Chapter 15
Pricing in Incomplete Models

We give an overview of various approaches to price derivatives in incomplete markets. In an incomplete market, there exists derivatives which cannot be exactly replicated. The second fundamental theorem of asset pricing (Theorem 13.3) states that if the market is arbitrage-free and complete, then there is only one equivalent martingale measure, and thus there is only one arbitrage-free price. When the arbitrage-free market is not complete, then there are many equivalent martingale measures, and thus there are many arbitrage-free prices.

In this chapter, we describe some approaches for choosing the equivalent martingale measure from a set of available equivalent martingale measures, in the case of an incomplete market. Chapter 16 is devoted to the study of quadratic hedging and pricing. In this chapter, we give only a short description of quadratic pricing, and concentrate to describe other methods.

Utility maximization provides a general method for the construction of an equivalent martingale measure. We show that the Esscher measure, which was used to prove the first fundamental theorem of asset pricing (Theorem 13.1), is related to the maximization of the expected utility, when the utility function is the exponential utility function. The concept of marginal rate of substitution provides a heuristic way to connect the utility maximization to the pricing of options. Minimizing the relative entropy between a martingale measure and the physical measure provides an equally natural way to construct an equivalent martingale measure. Again, we can show that minimizing the relative entropy is related to the maximizing the expected utility with the exponential utility function. It is of interest to compare the Esscher prices to the Black–Scholes prices: we see that the Esscher prices are close to the Black–Scholes prices for at-the-money calls, whereas for out-of-the-money calls the Esscher prices are lower.

We describe formulas for constructing an equivalent martingale measure by an absolutely continuous change of measure. These formulas are given for conditionally Gaussian returns, and for conditionally Gaussian logarithmic returns. An absolutely continuous change of measure shifts the market measure so that it becomes risk-neutral.

GARCH models provide good volatility predictions, and it is natural to ask whether GARCH models could be suitable for pricing options. We can apply the absolutely continuous change of measure to obtain an equivalent martingale measure in the GARCH market model. The standard GARCH(c015-math-001) model can be modified so that we obtain a model where the prices can be expressed almost in a closed form. We still need a numerical integration to compute the prices, but Monte Carlo simulation of price trajectories is not needed. The modified GARCH(c015-math-002) model was presented in Heston and Nandi (2000).

It is on interest to construct a nonparametric pricing method, and compare its properties to other methods. We construct a method which combines historical simulation, the Esscher measure, and conditioning on the current volatility.

An equivalent martingale measure can be deduced from the market prices of the options. This martingale measure could be called the implied martingale measure, because there is an analogy to the implied volatility. When the implied martingale measure is used, we have to assume that the market prices of the options are rational. This requires that we use option prices of liquid markets to estimate the implied martingale measure. The implied martingale measure which is deduced from the prices of liquid options can be used to price illiquid options.

Section 15.1 describes quadratic hedging and pricing. Section 15.2 describes pricing with the help of utility maximization. Section 15.3 considers pricing with the help of absolutely continuous changes of measures (Girsanov's theorem). Section 15.4 describes the use of a GARCH model in option pricing. Section 15.5 describes a method of nonparametric pricing which uses historical simulation. Section 15.6 discusses pricing with the help of estimating the risk-neutral density. Section 15.7 mentions quantile hedging.

Pricing in incomplete markets is studied in Duffie and Skiadas (1994), El Karoui and Quenez (1995), Karatzas (1996), and Gourioux et al. (1998). Bingham and Kiesel (2004, Chapter 7) discuss pricing in incomplete models, including mean–variance hedging and models driven by Lévy processes. Pricing of derivatives in the context of general econometric theory is presented in Magill and Quinzii (1996). Further references include Karatzas and Kou (1996).

15.1 Quadratic Hedging and Pricing

Quadratic hedging and pricing is discussed in detail in Chapter 16 (see also Föllmer and Schied, 2002, Definition 10.36, p. 393). At this point, we give a brief summary of the method.

Let us explain the idea of quadratic hedging using the case with one risky asset (c015-math-003) and two periods (c015-math-004). The initial wealth is c015-math-005 and the wealth obtained by trading with a bond and a stock is

equation

where c015-math-006 is the price of the bond, c015-math-007 is the price of the stock, c015-math-008 is the number of bonds in the portfolio, and c015-math-009 is the number of stocks in the portfolio. Our aim is to replicate the terminal value c015-math-010 of the contingent claim. We measure the quality of the approximation by the quadratic hedging error

equation

The minimization is done over self-financing trading strategies and over the initial wealth c015-math-011. The self-financing means that c015-math-012 and c015-math-013 satisfy

equation

The self-financing restriction connects the initial wealth c015-math-014 to the final wealth c015-math-015. The quadratic price is the initial wealth c015-math-016 that minimizes the quadratic hedging error.

The minimization is done easier when we use the value process, instead of the wealth process. Our final formulation for the general case c015-math-017 and c015-math-018 will be the following. In quadratic hedging, the quadratic hedging error

equation

is minimized among strategies c015-math-0191 and among the initial investment c015-math-022, where c015-math-023 is the discounted contingent claim. The terminal value of the gains process is defined by

equation

where c015-math-024 is the discounted price vector.

15.2 Utility Maximization

It turns out that an equivalent martingale measure can be found by looking at the portfolios which maximize the expected utility. Section 15.2.1 shows that the Esscher martingale measure is related to the maximization of the expected utility with the exponential utility function. Section 15.2.2 considers other utility functions. Section 15.2.3 shows that the Esscher measure is the equivalent martingale measure which minimizes the relative entropy with respect to the physical measure. Section 15.2.4 computes examples of Esscher prices. Section 15.2.5 discusses the heuristics of the marginal rate of substitution.

The use of Esscher transform and more general methods of utility maximization has to be combined with the estimation of the underlying distribution of the stock price process. This issue has been addressed in Bühlmann et al. (1996), Siu et al. (2004), Christoffersen et al. (2006), and Chorro et al. (2012), where parametric modeling was used. We address the issue in Section 15.5, where a nonparametric estimation is applied (see also Section 15.2.4).

