Chapter 16
Quadratic and Local Quadratic Hedging

Quadratic hedging was introduced in Sections 13.1.3 and 15.1. In quadratic hedging we find the best approximation of the option in the sense of the mean-squared error. Quadratic hedging is related to the idea of statistical arbitrage: The fair price is defined as such price that makes the probability of gains and losses small for the writer of the option.

Quadratic hedging makes it possible to price and hedge options in a completely nonparametric way. In quadratic hedging we can derive prices and hedging coefficients without any modeling assumptions, making only some rather weak assumptions about square integrability and about a bounded mean–variance trade-off. There are many ways to implement quadratic hedging nonparametrically. We use kernel estimation in our implementation.

Let c016-math-001 be the discounted payoff of an European option. For example, for a call option c016-math-002. In quadratic hedging the mean-squared hedging error

equation

is minimized among strategies c016-math-003 and among the initial investment c016-math-004. The terminal value of the gains process is defined by

equation

where c016-math-005 is the discounted price vector. The problem resembles least-squares linear regression, where c016-math-006 is the response variable, c016-math-007 are the explanatory variables, c016-math-008 is the intercept, and c016-math-009 are the regression coefficients. However, now we have a time series setting, where the “explanatory” variable c016-math-010 is observed at time c016-math-011. We can use the knowledge of the observed values of c016-math-012 in choosing c016-math-013. In the usual linear regression all regression coefficients are chosen at the same time: c016-math-014.

Quadratic hedging in discrete time is explained in monographs Föllmer and Schied (2002, p. 393), Bouchaud and Potters (2003), and Černý (2004b). Earlier studies include Föllmer and Schweizer (1989), Bouchaud and Sornette (1994), Schweizer (1994), and Schäl (1994). Early continuous time studies include Föllmer and Sondermann (1986) and Duffie and Richardson (1991).

We study quadratic hedging and pricing in three steps: first for the one period model, then for the two period model, and finally the formulas are given for the general multiperiod model. The multiperiod model contains as special cases the one and two period models, but we think that it is helpful to study the one and two period models separately, because in these models the formulas are more transparent and notationally more convenient than in the multiperiod model. The generalization from the two period model to the multiperiod model is straightforward.

Local quadratic hedging simplifies the minimization problem of quadratic hedging, and it can achieve easier computations. In the one period model quadratic hedging and local quadratic hedging are equivalent, but in the multiperiod models they are different.

The price and the hedging coefficients of quadratic hedging do not have a closed-form expression, but only a recursive definition. This recursive definition can be used in computations, but the implementation is not trivial. We implement only the local quadratic hedging and pricing. We need to estimate various conditional expectations. We estimate the conditional expectations using historical simulation: The time series of previous returns is used to construct a large number of price sequences, and conditional expectations are estimated as sample means over price sequences. The observed volatility is used as the conditioning variable. Separate methods are used for the case of independent and dependent increments.

We evaluate quadratic hedging by studying the distribution of the hedging errors. The distribution should be concentrated around zero as well as possible. The main observation is that even the simplest setting of local quadratic hedging with independent increments leads to a distribution of the hedging errors that is better concentrated around zero than the distribution of the hedging errors when Black–Scholes hedging with GARCH(c016-math-015) volatility is used.

Section 16.1 studies global quadratic hedging and pricing. Section 16.2 studies local quadratic hedging and pricing. Section 16.3 studies implementations of local quadratic hedging.

16.1 Quadratic Hedging

The exact solution for quadratic hedging can be given using backward induction. We present the solution in three steps: first for the one period model, second for the two period model, and third for the general model.

16.1.1 Definitions and Assumptions

We recall the notation from Section 13.2, and in particular from Section 13.2.2. We assume that there is only one risky asset: c016-math-016. The price process of the riskless bond is denoted by c016-math-017. We choose

equation

where c016-math-018. The notation is a short hand for c016-math-019, where c016-math-020 is the time between two steps, expressed in fractions of a year, and c016-math-021 is the annual interest rate. The time series of prices of the risky asset is denoted by c016-math-022. The complete price vector is denoted by

equation

A trading strategy is

equation

where the values c016-math-023 and c016-math-024 express the quantity of the bond and the risky asset held between c016-math-025 and c016-math-026.

16.1.1.1 Wealth and Value Processes

The wealth at time 0 is c016-math-027, and after that

equation

Under the condition of self-financing the wealth at time c016-math-028 was written in (13.5) as

equation

where c016-math-029.

The discounted price process is defined by

equation

We denote c016-math-030. The value process was defined in (13.8) as

equation

Under the condition of self-financing the value at time c016-math-031 can be written as

16.1 equation

where c016-math-033.

The gains process is defined as

equation

For a self-financing strategy

16.1.1.2 Quadratic Hedging

We use the terms “quadratic hedging” and “global quadratic hedging” to mean the same thing. The term “global quadratic hedging” is used when a distinction to “local quadratic hedging” is emphasized. We use the term “quadratic price” to mean the price that is implied by quadratic hedging.

Let c016-math-035 be the value of an European option. For example, c016-math-036. Let c016-math-037. A quadratic strategy c016-math-038 is a minimizer of

equation

over c016-math-039 and over self-financing strategies c016-math-040. We obtain the quadratic price

equation

In the general case the quadratic price is c016-math-041, but in our case c016-math-042. The bond coefficients c016-math-043 are determined from the equations

as noted in (13.9). The complete quadratic hedging strategy, which includes both the quantities of bond and stock, is given by

equation

16.1.1.3 Mean Self-Financing

We have defined quadratic hedging as a hedging strategy that minimizes the mean-squared error among self-financing strategies. It is also possible to define a version of quadratic hedging where the mean-squared error is minimized among so-called mean self-financing strategies. We consider this approach only in Section 16.2.3, where local quadratic hedging without self-financing is discussed.

The self-financing condition in (13.4) states that

We can dispose the restriction to the self-financing strategies, and assume only that the strategies are mean self-financing. Assumption

equation

is equivalent with

equation

Let us make the mean self-financing assumption

where

equation

Note that if we define the cumulative cost process by

equation

where

equation

then

equation

Schäl (1994) considers quadratic hedging in discrete time with mean self-financing strategies. The definition of mean self-financing in continuous time was given in Föllmer and Sondermann (1986).1

We noted that the equality c016-math-049 in (16.2) holds only for self-financing strategies. Nevertheless, it is possible to minimize the mean-squared error

equation

over strategies c016-math-050, which are not necessarily self-financing. Note that c016-math-051.

We have that

equation

Under the condition (16.5) of mean self-financing we have that

equation

Thus,

equation

Thus, the term “variance optimal hedging” can be used in this case. Also,

equation

and it is natural to call c016-math-052 the fair price.

16.1.1.4 Assumptions

We have to assume the square integrability of the relevant terms:

equation

The assumptions can be written as c016-math-053 and c016-math-054. In addition, we have to restrict ourselves to the square integrable trading strategies. That is, the value and the gains processes are assumed to satisfy

We say that the bounded mean–variance trade-off holds if

c016-math-057-almost surely, for c016-math-058, where c016-math-059 is a constant. Föllmer and Schied (2002, Theorem 10.39, p. 395) consider the existence and uniqueness of a quadratic hedging strategy. They show that with c016-math-060 risky assets the bounded mean–variance trade-off guarantees the existence of a quadratic strategy (which they call a variance-optimal strategy). The strategy is unique up to modifications in the set c016-math-061.

16.1.2 The One Period Model

We consider the pricing of an European option in the single period model. In the single period model the underlying security has value c016-math-062 at the beginning of the period and value c016-math-063 at the expiration of the option. The price c016-math-064 is a fixed number and c016-math-065 is a random variable. At time zero the option price is c016-math-066. The value of the European option at the expiration is denoted by c016-math-067. For example, in the case of a call option c016-math-068, where c016-math-069 is the strike price. In the single period model the option is hedged only once (at time 0).

16.1.2.1 Pricing in the One Period Model

In the one period model

16.8 equation
equation

We want to find c016-math-071 and c016-math-072 minimizing

equation

This is the population version of a linear least-squares regression with the explanatory variable c016-math-073 and the response variable c016-math-074. We obtain the solutions2

and

The hedging coefficient can be written as

because constant c016-math-082 can be removed from the covariance and the variance, and the discounting factors of the nominator and denominator cancel each other. Note that for a call option c016-math-083 and for a put option c016-math-084.3

We see that the quadratic price

16.13 equation

is obtained by subtracting a correction term from the expected value of the payoff of the option.

The value process is useful in the multiperiod model, but in the single period model we can use the wealth process as well. We can use the following equivalent formulation. The initial wealth is c016-math-091. This amount is invested in the bank account. The amount c016-math-092 is borrowed at the risk-free rate and this money is invested in stock, so that there are c016-math-093 stocks in the portfolio. The value of the portfolio at time 1 is

equation

We want to find values of c016-math-094 and c016-math-095 that minimize

16.14 equation

16.1.2.2 Scatter Plots of Stock Price Differences and Option Payoffs

Figure 16.1 shows a scatter plot of points c016-math-097 together with linear fits. We use the daily data of S&P 500 prices, described in Section 2.4.1.

We take c016-math-098 and c016-math-099, where c016-math-100 are the gross returns of S&P 500 over 10 trading days. Values c016-math-101 are the payoffs of call options with strike price c016-math-102, and the expiration time 10 trading days. In panel (a) c016-math-103, and in panel (b) c016-math-104. The green lines show the least-squares linear fit c016-math-105, where c016-math-106 and

equation

where we use the sample means, variances, and covariances. The red lines show the linear fit c016-math-107, where c016-math-108 is the Black–Scholes price, and c016-math-109 is the Black–Scholes delta. The volatility is estimated by the sample standard deviation. We see that when c016-math-110, there is hardly any difference between the least-squares fit and the Black–Scholes fit. When c016-math-111, then the least-squares hedging coefficient is higher than the Black–Scholes delta.

Graphical representation of Linear approximations of call payoffs.

Figure 16.1 Linear approximations of call payoffs. We show scatter plots of c016-math-112 for (a) c016-math-113 and (b) c016-math-114. The green curves show the least-squares fit and the red curves show the Black–Scholes fit.

16.1.2.3 Pricing in the One Period Model Continued

We derive the solution (16.10), (16.11) in a different way. The different way of deriving the solution is such that it can be generalized to multiperiod models. Furthermore, it helps us to find the equivalent martingale measure and the hedging error.

We need to find c016-math-115 and c016-math-116 minimizing

equation

where c016-math-117 and c016-math-118. We can write

For a fixed c016-math-120 the minimizer over c016-math-121 is

We have that

where we denote

and

equation

We see from (16.17) that the mean-squared error is minimized by choosing c016-math-125. Equation (16.16) implies that the optimal hedging coefficient is

where c016-math-127 is defined in (16.18). The formulas (16.18) and (16.19) are equivalent with the formulas (16.10) and (16.11).4

Note that formulas (16.18) and (16.19) define c016-math-128 in terms of c016-math-129, whereas (16.10) and (16.11) define c016-math-130 in terms of c016-math-131.

16.1.2.4 The Martingale Measure in the One Period Model

Let us assume that our one period model is arbitrage-free. Theorem 13.1 implies that there exists an equivalent martingale measure. Theorem 13.2 implies that any arbitrage-free price can be written as an expected value c016-math-132 for some equivalent martingale measure c016-math-133. Let us find the martingale measure c016-math-134, which is implied by the quadratic hedging.

The density of c016-math-135 with respect to the underlying physical measure is obtained from (16.18) as

where

equation

and

Now, we have found a measure c016-math-138 such that

equation

Note that in our notation c016-math-139.

The Martingale Measure for S&P 500 in the One Period Model

Let us estimate the equivalent martingale measure associated with quadratic hedging using S&P 500 daily data of Section 2.4.1.

A time series of increments c016-math-140 is not approximately stationary when the time series covers a long time period; see Figure 5.2. The time series of gross returns is nearly stationary, even when the time series extends over a long time period; see Figure 2.1(b). Thus, we can use historical simulation to create a time series of increments from the excess gross returns.

We take the interest rate c016-math-141. We consider a one-step model with the step of 20 days. The excess gross return is equal to the net return

equation

Now,

equation

Let

equation

be the increment. Our S&P 500 data provides a sample of identically distributed observations c016-math-142 from the distribution of c016-math-143. We use non-overlapping increments.

Let us estimate the density

equation

of the martingale measure with respect to underlying physical measure of c016-math-144. The estimate is

equation

where c016-math-145 and c016-math-146 are estimates of c016-math-147 and c016-math-148 in (16.21), obtained by replacing the expectations and variances by sample averages and sample variances.

The underlying physical density of c016-math-149 with respect to the Lebesgue measure can be estimated using the kernel estimate c016-math-150 of (3.43). The density of the martingale measure with respect to the Lebesgue measure can be estimated as

equation

Figure 16.2(a) shows the estimate c016-math-151 of the density of the martingale measure with respect to the physical measure (dark green). The red curve shows the case of the Esscher measure and the blue curve the case of the Black–Scholes measure. The blue curve shows the density of the risk-neutral log-normal density with respect to the estimated physical measure. These are taken from Figure 13.1. Panel (b) shows the density c016-math-152 (dashed dark green) and c016-math-153 (solid dark green). We show also the physical density (solid blue) and the risk-neutral density (dashed blue) in the Black–Scholes model.

Graphical representation of Martingale measure in the one-step model.

Figure 16.2 Martingale measure in the one period model. (a) The density of the quadratic martingale measure with respect to the physical measure (dark green), Esscher measure (red), and Black–Scholes measure (blue). (b) The kernel density estimate of the physical measure (solid dark green), and the corresponding quadratic martingale measure (dashed dark green). The log-normal physical measure and the corresponding risk-neutral log-normal density are depicted as solid blue and dashed blue curves, respectively.

The Martingale Measure in the One Period Binomial Model

Let us study the one period binomial model, as defined in Section 14.2.1. In this model at time 0 the stock has value c016-math-154, and at time 1 the stock can take values c016-math-155 and c016-math-156, where c016-math-157. The probability of the up movement is c016-math-158 and the probability of the down movement is c016-math-159 with c016-math-160. Let us denote

equation

We have that

equation

From (16.20) we obtain that the martingale measure c016-math-161 satisfies

equation

where c016-math-162, c016-math-163, and c016-math-164 or c016-math-165. We have that

equation

and

equation

Thus,

equation

which is equal to the martingale measure already derived in (14.18). In fact, the binomial model is a complete model and there is only one equivalent martingale measure.

16.1.3 The Two Period Model

We consider pricing and hedging of an European option in the two period model. The general multiperiod model is considered in Section 16.1.4, and this presentation includes the two period model as a special case. However, we think that it is easier to read the presentation of the multiperiod model when the two period model is presented first.

In the two period model the underlying security takes values c016-math-166, c016-math-167, and c016-math-168. The price c016-math-169 is a fixed number and c016-math-170 and c016-math-171 are random variables. The option is written at time 0, and it expires at time 2. Hedging is done at times 0 and 1 by choosing the hedging coefficients c016-math-172 and c016-math-173. The value of the European option at the expiration is denoted by c016-math-174. For example, in the case of a call option c016-math-175, where c016-math-176 is the strike price.

16.1.3.1 An Introduction to the Minimization Problem

The minimization problem can be solved either using the value process or by using the wealth process. The use of the value process is more convenient.