15.2.1 The Exponential Utility

The Esscher transformation was applied in the proof of Theorem 13.1 (the first fundamental theorem of asset pricing) to construct an equivalent martingale measure in an arbitrage-free market. Let us recall the definition of the Esscher measure for the case of one risky asset. Let c015-math-025 be the discounted stock price, where c015-math-026. Let

equation

Let

equation

where c015-math-027. Let

be the unique finite minimizer of c015-math-029 over c015-math-030. Let c015-math-031 and

equation

for c015-math-032. We define the probability measure

Now c015-math-034, c015-math-035, and c015-math-036 for c015-math-037. Hence c015-math-038 is a martingale difference and c015-math-039 is a martingale with respect to c015-math-040.

The Esscher martingale measure is related to the maximization of the expected utility with the exponential utility function. The exponential utility function is defined as

equation

where c015-math-041 is the parameter of risk aversion. We want to maximize

equation

over self-financing trading strategies c015-math-042, where

equation

The maximization is equivalent to the minimization of

equation

We can write

equation

where c015-math-043. Minimization of

equation

over c015-math-044-measurable c015-math-045 is equivalent to the minimization of

equation

over c015-math-046. Thus, we arrive at the minimizer c015-math-047 in (15.1), and we can define an equivalent martingale measure (15.2) with the help of these minimizers.

15.2.2 Other Utility Functions

The construction of the Esscher martingale measure can be generalized to cover other utility functions. For example, consider the one period model with c015-math-048 risky assets and let c015-math-049 be the vector of discounted net gains:

equation

where c015-math-050 are the prices of the risky assets at time c015-math-051. Föllmer and Schied (2002, Corollary 3.10) states that if the market is arbitrage-free, and the utility function c015-math-052 and a maximizer c015-math-053 of c015-math-054 satisfy certain assumptions,2 then

defines an equivalent martingale measure c015-math-068.

The proof is based on the fact that a martingale measure has to satisfy c015-math-069, and this follows for the measure defined in (15.3) because the maximizer satisfies the first-order condition c015-math-070. Note that the existence of a maximizer is implied by Föllmer and Schied (2002, Theorem 3.3).

The Esscher density

is a special case of (15.3). Indeed, when the utility function is the exponential utility function c015-math-072, where c015-math-073 and c015-math-074 is the risk aversion, then (15.3) gives the density

equation

Portfolio c015-math-075 maximizes the expected utility c015-math-076 if and only if c015-math-077 minimizes the moment generating function c015-math-078. Thus, we obtain the density in (15.4), and the martingale measure c015-math-079 is independent of the risk aversion c015-math-080.

Note that we have maximized c015-math-081 over c015-math-082, which is not the same as maximizing the expected utility of the wealth. However, in the one-period case the wealth is written in (9.9) as

equation

Let c015-math-083 be a strictly increasing, strictly concave, and continuous utility function. We want to find c015-math-084 which maximizes c015-math-085 over all c015-math-086 such that c015-math-087 is c015-math-088-almost surely in the domain of c015-math-089. Define

equation

The original utility maximization is equivalent to the maximization of

equation

among all c015-math-090 such that c015-math-091, where c015-math-092 is the domain of c015-math-093.

15.2.3 Relative Entropy

The Esscher martingale measure was shown to be related to the maximization of the expected utility with the exponential utility function. We can also show that the Esscher measure can be obtained by minimizing the Kullback–Leibler distance to the physical market measure.

The closeness of probability distributions can be measured by the relative entropy (the Kullback–Leibler distance). The relative entropy of a probability measure c015-math-094 with respect to a probability measure c015-math-095 is defined as

equation

when c015-math-096 dominates c015-math-097. When c015-math-098 does not dominate c015-math-099, then we define c015-math-100.

Föllmer and Schied (2002, Corollary 3.25) states the following result for the one-period model with c015-math-101 risky assets: When the market model is arbitrage-free, then there exists a unique equivalent martingale measure c015-math-102, which minimizes the relative entropy c015-math-103 over all c015-math-104, where c015-math-105 is the set of equivalent martingale measures. Furthermore, the density of c015-math-106 is the Esscher density

equation

where c015-math-107 is the minimizer of the moment generating function c015-math-108, and c015-math-109 is the vector of discounted net gains: c015-math-110, where c015-math-111 is the vector of prices at time c015-math-112, and c015-math-113 is the vector of prices at time c015-math-114.

15.2.4 Examples of Esscher Prices

We apply the data of S&P 500 daily prices, described in Section 2.4.1. We estimate the Esscher call prices when the time to expiration is c015-math-115 trading days. The estimation is done for the c015-math-116-period model. We apply nonsequential estimation: the Esscher measure is estimated using the complete time series, and then the prices are estimated using the complete time series together with the estimated Esscher measure. We take the risk-free rate c015-math-117.

We denote the observed historical prices by c015-math-118. We construct c015-math-119 sequences of prices:

equation

where

equation

Each sequence has length c015-math-120, and the initial prices are c015-math-121. We apply nonoverlapping sequences, and restrict ourselves to the values c015-math-122, c015-math-123.

Let us compute differences

for c015-math-125 and c015-math-126. These differences are a sample of identically distributed observations of price increments c015-math-127. Note that price increments are not a stationary time series when the time period is long; see Figure 5.2(a). However, in our construction we have made c015-math-128 sequences of prices, each price sequence starts at 100, and thus the price differences make an approximately stationary sequence; see Figure 5.2(b) and (c).

Let us assume that the price increments are independent. Let c015-math-129 be the sample average of c015-math-130. Let c015-math-131 be the minimizer of c015-math-132 over c015-math-133. Let

equation

and

equation

The density of the martingale measure c015-math-134 with respect to underlying physical measure c015-math-135 of the price increments c015-math-136 is estimated by

equation

Let c015-math-137 be the payoff of a contingent claim. The price implied by the measure c015-math-138 is

We have constructed nonoverlapping price increments. The Esscher martingale measure was estimated using these price increments. Next we estimate the price (15.6) using a sample average. Let the contingent claim be c015-math-140. The estimate of price c015-math-141 is

equation

where c015-math-142 are the price differences in (15.5).