The Minimization Using the Value Process

In the two period model the value process and the discounted contingent claim are defined as

equation

where c016-math-177, c016-math-178, c016-math-179, and c016-math-180. We want to find c016-math-181 minimizing

Notation c016-math-183 means the unconditional expectation with respect to the underlying measure c016-math-184, and we denote by c016-math-185, the conditional expectation, with respect to sigma-algebra c016-math-186:

equation

Unlike in the one period model this minimization problem cannot be considered as a usual population version of a linear least squares regression. We can consider c016-math-187 and c016-math-188 as explanatory variables and c016-math-189 as the response variables, but now intercept c016-math-190 and coefficient c016-math-191 are chosen at time 0, and coefficient c016-math-192 is chosen at time 1. In the usual regression problem all parameters are chosen at time 0.

The minimization problem can be solved in the following way. First, we find c016-math-193 minimizing

equation

Let the minimizer be c016-math-194. The notation indicates that the minimizer depends on c016-math-195 and c016-math-196. Second, we find c016-math-197 and c016-math-198 minimizing

equation
The Minimization Using the Wealth Process

The wealths at times 0, 1, and 2 are

equation

where c016-math-199, c016-math-200, c016-math-201, c016-math-202, c016-math-203, and c016-math-204. We want to find c016-math-205 so that

equation

is minimized, under the self-financing constraints. The minimization problem can be solved in the following way. First, we find c016-math-206 minimizing

equation

under the self-financing constraint

equation

Let the minimizer be c016-math-207. The notation indicates that the minimizer depends on c016-math-208. Second, we find c016-math-209 minimizing

equation

We can now see why it is easier to solve the problem using the value process: It is possible to apply unconstrained minimization when the value process is used.

16.1.3.2 Solving the Minimization Problem

Let us solve the problem of minimizing (16.22). We have that

equation

Since

equation

the minimizer over c016-math-210 is

We have that

where we denote

equation

and

equation

It holds that

equation

Similar calculations which lead to (16.23) show that the minimizer over c016-math-214 is

Finally, we have to find c016-math-216 minimizing

equation

The minimizer over c016-math-217 is

where

equation

Indeed, similar calculations which lead to (16.24) show that

equation

where

equation

We can summarize the results in the following proposition.

Note that hedging at time 1 is not done by coefficient c016-math-221 in (16.23). Instead, at time 1 we consider the one period model between times 1 and 2, and choose the hedging coefficient of the one period model, as in (16.16).

16.1.3.3 The Martingale Measure in the Two Period Model

Let us find the martingale measure c016-math-222, which is implied by the quadratic hedging. We have to find such measure c016-math-223 that the option price is the discounted expectation with respect to the measure c016-math-224:

equation

We obtain from (16.25) and (16.27) that the density of c016-math-225 with respect to the underlying physical measure c016-math-226 is

equation

where

equation
equation

In fact,

equation

which can be written as

equation

where we use the notation c016-math-227.

When the increments are independent, then the martingale measure c016-math-228 is defined by the density

equation

where

equation

and

equation

compare this to c016-math-229 and c016-math-230 as defined in (16.21).

A Martingale Measure for S&P 500 in the Two Period Model

Let us estimate the equivalent martingale measure associated with quadratic hedging using S&P 500 daily data of Section 2.4.1. Let us consider a two-step model with two steps of 10 days. We take interest rate c016-math-231. Let

equation

be the increments. Our S&P 500 data provides a sample of identically distributed observations from the distribution of c016-math-232. We use non-overlapping increments.

Let us estimate the density

equation

of martingale measure c016-math-233 with respect to the physical measure c016-math-234.

First, we have to estimate c016-math-235 and c016-math-236 using nonparametric regression. Let us denote

equation

Let c016-math-237 and c016-math-238 be the kernel regression estimates.5 Then, we obtain the estimate of c016-math-244 as

equation

The estimate of c016-math-245 is

equation

Second, we have to estimate c016-math-246, c016-math-247, and c016-math-248. The estimates c016-math-249, c016-math-250, and c016-math-251 are the sample averages. Then, we obtain the estimate of c016-math-252 as

equation

The estimate of c016-math-253 is

equation

Now, we have obtained the estimate

The density of the martingale measure with respect to the Lebesgue measure can be estimated as

equation

where c016-math-255 is a two-dimensional kernel density estimate of the underlying physical measure of c016-math-256. The kernel density estimator is defined in (3.43).

When the returns are assumed to be independent, then we use the estimate

where

equation

and c016-math-258 and c016-math-259 are the sample versions of

equation

In the sample versions we replace the means and variances with the sample means and sample variances.

Figure 16.3 shows estimates of the density of the quadratic martingale measure with respect to the physical measure. In panel (a) we show estimate (16.28), which does not assume independence, and in panel (b) we show estimate (16.29), which assumes independence. In our setting regression estimation is difficult, and assuming independence leads to a more stable result. It is clear that the regression estimation could be improved by applying separate methods for the prediction of the first moment c016-math-260 and for the second moment c016-math-261.

Graphical representation of The quadratic martingale measure: Two period model.

Figure 16.3 The quadratic martingale measure: Two period model. Estimates of the density of the quadratic martingale measure with respect to the physical measure. (a) Increments are not assumed independent. (b) Increments are assumed to be independent.

16.1.4 The Multiperiod Model

We have derived the optimal hedging coefficient and the fair price in the mean-squared error sense for the two period model in Section 16.1.3. It is straightforward to generalize the results from the two period model to a general multiperiod model. The hedging coefficients and the fair price are derived using dynamic programming (backward induction).

16.1.4.1 Pricing in the Multiperiod Model

Let c016-math-262 be the value process of a self-financing portfolio:

16.30 equation

where

equation

and c016-math-264. We want to find c016-math-265 minimizing

where c016-math-267 is the discounted value of the derivative at the expiration.6

When c016-math-271 minimize (16.31), then we say that c016-math-272 is the fair price in the mean-squared error sense and c016-math-273 is the optimal hedging coefficient in the mean squared error sense. The coefficients c016-math-274 are needed to derive c016-math-275 in our backward induction, but they do not equal the optimal hedging coefficients at times c016-math-276. Instead, at time c016-math-277 we need to make a new calculation of coefficients, say c016-math-278, where c016-math-279 is the optimal hedging coefficient at time c016-math-280.

The following theorem is proved in Černý (2004b, Section 13.4), where c016-math-281, so that the number of risky assets is allowed to be larger than one. Černý (2004a) is an article with the same result, and Bertsimas et al. (2001) contains a similar kind of result. A similar kind of proof can be found in Schäl (1994), who considers the case of mean self-financing strategies. The case of independent increments was considered by Wolczyńska (1998) and Hammarlid (1998).

The Minimal Hedging Error

The proof implies that the minimal hedging error is given by c016-math-301, defined recursively in (16.36). Indeed, from (16.35) and (16.37), we obtain that

equation
The Hedging Coefficients

The proof implies that the sequence of quadratic hedging coefficients is given by

equation

where c016-math-302. The coefficient c016-math-303 is applied at time 0, and the coefficients c016-math-304 will not be applied, because at time c016-math-305, we need to construct a model of c016-math-306 periods.

Excess Gross Returns and Quadratic Hedging

The formulas for the price and the hedging coefficient are written using the increment

equation

We can write the formulas as well using the excess gross return

equation

The formulas (16.32) and (16.34) can be written as

where c016-math-308 and

equation

The fair price in the mean-squared error sense is c016-math-309 and the optimal hedging coefficient at time c016-math-310 is

Indeed, we can multiply by c016-math-312 and divide by c016-math-313 both the nominators and the denominators, and these terms can be moved inside the conditional expectations c016-math-314, because they are c016-math-315-measurable.

16.1.4.2 The Martingale Measure

Let us assume that the model is arbitrage-free. Theorem 13.1 says that there exists an equivalent martingale measure. Let us find the martingale measure c016-math-316 associated with quadratic hedging. According to Theorem 13.2 the martingale measure is such that the option price is the discounted expectation with respect to the measure:

equation

The density of c016-math-317 with respect to the underlying physical measure is

equation

where

equation

In fact, (16.32) in Theorem 16.2 implies that

equation

which can be written as

equation

where we use the notation c016-math-318. We can derive a similar expression for the density using the excess gross returns c016-math-319 instead of increments c016-math-320: we can apply (16.38).

16.1.4.3 Simplifying Assumptions

We can simplify the price formula (16.32) and the hedging formula (16.34) making restrictive assumptions on the increments c016-math-321. These assumptions are the martingale assumption, the assumption of a deterministic mean–variance ratio, the assumption of independence, and the assumption of independence and identical distribution.

Similar simplifications can be made to the formulas (16.38) and (16.39) when the assumptions are made on the process c016-math-322 of the gross returns.7

Quadratic Hedging Under the Martingale Assumption

Let us assume that c016-math-331 is a martingale with respect to the underlying physical measure c016-math-332. Then,

equation

c016-math-333-almost surely. Thus,

equation

Now, we have that c016-math-334, and

equation

This implies that the option price is the expected value:

equation

The first hedging coefficient is

equation

where c016-math-335, and c016-math-336. The expression for c016-math-337 is the same as in the one period model; see (16.19), (16.10), and (16.12). Using the rule of iterated expectations we can also write

equation

We can derive the result easily without using Theorem 16.2. Indeed,

equation

where c016-math-338. Under the martingale assumption,

equation

Also,

equation

Thus, we obtain a sum of similar one period optimizations as in (16.15).

Deterministic Mean–Variance Ratio

Let us denote

equation

Let us assume that the ratio

equation

is deterministic for c016-math-339. This assumption is made in Föllmer and Schied (2002, Proposition 10.40, p. 396) to derive an expression for the variance-optimal hedging strategy. Note that the mean–variance ratio was used in (16.7) to formulate a sufficient condition for the existence and uniqueness of the variance-optimal hedging strategy (the bounded mean–variance trade-off). Under the assumption of a deterministic mean–variance ratio it holds that c016-math-340 in (16.33) is deterministic. In fact, now c016-math-341, and

equation

That is,

equation

Values c016-math-342 are defined recursively for c016-math-343 by

where we start at c016-math-345. The fair price in the mean-squared error sense is c016-math-346 and the optimal hedging coefficient is

Independent Increments

Let us assume that the increments of discounted prices

equation

are independent. Assume that the sigma-algebras are generated by the price process: c016-math-348. Then, the independence of increments implies that the conditional expectations reduce to unconditional expectations, and

equation

are deterministic. Thus, the ratio c016-math-349 is deterministic, and we obtain the price and hedging formulas (16.40) and (16.41).

i.i.d. Increments

Let us assume that the increments of discounted prices c016-math-350 are independent and identically distributed. Let us denote

equation

We have that

equation

The price and hedging formulas are obtained from (16.40) and (16.41). Values c016-math-351 are defined recursively for c016-math-352 by

equation

where we start at c016-math-353. The fair price in the mean-squared error sense is c016-math-354 and the optimal hedging coefficient is

equation

The density of the martingale measure c016-math-355 with respect to the underlying physical measure is

equation

where

equation

16.2 Local Quadratic Hedging

Local quadratic hedging applies a much simpler recursive scheme for minimizing the quadratic hedging error than global quadratic hedging of Section 16.1. Local quadratic hedging solves the minimization only approximately. This numerical error could be compensated if a more accurate statistical estimation is possible.

Local quadratic hedging reduces the minimization of quadratic hedging error to a series of minimizations in one period models. Thus, in the one period model global and local quadratic hedging are identical. We introduce local quadratic hedging using the two period model, and after that cover the multiperiod model.

16.2.1 The Two Period Model

We introduce local quadratic hedging using the two period model. In the two period model the value process and the discounted contingent claim are defined as

equation

where c016-math-356, c016-math-357, c016-math-358, and c016-math-359.

In local quadratic hedging the minimization is done in two steps.

  1. 1. First, we find c016-math-360 and c016-math-361 minimizing
    equation

    This is the population version of a linear least-squares regression with the response variable c016-math-362 and the explanatory variable c016-math-363. The minimizers are

    equation
  2. 2. Second, we find c016-math-364 and c016-math-365 minimizing
    equation

    This is the population version of a linear least-squares regression with the response variable c016-math-366 and the explanatory variable c016-math-367. The minimizers are

    equation

The minimization problems are easier to solve than in the case of global quadratic hedging. However, we are not able to minimize

equation

but only to minimize it approximately.

We can write the price of the discounted contingent claim obtained by local quadratic hedging as

The first hedging coefficient c016-math-369 can be written as

The hedging coefficients c016-math-371 and c016-math-372 give the number of stocks in the hedging portfolio. The number of bonds c016-math-373 and c016-math-374 are obtained from the self-financing restrictions as in (16.3):

equation

16.2.1.1 A Comparison to the Global Quadratic Hedging

To highlight the difference between the local and the global quadratic hedging, let us recall the global quadratic hedging of Section 16.1. In the global quadratic hedging we want to find c016-math-376 minimizing

16.45 equation

The minimization problem can be solved in two steps. First, we find c016-math-378 minimizing

equation

Let the minimizer be c016-math-379. The minimizer depends on c016-math-380 and c016-math-381. Second, we find c016-math-382 and c016-math-383 minimizing

equation

The quadratic price is c016-math-384.

16.2.1.2 The Martingale Measure

Let us study the martingale measure c016-math-385 implied by the local hedging. The density of the martingale measure with respect to the physical measure c016-math-386 is

equation

where

equation

with

equation

for c016-math-387. The derivation of the martingale measure is given in (16.55) for the multiperiod model.

We can write the density in terms of the excess gross return. Namely,

equation

where

equation

and

equation

for c016-math-388. This is possible, because the denominator and nominator can be multiplied by the square of c016-math-389, which is c016-math-390-measurable, and can be placed inside c016-math-391.

A Martingale Measure for S&P 500: Two Steps

Let us estimate the equivalent martingale measure associated with quadratic hedging using S&P 500 daily data of Section 2.4.1. Let us consider a two-step model with two steps of 10 days. We choose interest rate c016-math-392. Let

equation

be the price increments, where c016-math-393 is the current time. When c016-math-394 runs through a long time period the observations are not stationary, but we can use our S&P 500 data to provide a sample of identically distributed observations of c016-math-395. We use non-overlapping increments. The observations are

equation

Let us estimate the density

equation

of martingale measure c016-math-396 with respect to the physical measure c016-math-397.

First, we have to estimate c016-math-398 and c016-math-399 using nonparametric regression. Let us denote

equation

Let c016-math-400 and c016-math-401 be the kernel regression estimates.8 Then, we obtain the estimates of c016-math-407 and c016-math-408 as

equation

The estimate of c016-math-409 is

equation

Second, we have to estimate c016-math-410 and c016-math-411. The estimates c016-math-412 and c016-math-413 are the sample averages. Then, we obtain the estimates of c016-math-414 and c016-math-415 as

equation

The estimate of c016-math-416 is

equation

Now, we have obtained the estimate

equation

The density of the martingale measure with respect to the Lebesgue measure can be estimated as

equation

where c016-math-417 is a two-dimensional kernel density estimate of the underlying physical measure of c016-math-418. The kernel density estimator is defined in (3.43).

Figure 16.4 shows estimates of the density of the local quadratic martingale measure with respect to the physical measure. Panel (a) shows a contour plot and panel (b) shows a perspective plot.

Graphical representation of local quadratic martingale measure: Two period model.

Figure 16.4 A local quadratic martingale measure: Two period model. (a) A contour plot; (b) a perspective plot. Estimates of the density of the local quadratic martingale measure with respect to the physical measure.