Figure 15.1 compares the Esscher call prices to the Black–Scholes prices. The time to maturity is 20 trading days. Panel (a) shows the call prices as a function of moneyness c015-math-143. The Esscher prices are shown with a red curve. The Black–Scholes prices are shown with a black curve. The volatility of the Black–Scholes prices is taken as the annualized sample standard deviation over the complete sample. Panel (b) shows the ratio of Black–Scholes prices to Esscher prices as a function of c015-math-144. We see that the Black–Scholes prices are less than the Esscher prices, except for the in-the-money calls. The result confirms with Figure 13.2(a), which shows that the Esscher density takes smaller values than the density of the Black–Scholes martingale measure for large increments.

Graphical representation of Esscher call prices compared to Black-Scholes prices.

Figure 15.1 Esscher call prices compared to Black–Scholes prices. (a) Esscher prices (red) and Black–Scholes prices (black) as a function of c015-math-145. (b) The ratio of Black–Scholes prices to Esscher prices.

15.2.5 Marginal Rate of Substitution

A martingale measure can be derived by using an argument based on marginal rate of substitution, as presented in Davis (1997) or in Cochrane (2001).

The value at time c015-math-146, obtained by a self-financing trading, is written in (13.8) as

equation

Let us denote by

equation

the value which is obtained when the initial value is c015-math-147, and the self-financing trading strategy is c015-math-148. The objective is to maximize

equation

over self-financing trading strategies c015-math-149, where c015-math-150 is a utility function. Utility functions are discussed in Section 9.2.2. The fair price of a derivative could be defined to be such that diverting a little of funds into the derivative has a neutral effect on the investor's achievable utility. Let c015-math-151 be a discounted contingent claim. Let us denote

equation

where c015-math-152 is the price of c015-math-153. We define the fair price of c015-math-154 at time c015-math-155 to be the solution c015-math-156 of the equation

equation

It can be proved that

equation

where c015-math-157 is the maximizer of c015-math-158, and

equation

Here we assume that c015-math-159 is differentiable at each c015-math-160 and that c015-math-161. Now

equation

is called the stochastic discount factor (pricing kernel, change of measure, state-price density). We have a single discount factor which is pricing each different asset. The price reflects riskiness and the riskiness depends on the covariance of c015-math-162 with the pricing kernel, and not the variance. An asset that does badly in recession is less desirable than an asset that does badly in boom, when the assets are otherwise similar.

15.3 Absolutely Continuous Changes of Measures

Girsanov's theorem gives a formula for changing the physical market measure to an equivalent martingale measure. First, we describe formulas for changing the measure when the returns are conditionally Gaussian. Second, we consider conditionally Gaussian logarithmic returns. Note that a version of Girsanov's theorem in continuous time is given in (5.64).

15.3.1 Conditionally Gaussian Returns

Let us assume that the excess returns

equation

are conditionally Gaussian. We assume that they satisfy

equation

for c015-math-163, where

c015-math-165 is c015-math-166-measurable, and c015-math-167 and c015-math-168 are predictable. The notation in (15.7) means that the conditional distribution of c015-math-169, conditional on c015-math-170, under probability measure c015-math-171, is the standard normal distribution. Here we assume that c015-math-172, and c015-math-173. We assume that c015-math-174. The assumption on c015-math-175 implies that c015-math-176 is a sequence of independent random variables with c015-math-177.

15.3.1.1 The Martingale Measure

The equivalent martingale measure c015-math-178 is such that under c015-math-179, the excess returns satisfy

15.8 equation

where c015-math-181 are i.i.d. with c015-math-182.

15.3.1.2 Construction of the Martingale Measure

Let

equation

for c015-math-183. Now c015-math-184 and c015-math-185. We define the probability measure c015-math-186 on c015-math-187 by

Measure c015-math-189 is an equivalent martingale measure: the process c015-math-190 of discounted prices is a martingale.

The fact that c015-math-191 is a martingale measure follows from

which is equivalent to

equation

The proof can be found in Shiryaev (1999, p. 443). Let us give the main steps of the proof. Let c015-math-193 and c015-math-194 be the imaginary unit. Then,3

c015-math-199-almost surely, since

equation

and

equation

Equation (15.11) shows that the characteristic function of c015-math-200 under c015-math-201 is the characteristic function of c015-math-202, which leads to (15.10).

15.3.1.3 The Relation to the Esscher Measure

The Esscher transformation leads to the same equivalent martingale measure as the absolutely continuous change of measure, in the special case of Gaussian returns with unit variance. We consider the case of one risky asset. Let

equation

where c015-math-203 is the discounted stock price, c015-math-204, and c015-math-205. Now

equation

where c015-math-206. Then c015-math-207 minimizes c015-math-208 over c015-math-209. We have

equation

The Esscher martingale measure is c015-math-210, where c015-math-211, which is the same measure as (15.9) for c015-math-212.

15.3.2 Conditionally Gaussian Logarithmic Returns

Let us assume that the excess logarithmic returns are conditionally Gaussian:

equation

where c015-math-213, c015-math-214, and c015-math-215 satisfy the same assumptions as in the case of conditionally Gaussian excess returns. This implies that c015-math-216 is a sequence of independent random variables with c015-math-217. Here c015-math-218 is the discounted stock price, so that c015-math-219 is the excess logarithmic return:

equation

15.3.2.1 The Martingale Measure

The equivalent martingale measure c015-math-220 is such that under this measure the excess logarithmic returns satisfy

where c015-math-222 are i.i.d. with c015-math-223.

15.3.2.2 Construction of the Martingale Measure

Let

equation

for c015-math-224. Let us define measure c015-math-225 by

equation

It can be proved that c015-math-226 is an equivalent martingale measure.

We derive measure c015-math-227 using the approach of Shiryaev (1999, p. 449). Let us assume that

equation

where c015-math-228 has the form

equation

and c015-math-229 are c015-math-230-measurable. Let us choose c015-math-231 so that the discounted price c015-math-232 is a martingale, where

equation

Variables c015-math-233 must be such that

Indeed, we need to have

equation

We have

equation

Equality in (15.13) is equivalent to

equation

which is equivalent to

equation

Thus,

equation

and

equation

15.4 GARCH Market Models

When the market model is a GARCH model, then we can apply the absolutely continuous changes of measures of Section 15.3 to derive an equivalent martingale measure. We consider first the Heston–Nandi method, which applies numerical integration to compute the expectation with respect to the equivalent martingale measure. Second, we consider the method of Monte Carlo simulation for the computation of the expectation. Third, we compare the risk-neutral densities of the Heston–Nandi model and GARCH(c015-math-235) model.

The GARCH(c015-math-236) model is defined in (5.38). Pricing under GARCH models was considered in Duan (1995). The Heston–Nandi model was presented in Heston and Nandi (2000).

A related approach was followed in Aït-Sahalia et al. (2001), where a complete continuous time diffusion model was postulated for the stock price, the volatility was estimated, and the Girsanov's theorem was applied to obtain the risk-neutral measure, which can be used for pricing.

Chorro et al. (2012) consider GARCH models where the innovations follow a generalized hyperbolic distribution, instead of the standard normal distribution. They construct the equivalent martingale measure as in Gerber and Shiu (1994), by choosing the density of the martingale measure (with respect to the physical measure) to have an exponential affine parametrization.

15.4.1 Heston–Nandi Method

We follow Heston and Nandi (2000) and consider model (5.50). We assume that the logarithmic returns satisfy

equation

where c015-math-237 is the constant risk-free rate, c015-math-238 is predictable, c015-math-239 are i.i.d. c015-math-240, and

equation

where c015-math-241 is the skewness parameter. Because we assume conditionally Gaussian logarithmic returns, then (15.12) implies that there exists an equivalent martingale measure c015-math-242 which is such that under c015-math-243

equation

15.4.1.1 Prices of Call Options

The prices of call options have an expression which can be computed using numerical integration. Let

equation

be the payoff of a call option, and

equation

be the discounted payoff of the call option. It follows from Theorem 13.2 that an arbitrage-free price of the call option is

equation

The expectation c015-math-244 can be written in terms of the characteristic function. Let c015-math-245 be the characteristic function of c015-math-246, under the condition that the stock price at time 0 is c015-math-247, where

equation

is the moment generating function, as defined in (5.53). The formula of c015-math-248 involves the stock price c015-math-249, the interest rate c015-math-250, and time c015-math-251 to expiration. Then

where c015-math-253 denotes the real part of a complex number c015-math-254. The expression in (15.14) is not a closed form expression, but we need numerical integration to compute the values. Also the values of the moment generating function are computed using a recursive formula. The price (15.14) is analogous to the Black–Scholes price in (14.58), where the cumulative distribution function c015-math-255 of the standard normal distribution does not have a closed form expression.

Let us prove (15.14). Let c015-math-256 be the density of c015-math-257 with respect to Lebesgue measure. Let

equation

Now c015-math-258 is a density function because c015-math-259 and c015-math-260. Note that

equation

The moment generating function of c015-math-261 is

equation

Now,

equation

The characteristic function of c015-math-262 is c015-math-263. Thus,

equation

see Billingsley (2005, Theorem 26.2, p. 346). The characteristic function of c015-math-264 is c015-math-265, and a similar formula is obtained for c015-math-266. We have proved (15.14).

Figure 15.2 shows Heston–Nandi GARCH(c015-math-267) call prices divided by the Black–Scholes call prices as a function of the moneyness. Parameters c015-math-268, c015-math-269, c015-math-270, and c015-math-271 are estimated from S&P 500 daily data, described in Section 2.4.1. In panel (a), time to expiration is 20 trading days, and the annualized volatility takes values c015-math-272 (black), c015-math-273 (red), and c015-math-274 (blue). The solid lines have c015-math-275, which is about equal to the value estimated from the S&P 500 data. The dashed lines have c015-math-276. In panel (b), the annualized volatility is c015-math-277, and the expiration time takes values 5 days (black), 20 days (red), and 40 days (blue). The risk-free rate is c015-math-278.

Panel (a) shows that the Heston–Nandi prices are lower than the Black–Scholes prices when the volatility is high, but when the volatility is low, then the Heston–Nandi prices are higher than the Black–Scholes prices. When the moneyness increases then the ratio of prices approaches one. Panel (b) shows that when the time to expiration becomes shorter, then the ratio of Heston–Nandi prices to the Black–Scholes prices increases. The skewness parameter c015-math-279 has a smaller influence than the volatility and the time to expiration.

Graphical representation of The ratios of Heston-Nandi to Black-Scholes prices.

Figure 15.2 The ratios of Heston–Nandi to Black–Scholes prices. Shown are the Heston–Nandi call prices divided by the Black–Scholes prices as a function of moneyness c015-math-280. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-281 (black), c015-math-282 (red), and c015-math-283 (blue). (b) The annualized volatility is c015-math-284. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue). The solid lines have c015-math-285 and the dashed lines have c015-math-286.

15.4.1.2 Hedging Coefficients of Call Options

The hedging coefficient of a call option is

equation

where c015-math-287 is the strike price and c015-math-288 is the moment generating function of c015-math-289, as defined in (5.53). The formula of c015-math-290 involves the stock price c015-math-291, the interest rate c015-math-292, and time c015-math-293 to expiration. Here c015-math-294 is the number of trading days to the expiration.

Figure 15.3 plots the ratios of Heston–Nandi GARCH(c015-math-295) call hedging coefficients to Black–Scholes call hedging coefficients as a function of the moneyness. We have the same setting as in Figure 15.2. We see from panel (a) that for a moderate and large volatility the Heston–Nandi deltas are smaller for out-of-the-money options, and larger for in-the-money options, than the Black–Scholes deltas. For small volatility, the behavior is opposite. We see from panel (b) that the time to expiration has a similar kind of effect as the volatility.

Graphical representation of Ratios of Heston-Nandi deltas to Black-Scholes deltas.

Figure 15.3 Ratios of Heston–Nandi deltas to Black–Scholes deltas. Shown are Heston–Nandi call hedging coefficients divided by the Black–Scholes hedging coefficients as a function of moneyness c015-math-296. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-297 (black), c015-math-298 (red), and c015-math-299 (blue). (b) The annualized volatility is c015-math-300. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue). The solid lines have c015-math-301 and the dashed lines have c015-math-302.