16.2.2 The Multiperiod Model

Let

equation

be the discounted value of the derivative at the expiration. We define recursively values c016-math-419 and c016-math-420, c016-math-421, starting with the value c016-math-422. Let c016-math-423 and c016-math-424 be the minimizers of

equation

for c016-math-425, over c016-math-426, where

equation

This is a conditional population least-squares linear regression problem with the response variable c016-math-427 and the explanatory variable c016-math-428. The problem is similar to the minimization problem in the one period model of Section 16.1.2, but now we are conditioning on c016-math-429. The solutions are

equation

and

equation

Value c016-math-430 is the price suggested by local quadratic hedging and c016-math-431 is the hedging coefficient at time 0, which is suggested by local quadratic hedging.

We can write c016-math-432 using the undiscounted prices c016-math-433 and c016-math-434 as

16.46 equation

where

16.47 equation

The price can be written as:9

The hedging coefficients can be written as10

equation

When c016-math-441 are independent, then

where c016-math-443, and the price is

Also, similarly as in (14.34), when c016-math-445 are independent,

equation

where c016-math-446.

16.2.2.1 A Comparison to Black–Scholes

We compare the quadratic prices and hedging coefficients to the Black–Scholes prices and hedging coefficients. We assume the independence of increments and use formulas (16.49) and (16.50). We apply S&P 500 daily data of Section 2.4.1. The Black–Scholes prices and deltas are computed using the annualized standard deviation as the volatility.

Figure 16.5 compares quadratically optimal prices to Black–Scholes prices. Panel (a) shows the quadratically optimal prices (black) and the Black–Scholes prices (red) as a function of moneyness c016-math-447. Time to expiration is 20 trading days. Panel (b) shows the ratios of the quadratically optimal prices to the Black–Scholes prices as a function of moneyness. Time to expiration is 20 days (black), 40 days (red), 60 days (blue), and 80 days (green). We see from panel (b) that when the moneyness is less than one, then the quadratic prices are less than the Black–Scholes prices. When the moneyness is about 0.95, then increasing the time to expiration makes the ratio of the quadratic prices to the Black–Scholes prices increase.

Graphical representation of Call prices.

Figure 16.5 Call prices. (a) The quadratic prices (black) and the Black–Scholes prices (red) as a function of moneyness c016-math-448. (b) The ratios of the quadratic prices to the Black–Scholes prices. Time to expiration is 20 days (black), 40 days (red), 60 days (blue), and 80 days (green).

Figure 16.6 compares quadratic hedging coefficients to Black–Scholes hedging coefficients. Panel (a) shows the quadratic hedging coefficients (black) and the Black–Scholes hedging coefficients (red) as a function of moneyness c016-math-449. Time to expiration is 20 trading days. Panel (b) shows the ratios of the quadratic hedging coefficients to the Black–Scholes hedging coefficients as a function of moneyness. Time to expiration is 20 days (black), 40 days (red), 60 days (blue), and 80 days (green). We see from panel (b) that when the moneyness is about one, then increasing the time to expiration makes the ratio of the quadratic hedging coefficient to the Black–Scholes hedging coefficient increase. When the moneyness is less than 0.95 and the time to expiration is 20 days, then the quadratic hedging coefficient is much larger than the Black–Scholes hedging coefficient.

Graphical representation of Hedging coefficients.

Figure 16.6 Hedging coefficients. (a) The quadratic hedging coefficients (black) and the Black–Scholes hedging coefficients (red) as a function of moneyness c016-math-450. (b) The ratios of the quadratic hedging coefficients to the Black–Scholes hedging coefficients. Time to expiration is 20 days (black), 40 days (red), 60 days (blue), and 80 days (green).

16.2.2.2 Square Integrability

The local quadratic trading strategy needs to be square integrable, in the sense of assumption (16.6). The square integrability is studied in Föllmer and Schied (2002, Proposition 10.10, p. 377). In fact, in order to guarantee the satisfaction of (16.6), it is enough to assume

c016-math-452-almost surely, for c016-math-453, for a constant c016-math-454. Condition (16.51) of the bounded mean–variance trade-off appeared already in (16.7), where it was stated to guarantee the existence of a global quadratic trading strategy. Denote c016-math-455. Assumption (16.51) implies that c016-math-456 Thus we have for the local quadratic hedging coefficients c016-math-457 that

equation

where we used for the first equality the law of the iterated expectations, and for the third inequality the Cauchy–Schwarz inequality. Thus, the square integrability of c016-math-458 implies the square integrability of c016-math-459, which implies the square integrability of c016-math-460. The backward induction shows that the square integrability of c016-math-461 implies the square integrability of c016-math-462 for c016-math-463, under assumption (16.51).

16.2.2.3 The Equivalent Martingale Measure

Let us find the equivalent martingale measure c016-math-464 implied by the local quadratic hedging. The density of the martingale measure with respect to the physical measure c016-math-465 is

where

equation

with

In order that density c016-math-468 is positive we have to assume that

c016-math-470-almost surely on c016-math-471. Otherwise, c016-math-472 would be a signed measure and not a probability measure.

A Derivation of the Equivalent Martingale Measure

Let us show that

equation

This follows because we can write

Now, we have

equation

which can be written as

equation

where we use the notation c016-math-475. Note that (16.56) implies that

Characterizations of the Equivalent Martingale Measure

Measure c016-math-477 in (16.52) can be characterized as a minimal martingale measure. Föllmer and Schied (2002, Definition 10.21, p. 382) define a minimal martingale measure to be such measure c016-math-478 which is equivalent to c016-math-479, c016-math-480, and such that every square integrable c016-math-481-martingale c016-math-482 which is strongly orthogonal to c016-math-483 is also a c016-math-484-martingale. The strong orthogonality of c016-math-485 and c016-math-486 means that

equation

c016-math-487-almost surely, for c016-math-488.

Föllmer and Schied (2002, Theorem 10.22, p. 383) states that if c016-math-489 is a minimal martingale measure, then (16.57) holds. Föllmer and Schied (2002, Corollary 10.28, p. 388) states that there exists at most one minimal martingale measure.

Föllmer and Schied (2002, Theorem 10.30, p. 390) proves the existence and the uniqueness of a minimal martingale measure, and gives formula (16.58) for the density of the minimal martingale measure. Let us assume the condition (16.54) of positivity and the condition (16.51) of the bounded mean–variance trade-off. Then there exists a unique minimal martingale measure c016-math-490 with density

where c016-math-492 and

equation

with c016-math-493 and

equation

where c016-math-494 is defined in (16.53), and c016-math-495 is the martingale part of the Doob decomposition of c016-math-496. The Doob decomposition of c016-math-497 is

equation

where c016-math-498 is a martingale and c016-math-499 is predictable. The Doob decomposition is defined as

equation

where c016-math-500, c016-math-501, and c016-math-502; see Föllmer and Schied (2002, Proposition 6.1, p. 277).

Now we can show that the measure in (16.52) is the same as the minimal martingale measure in (16.58). Indeed,

equation

and

equation

16.2.3 Local Quadratic Hedging without Self-Financing

It is of interest to note that when we define a local quadratic hedging without self-financing, then the price will be the same, the hedging coefficients of the stocks will be the same, and only the hedging coefficients of the bonds will be different. A local quadratic hedging without self-financing can be defined in a similar way as the local quadratic hedging with self-financing, but we replace the value process with the wealth process.

16.2.3.1 Backward Induction

Let us consider the two period model with c016-math-503. Let us describe local quadratic hedging when the wealth process is used. The wealth at times 0, 1, and 2 is equal to

equation

The self-financing condition would state that c016-math-504 and c016-math-505 should be chosen so that

equation

In local quadratic hedging without self-financing we first find c016-math-506 and c016-math-507 minimizing

equation

This is the population version of a linear least-squares regression with the response variable c016-math-508 and the explanatory variable c016-math-509. The minimizers are

equation

Let us denote

equation

Term c016-math-510 is obtained by “discounting” term c016-math-511. Second, we find c016-math-512 and c016-math-513 minimizing

equation

This is the population version of a linear least-squares regression with the response variable c016-math-514 and the explanatory variable c016-math-515. The minimizers are

equation

The optimal price in the local quadratic sense is

equation

Term c016-math-516 is obtained by “discounting” term c016-math-517. The price can be written as

equation

The price is equal to the price which is obtained with the self-financing condition, as can be seen from (16.42). The first hedging coefficient can be written as

equation

The hedging coefficient is equal to the hedging coefficient in (16.43), which is obtained with the self-financing restriction.