15.4.1.3 Hedging Errors of Call Options

Figure 15.4 shows hedging errors for hedging a call option with moneyness c015-math-303.4 We use S&P 500 daily data, described in Section 2.4.1. Panel (a) shows tail plots and panel (b) shows kernel density estimates. We consider three cases: (1) The red plots show the case where the volatility in the Heston–Nandi formula is taken to be the current GARCH volatility in the Heston–Nandi model. (2) The green plots show the case where the volatility is the stationary volatility in (5.49). (3) The blue plots show the case of Black–Scholes hedging with GARCH(c015-math-312) volatility. The parameters are estimated sequentially. Time to expiration is 20 days. The hedging is done twice: at the beginning and at the 10th day. The data is divided into 20 days periods using nonoverlapping sequences. The risk-free rate is c015-math-313. The hedging is started after obtaining 8 years (2000 days) of observations. We see that the Black–Scholes hedging leads to a better tail distribution of the hedging errors: the losses are smaller and the gains are larger. Note that the hedging errors of Black–Scholes hedging look different than in Figure 14.24, because in Figure 14.24 the hedging is done daily, and overlapping sequences are used.

Graphical representation of Heston-Nandi hedging errors.

Figure 15.4 Heston–Nandi hedging errors. (a) Tail plots; (b) kernel density estimates of hedging errors. We show cases (1) the Heston–Nandi hedging with Heston–Nandi volatility (red), (2) the Heston–Nandi hedging with stationary volatility (green), and (3) the Black–Scholes hedging with GARCH(c015-math-314) volatility (blue).

15.4.2 The Monte Carlo Method

We assume that the excess returns follow a shifted GARCH(c015-math-315) model. The assumption that the excess logarithmic returns c015-math-316 follow a shifted GARCH(c015-math-317) model leads to similar prices, and we do not show results for this case.

It is assumed that

equation

where

equation

c015-math-318, c015-math-319, and c015-math-320 are i.i.d. with the standard normal distribution c015-math-321. Under the martingale measure c015-math-322, obtained by an absolutely continuous change of measure, the excess returns can be written as

Pricing under measure c015-math-324 can be done by estimating

equation

where c015-math-325 is the discounted contingent claim. We simulate c015-math-326 sequences

and use the estimate

equation

where c015-math-328 is the value of the discounted contingent claim for the trajectory c015-math-329. We consider call options, so that c015-math-330.

We need to simulate trajectories c015-math-331 under the dynamics in (15.15). This can be done by simulating sequences c015-math-332 of excess returns and sequences c015-math-333 of risk-free returns. The sequence of stock prices is obtained as

equation

In our simulation, we assume that the risk-free rates are zero, so that the risk-free gross returns are c015-math-334. To start the simulation we choose c015-math-335 to be the current GARCH(c015-math-336) volatility, and c015-math-337.

Figure 15.5 shows Monte Carlo approximations of call prices divided by the Black–Scholes price as a function of the number c015-math-338 of Monte Carlo samples. In panel (a), the moneyness is c015-math-339, and in panel (b), the moneyness is c015-math-340. The black curves show the GARCH(c015-math-341) prices, and the red curves show the Heston–Nandi GARCH(c015-math-342) prices. The red horizontal line shows the Heston–Nandi prices computed using the closed form expression (15.14). Time to expiration is 20 trading days. We have applied data of S&P 500 daily prices, described in Section 2.4.1. The initial standard deviation is the sample standard deviation of S&P 500 returns, and the GARCH(c015-math-343) parameters are estimated using S&P 500 data. The Black–Scholes volatility is the annualized sample standard deviation. The risk-free rate is c015-math-344. We see that the GARCH(c015-math-345) price is lower than the Black–Scholes price when the moneyness is one, and the GARCH(c015-math-346) price is higher than the Black–Scholes price when the moneyness is 0.95. The Heston–Nandi prices are lower than the prices in the standard GARCH(c015-math-347) model.

Graphical representation of Monte Carlo approximation of GARCH prices.

Figure 15.5 Monte Carlo approximation of GARCH prices. Approximations of call prices divided by the Black–Scholes price as a function of the number of Monte Carlo samples. (a) Moneyness is c015-math-348. (b) Moneyness is c015-math-349. We show the GARCH(c015-math-350) price ratios (black) and the Heston–Nandi GARCH(c015-math-351) price ratios (red), and the red horizontal line shows the Heston–Nandi price ratio computed using (15.14).

Figure 15.6 shows standard GARCH(c015-math-352) call prices divided by the Black–Scholes call prices as a function of the moneyness. Parameters c015-math-353, c015-math-354, and c015-math-355 are estimated from S&P 500 daily data, described in Section 2.4.1. In panel (a), time to expiration is 20 trading days, and the annualized volatility takes values c015-math-356 (black), c015-math-357 (red), and c015-math-358 (blue). In panel (b), the annualized volatility is c015-math-359, and the expiration time takes values 5 days (black), 20 days (red), and 40 days (blue). The risk-free rate is c015-math-360.

Panel (a) shows that the standard GARCH(c015-math-361) prices are lower than the Black–Scholes prices when the volatility is high, but when the volatility is low, then the standard GARCH(c015-math-362) prices are higher than the Black–Scholes prices. When the moneyness increases then the ratio of prices approaches one. Panel (b) shows that when the time to expiration becomes shorter, then the ratio of the standard GARCH(c015-math-363) prices to the Black–Scholes prices increases.

Image described by caption and surrounding text.

Figure 15.6 The ratios of standard GARCH(c015-math-364) prices to Black–Scholes prices. Shown are the standard GARCH(c015-math-365) call prices divided by the Black–Scholes price as a function of moneyness c015-math-366. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-367 (black), c015-math-368 (red), and c015-math-369 (blue). (b) The annualized volatility is c015-math-370. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue).

Figure 15.7 shows standard GARCH(c015-math-371) call prices divided by the Heston–Nandi call prices as a function of the moneyness. The setting is the same as in Figure 15.6.