16.2.3.2 A Comparison with the Case of Self-Financing

We have seen that the price c016-math-518 and the hedging coefficients c016-math-519 and c016-math-520 are the same whether the self-financing restriction is imposed or not. What about the coefficients c016-math-521 and c016-math-522? The quantities of the bonds are given by

equation

The quantities can be compared to the quantities when the self-financing condition holds, given in (16.44) as

equation

We see that c016-math-523 are equal, but c016-math-524 are different.11

16.2.3.3 Mean Self-Financing

We have obtained a hedging strategy that is not self-financing, but it is mean self-financing, as defined in (16.5). Indeed,

equation

because c016-math-529 and c016-math-530 since c016-math-531 is c016-math-532-measurable and c016-math-533 is c016-math-534-measurable.

16.3 Implementations of Local Quadratic Hedging

We have derived formulas for the quadratic price and the quadratic hedging coefficient. The formulas are not in a closed form but their application requires numerical methods. In addition, the formulas depend on the knowledge of the unknown data generating mechanism, and we need to use statistical methods to estimate the data generating mechanism.

We implement only the local quadratic hedging, both for the case when the increments are assumed to be independent, and for the case when the increments are assumed to be dependent.

Section 16.3.1 describes the basic setting of historical simulation. Section 16.3.2 describes numerical and statistical methods for the case of independent increments. Section 16.3.3 considers the case of dependent increments. Section 16.3.4 compares the implementations of quadratic pricing and hedging to some benchmarks.

16.3.1 Historical Simulation

To implement quadratic hedging we use historical simulation. Analogously, Monte Carlo simulation could be applied. In Monte Carlo simulation a statistical model is imposed, and sequences of observations are generated from the model. In historical simulation only the previous observations are used.

A similar type of implementation has been described in Potters et al. (2001), where price functions c016-math-535 and hedging functions c016-math-536 are estimated using an expansion with basis functions, whereas we use kernel estimation. Also, we implement a method where the price function and the hedging function depend on volatility, so that they have the form c016-math-537 and c016-math-538.

16.3.1.1 Generating Sequences of Observations

We denote the time series of observed historical daily prices by c016-math-539. The price c016-math-540 is the current price. We construct c016-math-541 sequences of prices:

where

equation

Each sequence consists of c016-math-543 values, and the initial price in each sequence is c016-math-544.

We may choose to use less than c016-math-545 sequences, to make computation faster. Note that c016-math-546 sequences are overlapping, so that the use of the all possible c016-math-547 sequences may not increase statistical accuracy much, as compared for using a lesser number of sequences. We may construct c016-math-548 sequences of prices, and to get non-overlapping sequences we may choose c016-math-549, and choose index c016-math-550 to take the values c016-math-551, for c016-math-552.

16.3.1.2 The State Variable

With sequence c016-math-553 of prices there is an associated sequence c016-math-554 of state variables. Each c016-math-555 can be a vector. We have constructed sequences c016-math-556, which all start at the current stock price c016-math-557. The values of the state variables that correspond to sequence c016-math-558 are

equation

To utilize the information in the state variables, we use only those sequences c016-math-559 that are such that at time c016-math-560 the value of the vector c016-math-561 of state variables is close to the current value c016-math-562 of the state variables. Let c016-math-563 be the collection of those times:

equation

where c016-math-564 is the radius of the window, and c016-math-565 is the Euclidean distance.

For example, we can choose the state variable to be the logarithm of the current prediction of volatility:

equation

where c016-math-566 is estimated using the observed prices c016-math-567. For instance, we can apply the GARCH(c016-math-568) volatility estimate.12Then c016-math-584 is defined as

equation

where c016-math-585 is the radius of the window. This is similar to the nonparametric GARCH-pricing in Section 15.3.

16.3.1.3 Heuristic Discussion

We want to solve a series of linear regression problems

for c016-math-587, where c016-math-588. These regression problems are conditional on c016-math-589 and c016-math-590. The solutions are functions c016-math-591 and c016-math-592, where c016-math-593 is the discounted value of the stock and c016-math-594 is the value of the state variable. The sample version of the regression problem is

equation

where c016-math-595, and c016-math-596.

In analogy, consider first the standard linear regression model

equation

Assume we observe c016-math-597, c016-math-598, from this model. Then,

equation

and we can estimate the constants c016-math-599 and c016-math-600. Our setting resembles the model

equation

where c016-math-601 is an additional random variable, and c016-math-602 and c016-math-603 are functions. Assume that we observe c016-math-604, c016-math-605, from this model. Then

equation

In order to estimate values c016-math-606 and c016-math-607 for a fixed c016-math-608 we cannot use the standard linear regression, because there are no observations from model c016-math-609. Instead, we can estimate the functions c016-math-610 and c016-math-611 by localizing into the neighborhood of c016-math-612. Let c016-math-613. Now, we can use linear regression for the observations

equation

Note that we need to estimate functions c016-math-614 and c016-math-615 only at the points c016-math-616 and c016-math-617, c016-math-618.

We need to estimate functions c016-math-619 and c016-math-620 of two arguments (where c016-math-621 may be a vector). This can be done in two ways.

  1. 1. We can localize with respect to both c016-math-622 and c016-math-623.
  2. 2. We can estimate function

    Now, it is possible to avoid localization with respect to c016-math-625, make the localization only with respect to c016-math-626, and have available more observations to make the estimation. This is possible for certain c016-math-627.

Estimation of (16.61) is done by changing model (16.60) to model

where

equation

An estimate c016-math-629 leads to estimate

equation

Now the localization with respect to c016-math-630 is not necessary when c016-math-631, since c016-math-632. In this case we can ignore the current level of the trajectory of the stock price.

16.3.2 Local Quadratic Hedging Under Independence

We apply the local quadratic hedging and assume independence of the increments. The hedging coefficients are given by the formula (16.49) as

equation

where c016-math-633. The price is given by the formula (16.50) as

equation

Here c016-math-634 is the discounted payoff of the derivative, c016-math-635, and c016-math-636 are the discounted prices of the risky asset.

Let us assume for notational simplicity that the risk-free rate is zero, so that

equation

The price c016-math-637 involves only unconditional expectations, whereas hedging coefficients c016-math-638 involve conditional expectations. In our implementation of the case of independent increments we make two simplifications, as compared to the previous heuristic discussion. First, the conditioning with respect to the state variable is done only at the time c016-math-639 of writing the option. Second, we do not need to move to the model (16.62), but we can handle the conditioning on the stock price c016-math-640 by renormalizing the tails of sequences c016-math-641 so that they start with value c016-math-642. This is possible because in the case of independent increments the intermediate values c016-math-643 for c016-math-644 do not appear.

16.3.2.1 Unconditional Expectations

To estimate the unconditional expectations c016-math-645 we apply sequences c016-math-646 in (16.59), where c016-math-647. These sequences give us the differences and the terminal values

equation

where c016-math-648. We estimate the unconditional expectations by

equation

16.3.2.2 Conditional Expectations

To estimate the hedging coefficients c016-math-649, when the stock price is c016-math-650, we renormalize the tails of the sequences. We define sequences such that the initial price is c016-math-651, and the number of observations in each sequence is c016-math-652, where c016-math-653, that is, the length of the sequences is c016-math-654. We define

where

equation

c016-math-656 Now the initial price in each sequence is c016-math-657.

We estimate the conditional expectations c016-math-658 by applying sequences c016-math-659 in (16.63), where c016-math-660, c016-math-661, and c016-math-662. Each such sequence gives the differences and the terminal values

equation

The conditional expectations c016-math-663 are estimated by

equation

These estimates lead to the estimates of covariances and variances,13 and we obtain estimates of c016-math-666, which are used to produce an estimate of c016-math-667.

16.3.2.3 Comparison to Black–Scholes

We compare prices and hedging coefficients of local quadratic hedging (with independence assumption) to the Black–Scholes prices and hedging coefficients.

Comparison to Black–Scholes Prices

Figure 16.7 shows the ratios of the locally quadratic prices (under independence) to the Black–Scholes prices as a function of the annualized volatility. In panel (a) moneyness is c016-math-668 and in panel (b) c016-math-669. The smoothing parameter is c016-math-670 (black), c016-math-671 (red), and c016-math-672 (blue). The time to expiration is 20 trading days.

Graphical representation of Call price ratios as a function of volatility.

Figure 16.7 Call price ratios as a function of volatility. The ratios of the locally quadratic prices under independence to the Black–Scholes prices as a function of the annualized volatility. (a) Moneyness is c016-math-673; (b) c016-math-674. The smoothing parameter is c016-math-675 (black), c016-math-676 (red), and c016-math-677 (blue).

Figure 16.8 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the moneyness c016-math-678. In panel (a) the time to expiration is c016-math-679 trading days, and in panel (b) c016-math-680. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). The smoothing parameter is c016-math-681.

Graphical representation of Call price ratios as a function of moneyness.

Figure 16.8 Call price ratios as a function of moneyness. The ratios of the locally quadratic prices under independence to the Black–Scholes prices as a function of moneyness c016-math-682. (a) c016-math-683; (b) c016-math-684. The annualized volatility is c016-math-685 (black), c016-math-686 (red), and c016-math-687 (blue).

Figure 16.9 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the smoothing parameter c016-math-688. In panel (a) moneyness is c016-math-689 and in panel (b) c016-math-690. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). Time to expiration is c016-math-691 trading days.

Graphical representation of Call price ratios as a function of the smoothing parameter.

Figure 16.9 Call price ratios as a function of the smoothing parameter. The ratios of the locally quadratic prices under independence to the Black–Scholes prices as a function of smoothing parameter c016-math-692. (a) Moneyness is c016-math-693; (b) c016-math-694. The annualized volatility is c016-math-695 (black), c016-math-696 (red), and c016-math-697 (blue).

Comparison to Black–Scholes Deltas

Figure 16.10 shows the ratios of the locally quadratic hedging coefficients (under independence) to the Black–Scholes deltas as a function of the annualized volatility. In panel (a) moneyness is c016-math-698 and in panel (b) c016-math-699. The smoothing parameter is c016-math-700 (black), c016-math-701 (red), and c016-math-702 (blue). The time to expiration is 20 trading days.

Graphical representation of Call hedging coefficient ratios as a function of volatility.

Figure 16.10 Call hedging coefficient ratios as a function of volatility. The ratios of the locally quadratic hedging coefficients under independence to the Black–Scholes deltas as a function of the annualized volatility. (a) Moneyness is c016-math-703; (b) c016-math-704. The smoothing parameter is c016-math-705 (black), c016-math-706 (red), and c016-math-707 (blue).

Figure 16.11 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the moneyness c016-math-708. In panel (a) time to expiration is c016-math-709 trading days, and in panel (b) c016-math-710. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). The smoothing parameter is c016-math-711.

Graphical representation of Call hedging coefficient ratios as a function of moneyness.

Figure 16.11 Call hedging coefficient ratios as a function of moneyness. The ratios of the locally quadratic prices under independence to the Black–Scholes prices as a function of moneyness c016-math-712. (a) c016-math-713; (b) c016-math-714. The annualized volatility is c016-math-715 (black), c016-math-716 (red), and c016-math-717 (blue).

Figure 16.12 shows the ratios of the locally quadratic hedging coefficients to the Black–Scholes deltas as a function of the smoothing parameter c016-math-718. In panel (a) moneyness is c016-math-719 and in panel (b) c016-math-720. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). Time to expiration is c016-math-721 trading days.

Graphical representation of Call hedging coefficient ratios as a function of smoothing parameter.

Figure 16.12 Call hedging coefficient ratios as a function of smoothing parameter. The ratios of the locally quadratic hedging coefficients under independence to the Black–Scholes deltas as a function of smoothing parameter c016-math-722. (a) Moneyness is c016-math-723; (b) c016-math-724. The annualized volatility is c016-math-725 (black), c016-math-726 (red), and c016-math-727 (blue).

16.3.3 Local Quadratic Hedging under Dependence

We apply local quadratic hedging without assuming independence of increments. We need to estimate the sequences

16.64 equation

where c016-math-729. The recursion starts with the known value c016-math-730.

Let us assume for notational simplicity that the risk-free rate is zero, so that

equation

Let us denote by c016-math-731 the observed values of the stock c016-math-732, and let us denote by c016-math-733 the observed values of the state variable c016-math-734. The values in sequence c016-math-735, defined in (16.59), are denoted by

equation

where c016-math-736. The corresponding sequence of the values of the state variables is

equation

In the case of local quadratic hedging under independence the formulas did not involve the intermediate values c016-math-737 for c016-math-738. This simplified the computations, and we needed only to renormalize the tails of the price trajectories. Now we use a technique where we move from the increments c016-math-739 to the net returns c016-math-740 of stock prices, and from the values c016-math-741 to the values c016-math-742, where

equation

We can write

equation

We need to estimate the conditional expectations

equation

The conditional expectations are interpreted as

equation

16.3.3.1 The Steps of Backward Induction

Consider step c016-math-743. Assume that we have produced estimates

equation

where

equation

Step c016-math-744 is the first step of the backward induction. When c016-math-745, then in the case of a call option

equation

where c016-math-746, c016-math-747, are the terminal values of sequences c016-math-748.

We estimate the conditional expectations first with local averaging, and then generalize to kernel estimation.

Local Averaging

Let

equation

Set c016-math-749 contains those indexes c016-math-750 for which the c016-math-751th element c016-math-752 is close to c016-math-753, where c016-math-754. Note that c016-math-755. Let

equation

We estimate for each c016-math-756 the conditional expectations by14

equation
equation

These estimates lead to the estimates of covariances and variances, and we obtain an estimate of c016-math-759. An estimate for c016-math-760 is obtained by

equation

for c016-math-762 and c016-math-763, where c016-math-764. Note that at step c016-math-765 set c016-math-766 is a singleton:

equation

The price is obtained as

equation

and the first hedging coefficient is

equation
Kernel Estimation

The estimation of the conditional expectations can be done by using kernel estimation. We estimate for each c016-math-767 the conditional expectations using the estimators

equation
equation

where the weights are defined as

equation

c016-math-768 is the scaled kernel c016-math-769, c016-math-770 is a kernel function, where c016-math-771 is the dimension of vector c016-math-772. The previous method of local averaging is obtained as a special case when the kernel function is chosen as

equation

These estimates lead to the estimates of covariances and variances, and we obtain an estimate of c016-math-773. An estimate of c016-math-774 is obtained using the formula (16.65).

16.3.3.2 Comparison to Black–Scholes

We compare both prices and hedging coefficients of local quadratic hedging to the Black–Scholes prices and hedging coefficients.

Comparison to Black–Scholes Prices

Figure 16.13 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the annualized volatility. In panel (a) moneyness is c016-math-775 and in panel (b) c016-math-776. The smoothing parameter is c016-math-777 (black), c016-math-778 (red), and c016-math-779 (blue). Time to expiration is 20 trading days.

Graphical representation of Call price ratios as a function of volatility under dependence.

Figure 16.13 Call price ratios as a function of volatility under dependence. The ratios of the locally quadratic prices to the Black–Scholes prices as a function of the annualized volatility. (a) Moneyness is c016-math-780; (b) c016-math-781. The smoothing parameter is c016-math-782 (black), c016-math-783 (red), and c016-math-784 (blue).