Image described by caption and surrounding text.

Figure 15.7 The ratios of standard GARCH(c015-math-372) prices to Heston–Nandi prices. Shown are the standard GARCH(c015-math-373) call prices divided by the Heston–Nandi prices as a function of moneyness c015-math-374. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-375 (black), c015-math-376 (red), and c015-math-377 (blue). (b) The annualized volatility is c015-math-378. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue).

15.4.3 Comparison of Risk-Neutral Densities

We study the risk-neutral distributions of the Heston–Nandi model when the parameters change. Also, we compare the risk-neutral distributions of the Heston–Nandi model to the risk-neutral distributions of the standard GARCH(c015-math-379) model.

In the GARCH models, the physical distribution of stock prices is given by

equation

where c015-math-380. The volatility c015-math-381 is defined differently in the standard GARCH(c015-math-382) model and in the Heston–Nandi model. A risk-neutral distribution of stock prices is given by

equation

where c015-math-383. We can estimate the distribution of c015-math-384 by first estimating the parameters of the model, second generating c015-math-385 Monte Carlo trajectories which give c015-math-386 observations c015-math-387, and finally using a kernel density estimator. We use S&P 500 daily data, described in Section 2.4.1.5

Figure 15.8 shows the estimated risk-neutral densities. Panel (a) compares the Heston–Nandi model with the standard GARCH(c015-math-391) model. The red curve shows the risk-neutral distribution of the Heston–Nandi model, the black curve shows the risk-neutral distribution of the standard GARCH(c015-math-392) model, and the green curve shows the risk-neutral distribution of the Black–Scholes model.6 Panel (b) studies the effect of the skewness parameter c015-math-397. The red curve shows the case c015-math-398 about seven, which is the value estimated from data. The orange curve shows the case c015-math-399, but it is very close to the case c015-math-400. The blue curve shows the case c015-math-401. We see that the green density is skewed so that large negative returns and moderate positive return are more probable than in the case of c015-math-402.

Graphical representation of Risk-neutral densities.

Figure 15.8 Risk-neutral densities. (a) Heston–Nandi model (red), standard GARCH(c015-math-403) model (black), and the Black–Scholes model (green). (b) Heston–Nandi risk-neutral densities for c015-math-404 (red), c015-math-405 (orange), and c015-math-406 (blue).

Figure 15.9 shows the risk-neutral densities of Figure 15.8 divided by the Black–Scholes risk-neutral density, which is shown with green in Figure 15.8(a). Panel (a) shows the Heston–Nandi (red) and standard GARCH(c015-math-407 (black) risk-neutral densities divided by the Black–Scholes risk-neutral density. Panel (b) shows the Heston–Nandi risk-neutral densities for c015-math-408 (red), c015-math-409 (orange), and c015-math-410 (blue), divided by the Black–Scholes risk-neutral density. Panel (a) shows that the risk-neutral densities in the GARCH models take higher values at the center than the Black–Scholes risk-neutral density, and the ratio has a hat shape. Panel (b) shows that the densities are close to each other for c015-math-411 and c015-math-412, but when c015-math-413, then the skewness is visible.

Graphical representation of Risk-neutral densities: Ratios.

Figure 15.9 Risk-neutral densities: Ratios. (a) Heston–Nandi (red) and standard GARCH(c015-math-414) (black) risk-neutral density ratios. (b) Heston–Nandi risk-neutral density ratios for c015-math-415 (red), c015-math-416 (orange), and c015-math-417 (blue).

15.5 Nonparametric Pricing Using Historical Simulation

It is interesting to compare nonparametric pricing to the Black–Scholes pricing and to the GARCH pricing methods. We define nonparametric pricing combining three elements: (1) historical simulation, (2) Esscher transformation, (3) conditioning with the current volatility, which is taken to be the GARCH(c015-math-418) volatility.

We have used Monte Carlo simulation to create sequences of prices in (15.16). Analogously, historical prices can be used to create sequences of prices. The Esscher transformation was applied in the proof of Theorem 13.1 (the first fundamental theorem of asset pricing) to construct an equivalent martingale measure in an arbitrage-free market. In Section 15.2, the Esscher transformation was shown to be related to utility maximization.

We consider c015-math-419-period model, with time to expiration being c015-math-420 trading days. Our price data is c015-math-421. We construct c015-math-422 sequences of prices:

equation

where

equation

Each sequence has length c015-math-423, and the initial prices are c015-math-424.

15.5.1 Prices

We define first the unconditional price and then the price which conditions on the current volatility.

The unconditional price is computed by

equation

where c015-math-425 is the value of the discounted contingent claim for the trajectory c015-math-426, and c015-math-427 is the estimated Esscher density. For example, in the case of a call option, c015-math-428. The estimated Esscher density is c015-math-429 with

equation

where c015-math-430 and

equation

Value c015-math-431 is the minimizer of

equation

over c015-math-432.

Next we define the conditional price. Let c015-math-433 be the current estimated GARCH(c015-math-434) volatility. The price is estimated as

equation

where c015-math-435 is the value of the discounted contingent claim for the trajectory c015-math-436, and c015-math-437 is the Esscher density. The weight is defined as

equation

where c015-math-438 is the estimated GARCH(c015-math-439) volatility at day c015-math-440. Furthermore, c015-math-441 is the scaled kernel

equation

where c015-math-442 is a kernel function. We choose

equation

We apply the S&P 500 daily data, described in Section 2.4.1.

Figure 15.10 shows nonparametric call prices divided by the Black–Scholes call price as a function of the smoothing parameter. In panel (a) moneyness c015-math-443, and in panel (b) c015-math-444. The annualized volatility takes values c015-math-445 (black), c015-math-446 (red), and c015-math-447 (blue). Time to expiration is c015-math-448 trading days, risk-free rate is c015-math-449. The volatility is the sample standard deviation. We see that the price ratios are higher for small volatility. The prices stabilize when c015-math-450.

Graphical representation of Nonparametric prices.

Figure 15.10 Nonparametric prices. Ratios of nonparametric call prices to the Black–Scholes price as a function of the smoothing parameter. (a) Moneyness is c015-math-451. (b) Moneyness is c015-math-452. The annualized volatility takes values c015-math-453 (black), c015-math-454 (red), and c015-math-455 (blue).

Figure 15.11 shows nonparametric call prices divided by the Black–Scholes call prices as a function of the moneyness c015-math-456. In panel (a), time to expiration is 20 trading days, and the annualized volatility takes values c015-math-457 (black), c015-math-458 (red), and c015-math-459 (blue). In panel (b), the annualized volatility is c015-math-460, and the expiration time takes values 5 days (black), 20 days (red), and 40 days (blue). The risk-free rate is c015-math-461.

Graphical representation of The ratios of nonparametric prices to Black-Scholes prices.

Figure 15.11 The ratios of nonparametric prices to Black–Scholes prices. Shown are the nonparametric call prices divided by the Black–Scholes prices as a function of moneyness c015-math-462. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-463 (black), c015-math-464 (red), and c015-math-465 (blue). (b) The annualized volatility is c015-math-466. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue).

15.5.2 Hedging Coefficients

We compute the nonparametric hedging coefficients by approximating the derivative of the price numerically. The numerical approximation is done by the difference quotient. The price is computed at the stock price c015-math-467 and c015-math-468, and the hedging coefficient is taken as

where c015-math-470 is small.

Figure 15.12 shows nonparametric hedging coefficients divided by the Black–Scholes delta as a function of the smoothing parameter. In panel (a) moneyness c015-math-471, and in panel (b) c015-math-472. Time to expiration is c015-math-473 trading days, risk-free rate is c015-math-474. The volatility is the sample standard deviation. Parameter c015-math-475 in (15.17) is taken as c015-math-476 (black, red, and blue). We see that when c015-math-477, then the nonparametric deltas are larger than the Black–Scholes deltas; when c015-math-478, then the nonparametric deltas are smaller than the Black–Scholes deltas.

Graphical representation of Nonparametric deltas.

Figure 15.12 Nonparametric deltas. Ratios of nonparametric call deltas to the Black–Scholes delta as a function of the smoothing parameter. (a) Moneyness is c015-math-479. (b) Moneyness is c015-math-480. Parameter c015-math-481 in (15.17) is taken as c015-math-482 (black, red, and blue).

Figure 15.13 shows nonparametric call deltas divided by the Black–Scholes call deltas as a function of the moneyness c015-math-483. In panel (a), time to expiration is 20 trading days, and the annualized volatility takes values c015-math-484 (black), c015-math-485 (red), and c015-math-486 (blue). In panel (b), the annualized volatility is c015-math-487, and the expiration time takes values 5 days (black), 20 days (red), and 40 days (blue). The risk-free rate is c015-math-488.

Graphical representation of The ratios of nonparametric deltas to Black-Scholes deltas.

Figure 15.13 The ratios of nonparametric deltas to Black–Scholes deltas. Shown are the nonparametric call deltas divided by the Black–Scholes deltas as a function of moneyness c015-math-489. (a) Time to expiration is 20 trading days. The annualized volatility takes values c015-math-490 (black), c015-math-491 (red), and c015-math-492 (blue). (b) The annualized volatility is c015-math-493. The time to expiration takes values 5 days (black), 20 days (red), and 40 days (blue).

Figure 15.14 shows hedging errors of nonparametric hedging. Panel (a) shows tail plots of the empirical distribution function and panel (b) shows kernel density estimates. Time to expiration is c015-math-494 trading days. The blue curves show the case where hedging is done once, and the red curves show the case where hedging is done twice. We show also Black–Scholes hedging errors: in the green curves, hedging is done once and in the violet curves, the hedging is done twice. The risk-free rate is c015-math-495. The volatility is the sample standard deviation. The smoothing parameter is c015-math-496. Panel (a) shows that the Black–Scholes hedging performs better in the tails, but panel (b) shows that the nonparametric hedging performs better in the central area.

Graphical representation of Nonparametric hedging errors.

Figure 15.14 Nonparametric hedging errors. (a) Tail plots; (b) kernel density estimates. Nonparametric hedging is done once (blue) and twice (red). The Black–Scholes hedging is done once (green) and twice (violet).

15.6 Estimation of the Risk-Neutral Density

Let us consider an European call option

equation

where c015-math-497 is the price of the stock at the expiration, and c015-math-498 is the strike price. We assume that the risk-free rate is c015-math-499. Theorem 13.2 states that arbitrage-free prices of an European option c015-math-500 can be written as

equation

where the expectation is with respect to an equivalent martingale measure c015-math-501. We assume that the distribution of c015-math-502 under c015-math-503 has density c015-math-504 with respect to the Lebesgue measure. The density depends on the initial stock price c015-math-505. The price of the call option can be written as

Differentiating with respect to c015-math-507, we get

equation

where we used the fact c015-math-508. Differentiating second time, we get

We can apply (15.19) to compute an approximation to the density c015-math-510, when prices are observed for several strike prices c015-math-511. Note that (15.18) shows that the problem can be considered as deconvolution problem, where the pricing function

equation

should be inverted in order to obtain c015-math-512.7

The density c015-math-529 is a risk-neutral distribution for c015-math-530. Similarly, we can consider estimating the risk-neutral distribution of c015-math-531 for c015-math-532. Thus, we are able to estimate all marginal distributions of the prices process c015-math-533, but not the complete risk-neutral distribution c015-math-534.

15.6.1 Deducing the Risk-Neutral Density from Market Prices

Let us denote by c015-math-535 the option price when the strike price is c015-math-536. The market prices provide values c015-math-537 for strike prices c015-math-538. These observations can be used to deduce the risk-neutral density of c015-math-539, implied by the market prices. The implied risk-neutral density can be estimated using liquid options, and the estimated density can be used to price illiquid options. It is possible that the market prices are not fair. On the other hand, market prices can incorporate information which is difficult to obtain by statistical procedures. For example, market prices can incorporate information about event risks, like information about the elections in the near future.