Figure 16.14 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the moneyness c016-math-785. In panel (a) time to expiration is c016-math-786 trading days, and in panel (b) c016-math-787. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). The smoothing parameter is c016-math-788.

Graphical representation of Call price ratios as a function of moneyness: Dependence.

Figure 16.14 Call price ratios as a function of moneyness: Dependence. The ratios of the locally quadratic prices under dependence to the Black–Scholes prices as a function of moneyness c016-math-789. (a) c016-math-790; (b) c016-math-791. The annualized volatility is c016-math-792 (black), c016-math-793 (red), and c016-math-794 (blue).

Figure 16.15 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the smoothing parameter c016-math-795. In panel (a) moneyness is c016-math-796 and in panel (b) c016-math-797. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). Time to expiration is c016-math-798 trading days.

Image described by caption and surrounding text.

Figure 16.15 Call price ratios as a function of the smoothing parameter: Dependence. The ratios of the locally quadratic prices under dependence to the Black–Scholes prices as a function of smoothing parameter c016-math-799. (a) Moneyness is c016-math-800; (b) c016-math-801. The annualized volatility is c016-math-802 (black), c016-math-803 (red), and c016-math-804 (blue).

Comparison to Black–Scholes Deltas

Figure 16.16 shows the ratios of the locally quadratic hedging coefficients (under dependence) to the Black–Scholes deltas as a function of the annualized volatility. In panel (a) moneyness is c016-math-805 and in panel (b) c016-math-806. The smoothing parameter is c016-math-807 (black), c016-math-808 (red), and c016-math-809 (blue). Time to expiration is 20 trading days.

Image described by caption and surrounding text.

Figure 16.16 Call hedging coefficient ratios as a function of volatility: Dependence. The ratios of the locally quadratic hedging coefficients under dependence to the Black–Scholes deltas as a function of the annualized volatility. (a) Moneyness is c016-math-810; (b) c016-math-811. The smoothing parameter is c016-math-812 (black), c016-math-813 (red), and c016-math-814 (blue).

Figure 16.17 shows the ratios of the locally quadratic prices to the Black–Scholes prices as a function of the moneyness c016-math-815. In panel (a) time to expiration is c016-math-816 trading days, and in panel (b) c016-math-817. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). The smoothing parameter is c016-math-818.

Image described by caption and surrounding text.

Figure 16.17 Call hedging coefficient ratios as a function of moneyness: Dependence. The ratios of the locally quadratic prices under dependence to the Black–Scholes prices as a function of moneyness c016-math-819. (a) c016-math-820; (b) c016-math-821. The annualized volatility is c016-math-822 (black), c016-math-823 (red), and c016-math-824 (blue).

Figure 16.18 shows the ratios of the locally quadratic hedging coefficients to the Black–Scholes deltas as a function of the smoothing parameter c016-math-825. In panel (a) moneyness is c016-math-826 and in panel (b) c016-math-827. The annualized volatility is 0.1 (black), 0.2 (red), and 0.3 (blue). Time to expiration is c016-math-828 trading days.

Image described by caption and surrounding text.

Figure 16.18 Call hedging coefficient ratios as a function of smoothing parameter: Dependence. The ratios of the locally quadratic hedging coefficients under dependence to the Black–Scholes deltas as a function of smoothing parameter c016-math-829. (a) Moneyness is c016-math-830; (b) c016-math-831. The annualized volatility is c016-math-832 (black), c016-math-833 (red), and c016-math-834 (blue).

16.3.4 Evaluation of Quadratic Hedging

Figure 16.19 shows (a) tail plots and (b) kernel density estimates of hedging errors for call options.15 The blue curves show the case of Black–Scholes hedging. The local quadratic hedging is done assuming independence and the smoothing parameter is c016-math-845 (red), c016-math-846 (dark green), c016-math-847 (purple), and c016-math-848 (orange). The volatility is in all cases the GARCH(c016-math-849) volatility. The moneyness of call options is c016-math-850. Time to maturity is 20 days and hedging is done every day. Tail plots are defined in Section 3.2.1 and the kernel density estimator is defined in Section 3.2.2. We apply the standard normal kernel function and the smoothing parameter of the density estimator is chosen by the normal reference rule. We see from panel (b) that smoothing parameters c016-math-851 lead to similar results, but smoothing parameter c016-math-852 leads to a more dispersed distribution. Black–Scholes hedging leads to a more concentrated distribution than the quadratic hedging with independence assumption, but the quadratic hedging leads to a distribution which is skewed to the right in the central area of the distribution, which means that there are more gains than losses for the hedger of the option.

Graphical representation of Distribution of hedging errors: Local quadratic with independence.

Figure 16.19 Distribution of hedging errors: Local quadratic with independence. Shown are (a) tail plots and (b) kernel density estimates of hedging errors. Black–Scholes hedging (blue), local quadratic hedging with independence using the smoothing parameter c016-math-853 (red), c016-math-854 (dark green), c016-math-855 (purple), and c016-math-856 (orange).

Image described by caption and surrounding text.

Figure 16.20 Distribution of hedging errors when hedging is done once: Local quadratic under independence and dependence. Shown are (a) tail plots and (b) kernel density estimates of hedging errors. Black–Scholes hedging (blue), local quadratic hedging with independence (green), and local quadratic hedging with dependence (red).

Figure 16.20 considers the case of hedging only once. Panel (a) shows tail plots and panel (b) shows kernel density estimates of hedging errors. The moneyness of call options is c016-math-857. Time to maturity is 20 days. The blue curves show the case of Black–Scholes hedging. The red curves show the case of local quadratic hedging assuming dependence. The green curves show the case of local quadratic hedging assuming independence. The smoothing parameter is in both cases c016-math-858. The volatility is in all cases the GARCH(c016-math-859) volatility.

Image described by caption and surrounding text.

Figure 16.21 Distribution of hedging errors with daily hedging: Local quadratic under independence and dependence. Shown are (a) tail plots and (b) kernel density estimates of hedging errors. Black–Scholes hedging (blue), local quadratic hedging with independence (green), and local quadratic hedging with dependence (red).

Figure 16.21 shows (a) tail plots and (b) kernel density estimates of hedging errors. The moneyness of call options is c016-math-860. Time to maturity is 20 days and hedging is done every day. The blue curves show the case of Black–Scholes hedging. The red curves show the case of local quadratic hedging assuming dependence. The green curves show the case of local quadratic hedging assuming independence. The smoothing parameter is in both cases c016-math-861. The volatility is in all cases the GARCH(c016-math-862) volatility.

equation

From the second equation we get

From the first equation we get c016-math-078. Inserting this to (16.9) gives

equation

It holds that

equation
equation

Inserting this to (16.19) gives

equation
equation

where c016-math-239, c016-math-240, are the observations from the distribution of c016-math-241,

equation

c016-math-242 is a kernel function and c016-math-243 is a smoothing parameter.

equation

where c016-math-269 is the wealth, and c016-math-270 is the value of the option at the expiration.

equation
equation

where c016-math-402, c016-math-403, are the observations of c016-math-404, c016-math-405 is a kernel function and c016-math-406 is a smoothing parameter.

equation

On the other hand, c016-math-438

equation

and the generalization of (16.48) to c016-math-440:

equation
equation

Let us define the objective function

equation

which is to be minimized with respect to c016-math-525, c016-math-526, c016-math-527, and c016-math-528. However, this optimization problem is not easier to solve than the global quadratic optimization.

equation

where c016-math-570, c016-math-571. Prediction c016-math-572 is made at time c016-math-573, and it predicts the volatility at time c016-math-574. In order to obtain initial estimates of parameters c016-math-575, c016-math-576, c016-math-577, and an initial value c016-math-578 for the volatility, we assume that there are available observations c016-math-579. It is reasonable to update the estimates c016-math-580, c016-math-581, and c016-math-582 sequentially, using data c016-math-583.

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

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