Aït-Sahalia and Lo (1998) considered the observations c015-math-540 as coming from a regression model

equation

where c015-math-541 is the true pricing function and c015-math-542 is random noise. They used semiparametric regression to estimate the true pricing function, and then took the second derivative to obtain the risk-neutral density. They considered the pricing function to have five arguments:

equation

where c015-math-543 is the stock price, c015-math-544 is the time to expiration, c015-math-545 is the risk-free rate for that maturity, and c015-math-546 is the dividend yield for the asset. They used two dimensional kernel regression to estimate the implied volatility, as a function of moneyness and time to maturity. Then they applied the Black–Scholes formula to obtain the pricing function.

Aït-Sahalia and Duarte (2003) estimated the pricing function using a combination of constrained univariate least squares regression and smoothing. The constrained regression is useful because the pricing function is increasing and convex as a function of the strike price. The convexity follows from (15.19), because c015-math-547 implies that the second derivative of the pricing function is nonnegative, which implies convexity.

15.6.2 Examples of Estimation of the Risk-Neutral Density

We can use risk-neutral densities to give insight about a pricing method. Some pricing methods are such that the risk-neutral density can be expressed in a closed form (Black–Scholes model), or we can simulate observations from the risk-neutral density and estimate the risk-neutral density based on the simulated observations (Heston–Nandi and the standard GARCH(c015-math-548) model). See Figures 15.8 and 15.9 for risk-neutral densities in the Black–Scholes, Heston–Nandi and the standard GARCH(c015-math-549) model. The nonparametric pricing using historical simulation is described in Section 15.5. The nonparametric pricing is such that the risk-neutral density is not easy to find directly, but it can be deduced from the prices, by inverting the information in the prices.

We compute the risk-neutral densities for four methods: (1) the Heston–Nandi pricing described in Section 15.4.1, (2) the GARCH(c015-math-550) pricing described in Section 15.4.2, (3) the nonparametric pricing using historical simulation described in Section 15.5, and (4) the Black–Scholes pricing.

Let us denote by c015-math-551 the price when the strike price is c015-math-552. Let us observe prices c015-math-553 for strike prices c015-math-554. The first derivative is approximated by

equation

for c015-math-555. The approximations of the second derivative give estimates of the density:

for c015-math-557.

The numerical differentiation which leads to the density in (15.20) can lead to a unsmooth density. We can smooth the density by using two-sided moving averages. The smoothing is done below in the case of the nonparametric pricing. We apply below the S&P 500 daily data of Section 2.4.1.

Figure 15.15 shows the price functions. Panel (a) shows pricing functions as a function of strike price and panel (b) shows the ratios of the prices to the Black–Scholes prices. In panel (a), the prices are for the standard GARCH(c015-math-558) pricing (black), for nonparametric pricing (orange), for Heston–Nandi version of GARCH(c015-math-559) pricing (red), and for the Black–Scholes pricing (dark green). In panel (b), the prices are divided by the Black–Scholes prices. The number of Monte Carlo samples in GARCH pricing is c015-math-560. The smoothing parameter of nonparametric GARCH pricing is c015-math-561. The volatility is the sample standard deviation in all three cases.

Graphical representation of Price functions.

Figure 15.15 Price functions. (a) The standard GARCH(c015-math-562) pricing (black), nonparametric pricing (orange), Heston–Nandi pricing (red), and Black–Scholes pricing (dark green). (b) The prices divided by the Black–Scholes prices.

Figure 15.16 shows densities of c015-math-563 under risk-neutral measures. Panel (a) shows densities with respect to the Lebesgue measure and panel (b) shows the ratios of the densities to the Black–Scholes density. In panel (a), the densities are for the standard GARCH(c015-math-564) pricing (black), for nonparametric pricing (orange), for Heston–Nandi version of GARCH(c015-math-565) pricing (red), and for the Black–Scholes pricing (dark green). In panel (b), the densities are divided by the Black–Scholes density.

Graphical representation of Risk-neutral densities.

Figure 15.16 Risk-neutral densities. (a) The standard GARCH(c015-math-566) pricing (black), nonparametric pricing (orange), Heston–Nandi pricing (red), and the Black–Scholes pricing (dark green). (b) The densities are divided by the Black–Scholes density.

15.7 Quantile Hedging

Let c015-math-567 be a discounted European contingent claim. A fair price of c015-math-568 could be considered as the initial investment c015-math-569 such that there exists a trading strategy which leads to value c015-math-570 which is close to c015-math-571. This is the basic idea of quadratic hedging, which minimizes c015-math-572 (see Section 15.1). However, from the point of view of the writer of the option it is desirable that c015-math-573. This leads to the definition of quantile hedging, where the probability of c015-math-574 is maximized.

Let c015-math-575 be a bound for the initial investment. We want to find a self-financing trading strategy whose value process maximizes the probability

equation

among all those strategies which satisfy

  1. 1. c015-math-576,
  2. 2. c015-math-577 for c015-math-578.

We have to assume that c015-math-579 is not too large, since letting c015-math-580 to be large would allow very expensive strategies. We assume that c015-math-581, where c015-math-582 is the set of arbitrage-free prices, as characterized in Theorem 13.2. Föllmer and Schied (2002, Section 8.1, p. 245) studies quantile hedging; see also Föllmer and Leukert (1999).

equation

where c015-math-021 is the set of square integrable random variables.

equation

c015-math-197-almost surely. Thus, the first equality follows from the definition of the conditional expectation. For more details, see Föllmer and Schied (2002, Proposition A.12, p. 405) and Shiryaev (1999, p. 438), where terms “conversion lemma” and “generalized Bayes' formula” are used for this equality.

equation

where

equation

the risk-free rate is c015-math-305, c015-math-306 is the price of the option, c015-math-307 is the terminal value of the option, c015-math-308 are the hedging coefficients, c015-math-309 are the stock prices, the current time is denoted by 0, the time to expiration is c015-math-310 days, and hedging is done daily. When hedging is done with a lesser frequency, then we use formula (14.79) for c015-math-311.

equation

where c015-math-394 are i.i.d. Thus, the risk-neutral distribution is the log-normal distribution

equation

where c015-math-395 is the number of trading days to the expiration. Parameter c015-math-396 is estimated by the sample standard deviation from the daily logarithmic returns.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset