Chapter 14
Pricing by Arbitrage

Pricing by arbitrage means that an asset is priced with the unique arbitrage-free price. Pricing by arbitrage can be applied in two different settings: (1) We can price by arbitrage linear securities, like forwards and futures, in any markets. (2) We can price by arbitrage nonlinear securities, like options, in complete markets. We discuss both of these cases in this chapter.

If two assets have the same terminal value with probability one, then the assets should have the same price. Otherwise, we could obtain a risk-free profit by selling the more expensive asset and by buying the cheaper asset. Linear assets, like futures, can be defined as a linear function of the underlying assets. Thus, they can be replicated, and their price is the initial value of the replicating portfolio. Nonlinear assets, like options, can be replicated only under restrictive assumptions on the markets (under the assumption of complete markets). However, the restrictive assumptions are often not too far away from the real properties of the markets.

The concepts of an arbitrage-free market and a complete market were studied in Chapter 13. The results of Chapter 13 imply that if the market is arbitrage-free and complete, then there is only one arbitrage-free price. In fact, Theorem 13.2 states that the arbitrage-free prices are obtained as expected values with respect to the equivalent martingale measures, and Theorem 13.3 (the second fundamental theorem of asset pricing) states that if the market is arbitrage-free and complete, then there is only one equivalent martingale measure.

Section 14.1 discusses pricing of futures, the put–call parity, and the American call options.

Section 14.2 studies binary models. In these models the price of a stock can at any time move only one step up, or one step down. The binary models are complete models, thus a derivative has a unique arbitrage-free price. These prices can be easily computed. The prices are called the Cox–Ross–Rubinstein prices, and they were introduced in Cox et al. (1979).

Section 14.2.3 studies asymptotics of multiperiod binary models, as the number of periods increases, and the length of the periods decreases. A multiperiod binary model converges to a log-normal model. In the log-normal model the stock price follows a geometric Brownian motion. The geometric Brownian motion is a market model that does not exactly describe the actual markets but it can be rather close to the actual markets. The Black–Scholes prices are obtained as the limits of the Cox–Ross–Rubinstein prices. The derivation of the Black–Scholes prices from the Cox–Ross–Rubinstein prices is both elegant and it leads to efficient numerical recipes, although there are many other derivations of the Black–Scholes price (see Section 14.3.3).

Section 14.3 studies the properties of the Black–Scholes prices. Section 14.4 studies Black–Scholes hedging. Section 14.5 studies combining the Black–Scholes hedging with time changing volatility estimates. This method of option pricing provides a benchmark against which we can evaluate competing pricing and hedging methods.

14.1 Futures and the Put–Call Parity

Futures can be priced by arbitrage regardless whether the market is complete or not. The payoff of a futures contract is a linear function of the underlying, and thus a futures contract can be replicated even in an incomplete market. The put–call parity provides a related example of pricing by arbitrage, which works whether the market is complete or not: the payoff of a put subtracted from the payoff of a call is a linear function of the underlying.

14.1.1 Futures

We consider pricing of stock futures, currency forwards, forward zero-coupon bonds, and forward rate agreements.

14.1.1.1 Stock Futures

A futures contract on a stock is made at time c014-math-001. The contract specifies that the buyer of the contract has to buy the stock at a later time c014-math-002 with price c014-math-003. We assume that the stock does not pay dividends during the time period from c014-math-004 to c014-math-005. Let us denote with c014-math-006 the price of the stock at time c014-math-007. The value of the futures contract at the expiration time c014-math-008 is

equation

because the buyer of the futures contract gives away c014-math-009 and receives c014-math-010.

The price of a zero-coupon bond is taken as c014-math-011, where c014-math-012, c014-math-013 is the annualized risk-free rate, and c014-math-014 is the time between c014-math-015 and c014-math-016 in fractions of a year. This is the method of continuous compounding.1

What is the fair value c014-math-023 of the futures contract for c014-math-024? We may replicate the futures contract buy buying the stock and borrowing the amount c014-math-025. At time c014-math-026 the value of this portfolio is

equation

with probability one. At time c014-math-027 the value of this portfolio is c014-math-028 with probability one. Thus,

for c014-math-030, by the law of one price, to exclude arbitrage.

When an investor enters a futures contract in a futures exchange, this does not imply any cash flows, although the exchange requires from the investor a liquid collateral in order to secure a possible future payment. The fair forward price is called such value of c014-math-031 that makes the value c014-math-032 of the futures contract zero. To get c014-math-033 we need that

The future prices that are quoted in a futures exchange are the values c014-math-035 (which are determined by the supply and demand). Numbers c014-math-036 are called futures prices or forward prices.

A buyer of a stock future or a stock index future saves the carrying costs but looses the possible stock dividends. When the annualized dividend rate is known to be c014-math-037, then the fair future price is

equation

14.1.1.2 Currency Forwards

A currency forward is made at time c014-math-038. We denote with c014-math-039 the exchange rate USD/euro at time c014-math-040. Let c014-math-041 be an exchange rate. The contract stipulates that the buyer will buy c014-math-042 US dollars with euros at time c014-math-043, using the USD/euro exchange rate c014-math-044. The value of the contract for the buyer at time c014-math-045 is

Indeed, the buyer uses c014-math-047 euros to buy c014-math-048 dollars. Then the buyer exchanges the dollars to euros to obtain c014-math-049 euros. The profit of the buyer is c014-math-050.

What is the price c014-math-051 of the currency forward for c014-math-052? We can replicate the currency forward using zero-coupon bonds. Let c014-math-053 be the USD price of the US zero-coupon bond and let c014-math-054 be the euro price of an European zero-coupon bond. The zero-coupon bonds are such that c014-math-055 USD and c014-math-056. Let us consider the portfolio with c014-math-057 units of US zero-coupon bonds and with c014-math-058 units of European zero-coupon bonds. The value of the portfolio at time c014-math-059 is

equation

We choose

equation

The value of the portfolio a time c014-math-060 is

Since (14.3) and (14.4) are equal, the portfolio replicates the currency forward, and the price c014-math-062 of the currency forward is equal to c014-math-063 at time c014-math-064.

The price is in euros

equation

To make c014-math-065 we need to choose the exchange rate

equation

14.1.1.3 Forward Zero-Coupon Bonds

A time c014-math-066 two parties make an agreement to exchange at a future time c014-math-067 a zero-coupon bond whose maturity is at c014-math-068, with a cash payment c014-math-069. At time c014-math-070 only the agreement is made, and at time c014-math-071 the cash payment is made and the zero-coupon bond is received. This is called a forward zero-coupon bond. Forward zero-coupon bonds are considered in Section 18.2.1.

Let c014-math-072 be the price of the forward zero-coupon bond. Let c014-math-073 and c014-math-074 be the prices of zero-coupon bonds with maturities c014-math-075 and c014-math-076. Then,

equation

Consider a portfolio with one unit of c014-math-077 and short of c014-math-078 units of c014-math-079. The portfolio has value

equation

for c014-math-080. Thus, c014-math-081 and we have replicated the forward zero-coupon bond. Thus, at time c014-math-082 the prices are equal:

equation

To make c014-math-083, we need to choose the cash payment c014-math-084 as

14.5 equation

14.1.1.4 Forward Rate Agreements

A forward rate agreement allows to change a floating Libor rate against a fixed rate. The contract is made at time c014-math-086. The reset time of the Libor is c014-math-087. At this time the Libor rate is settled to be c014-math-088, where c014-math-089 is the maturity of the Libor. The agreement stipulates that the buyer of the contract pays the payment with the fixed rate c014-math-090, and receives the payment with the Libor rate c014-math-091. Thus, the payoff for the buyer at time c014-math-092 is

where c014-math-094 is the notional, and c014-math-095 is the time between c014-math-096 and c014-math-097. Forward rate agreements are considered in Section 18.2.2.

Let c014-math-098 and c014-math-099 be zero-coupon bonds. The forward rate agreement can be replicated by the trading strategy

equation

where c014-math-100 is the strategy of buying c014-math-101 and reinvesting the payoff at time c014-math-102 with the prevailing c014-math-103-period Libor rate. Thus, the payoff of c014-math-104 at time c014-math-105 is equal to c014-math-106. Thus, the trading strategy gives the payoff (14.6). The present value of the forward rate agreement is equal to

14.7 equation

To make c014-math-108, we need to choose

equation

14.1.1.5 Backwardation and Contango

Backwardation refers to the price relation where the spot price is higher than the forward price and contango refers to the case where the spot price is lower than the forward price:

equation

At the expiration, the spot price and the futures price should be equal (to prevent arbitrage). Thus, the contango relationship means that distant delivery months trade at a greater price than near-term delivery months (the term structure is upward-sloping). Depending on the type of the underlying, either the contango or the backwardation is typical.

In the case of stock futures, or stock index futures, the theoretically fair price of a futures contract is greater than the spot price, according to (14.2). Thus, contango is typical for stock futures. However, futures prices c014-math-109 can differ from the theoretically fair price, and thus both contango and backwardation are possible. In the case of stock futures we can change the terminology in the following way:

equation

According to this terminology we do not compare futures prices to the spot prices but to the theoretically fair prices.2

When the underlying is a commodity, then there are borrowing costs for the seller of the futures contract, similarly as in the case of stock futures. In addition, there are costs for storing the commodity, for the seller of the futures contract. Thus, contango is typical for futures on commodities, and contango is typically more profound than for stock futures.

The buyer of a treasury bond future saves carrying costs (determined by the short-term rate) but looses the interest received by the bond owners (determined by the long-term rate). If the short-term rate is significantly less than the long-term rate, then the distant delivery months may trade at a lesser price than the near-term delivery months: this is backwardation.

14.1.2 The Put–Call Parity

The price of a put can always be expressed in terms of the price of a call, and conversely. We have the put–call parity:

where c014-math-111 is the common strike price of the call and put, c014-math-112 is the annualized interest rate, c014-math-113 is the expiration time, and c014-math-114 is the time from c014-math-115 to c014-math-116 in fractions of a year. It is clear that at the expiration we have c014-math-117. The put–call parity extends this relation for times c014-math-118 before the expiration time c014-math-119. Thus, we do not need to know fair values for c014-math-120 and c014-math-121 in order to have an expression for their difference.3

14.1.2.1 A Derivation of the Put–Call Parity

Consider portfolio c014-math-122 obtained by buying the call and writing the put:

equation

At the expiration, we have c014-math-123. Indeed,

equation

Let c014-math-124 be the forward contract to buy stock at time c014-math-125 with price c014-math-126. This forward contract was valued in (14.1), where we showed that

equation

The replication was obtained by buying the stock and borrowing the amount c014-math-127. At the expiration, we have c014-math-128. Since c014-math-129 with probability one, we have

equation

for all times c014-math-130 before c014-math-131, to exclude arbitrage. This is the statement (14.8) that we wanted to prove.

14.1.2.2 Bounds for the Call Price

We have bounds for the call price c014-math-132:

The upper bound c014-math-134 is obvious, since the right to buy a stock must be less valuable than the stock itself. To prove the lower bound, we can note that c014-math-135 is obvious, since the right to buy a stock involves no obligations.4 The put–call parity and the fact that c014-math-140 gives

equation

We have proved the lower bound.5

The lower bound in (14.9) implies that

for c014-math-148. This follows because c014-math-149 when c014-math-150. The inequality (14.10) leads to the definition of the time value of the option; see (2.6). The lower bound says that the value of the option is larger than the intrinsic value c014-math-151. The difference of the price of the option and its intrinsic value is called the time value of the European option. For an European put option the lower bound fails unless c014-math-152. As a consequence of the put–call parity, the time-value of a put option whose intrinsic value is large (the option is in the money), is usually negative.

14.1.3 American Call Options

An American call option has the same price as the corresponding European call option, when the stock does not pay dividends. On the other hand, an American put option has a different price than the corresponding European put option.

The put–call parity was derived for European options. However, this parity can be used to show that

14.11 equation

when the stock does not pay dividends, where c014-math-154 is the price of an American call option and c014-math-155 is the price of the corresponding European call option. We know that c014-math-156, because an American option has more rights than the corresponding European option. The lower bound in (14.9) implies

equation

where c014-math-157 is the cash flow generated by the exercise of the American call option. Since c014-math-158, an early exercise is always suboptimal and one should sell the American call option and not exercise it. Since it is not optimal to exercise the option, the possibility for the early exercise is not worth of anything, and the American call option has the same value as the European call option.6

However, an American put option is in general worth more than the corresponding European put option. For the American options we have

equation

The difference between the put and call options comes from the fact that the value of a put option increases as the price of the stock decreases. When the value of the stock decreases, then the absolute price changes become smaller. We can reach the point where the stock takes so small values that further decreases in the stock price would not give a better rate of return to the put option than the risk-free rate. At that point it is better to exercise the option and invest in the risk-free rate. With calls the situation reverses because as the stock price increases, absolute price changes increase. Pricing of American put options in the multiperiod binary model is discussed in Section 14.2.4.

14.2 Pricing in Binary Models

We consider pricing and hedging of derivatives in binary models. In a binary model the stock can at any given time go only one step higher or one step lower. Section 14.2.1 studies one-period binary models. Section 14.2.2 uses one-period binary models as building blocks for multiperiod binary models. A multiperiod binary model approximates the Black–Scholes model, as is shown in Section 14.2.3.

14.2.1 The One-Period Binary Model

First, we describe the one-period binary model. Second, we find the fair price and the optimal hedging coefficient. Third, we find the equivalent martingale measure. Fourth, we provide additional pricing and hedging formulas.

14.2.1.1 A Description of the One-Period Binary Model

In the single period model, there are time points c014-math-167 and c014-math-168. The market consist of stock c014-math-169, bond c014-math-170, and contingent claim c014-math-171.

The bond takes value 1 at c014-math-172 and value c014-math-173 at c014-math-174:

equation

where c014-math-175 is the risk-free rate.

The initial value of the stock is c014-math-176. The stock can take only two values at time 1. At c014-math-177 the value is c014-math-178 with probability c014-math-179 and the value is c014-math-180 with probability c014-math-181:

equation

where c014-math-182. We assume that

Note that if c014-math-184, then it would not be rational to invest in the bond, and if c014-math-185, then it would not be rational to invest in the stock. The “rationality” can be formalized as the absence of arbitrage.

The contingent claim c014-math-186 takes two possible values c014-math-187 and c014-math-188 at c014-math-189:

equation

For example, in the case of a call option we have c014-math-190 for some strike price c014-math-191. Then c014-math-192 and c014-math-193.

We want to find a fair value c014-math-194 for the derivative at c014-math-195. Also, we want to find the optimal hedging coefficient c014-math-196, which is used to hedge the position of the writer of the option.

14.2.1.2 Pricing and Hedging in the One-Period Binary Model

We replicate the contingent claim c014-math-197 with a portfolio

equation

The portfolio consists of c014-math-198 units of the bond and c014-math-199 units of the stock. The portfolio takes values

equation

To obtain c014-math-200 we need to choose c014-math-201 and c014-math-202 so that

equation

The first equation leads to

Inserting this value of c014-math-204 to the second equation gives

The law of one price implies that the arbitrage-free price c014-math-206 of the contingent claim equals the value c014-math-207 of the replicating portfolio at time c014-math-208: c014-math-209, and thus

14.15 equation

where c014-math-211 and c014-math-212 are defined in (14.13) and (14.14).

The number c014-math-213 of bonds can be written as c014-math-214. Since c014-math-215, we can write

We can interpret the replicating portfolio from the point of view of the writer of the option as the portfolio where the writer receives the option premium c014-math-217, invests c014-math-218 in the bank account, borrows the amount c014-math-219 from the bank, and invests c014-math-220 in the stock.

14.2.1.3 The Equivalent Martingale Measure

Price c014-math-221 can be written as the expectation with respect to the equivalent martingale measure. We have

where

14.18 equation

and c014-math-224 means the expectation with respect to the probability measure c014-math-225 with

equation

Probability measure c014-math-226 is called a risk-neutral measure because

equation

and it is called a martingale measure because

equation

Note that condition (14.12) guarantees that c014-math-227, so that c014-math-228 is equivalent to c014-math-229.

14.2.1.4 Further Pricing and Hedging Formulas

It is of interest that the price can be written as

where

equation

Indeed, similarly to (14.13), we get

equation

which can be combined with (14.13) to get

equation

Combining this formula for c014-math-231 with the formula c014-math-232 shows (14.19). In (14.19), we have written the derivative price as a discounted expectation of the derivative price, with an additional correction term.

The hedging coefficient can be written as

To derive (14.20), note that

equation

Thus, (14.20) is equal to (14.14). We can also write

where c014-math-235 is the gross return. This way of writing the hedging coefficient appears in (16.10), where quadratic hedging is considered.

14.2.2 The Multiperiod Binary Model

We start with a description of the multiperiod binary model, and then proceed to the pricing formulas, derive the equivalent martingale measure, and give hedging formulas.

14.2.2.1 A Description of the Multiperiod Binary Model

The market consists of stock c014-math-236, bond c014-math-237, and contingent claim c014-math-238. We define the discrete time processes

14.22 equation

Note that in Section 13.2 we denoted the time steps of the discrete time markets by c014-math-240. We have changed the notation, because we construct a time series that approximates the geometric Brownian motion on c014-math-241. The approximation is done by dividing interval c014-math-242 to c014-math-243 periods of equal lengths and letting c014-math-244.

The bond takes value c014-math-245 at step c014-math-246, where c014-math-247 is the annualized interest rate, and c014-math-248 is the time between two periods in fractions of a year.

At step c014-math-249, the stock can take c014-math-250 values

where

equation

The stock price c014-math-252 is a random variable with

where c014-math-254,

The stochastic process of stock prices can be described by a recombining binary tree, as in Figure 14.1, where c014-math-255. At step c014-math-256, the stock takes value c014-math-257 (in the figure c014-math-258). If the value of the stock at time c014-math-259 is c014-math-260, then at step c014-math-261 the stock can take values c014-math-262 and c014-math-263, so that

equation

The probabilities of the up and down movements are c014-math-264 and c014-math-265:

equation
Illustration of A recombining binary tree.

Figure 14.1 A recombining binary tree. There are c014-math-266 periods and the initial value is c014-math-267.

The derivative can take at the expiration c014-math-268 values c014-math-269, c014-math-270. The random variable c014-math-271 takes the value c014-math-272, when c014-math-273. For example, when the contingent claim is the call option with c014-math-274, where c014-math-275 is the strike price, then c014-math-276. We have

We want to find the arbitrage-free price c014-math-278 of the derivative at time c014-math-279.

14.2.2.2 Pricing in the Multiperiod Binary Model

The evolution of the stock price in a multiperiod binary model has been described using a recombining binary tree. The price in a multiperiod binary model can be found by backward induction. We know the price at the expiration, when c014-math-280. We can use the single period model to calculate the price at step c014-math-281, and go backwards step by step to obtain the price at step c014-math-282. The price c014-math-283 of the derivative is calculated using the backwards induction with the following steps.

  1. 1. At the expiration step c014-math-284, the prices of the derivative are given by c014-math-285, c014-math-286.
  2. 2. Let the current step be c014-math-287 and the current node c014-math-288. Then the two possible prices for the derivative are c014-math-289 and c014-math-290. We can use the single period model to calculate the price at step c014-math-291. We get the price from (14.17) as
    14.26 equation
  3. where
    equation
  4. and the second equality follows from c014-math-293 and c014-math-294.

We have described a recursive algorithm for the computation of the price, but we can also obtain the following explicit expression for the price.

The arbitrage-free price of the derivative is obtained not only at step c014-math-302, but the price is obtained at all steps c014-math-303. In fact, the price of the derivative, under the condition that the stock price at step c014-math-304 is c014-math-305, is given by the formula

14.2.2.3 The Equivalent Martingale Measure

Let us define probability measure c014-math-307 by

equation

where c014-math-308. Measure c014-math-309 is obtained from the physical measure c014-math-310, defined in (14.24), by replacing the probability c014-math-311 of an up-movement by probability c014-math-312. The price in (14.27) can be written as the expectation

where c014-math-314 is the random variable taking value c014-math-315, when c014-math-316 takes value c014-math-317. The measure c014-math-318 is called the equivalent martingale measure, or the risk-neutral measure.

The price c014-math-319 of the derivative given in (14.28), under the condition that the stock price at step c014-math-320 is c014-math-321, can be written as

Accordingly, we can define random variables

equation

Defining

equation

we obtain a more elegant formula

14.2.2.4 Hedging in the Multiperiod Binary Model

The following theorem gives the hedging coefficients of the replicating portfolio. The replication means that the wealth process of the trading strategy obtains the value of the derivative with probability one, or equivalently, the value process of the trading strategy obtains the discounted value of the derivative with probability one. The value process of a self-financing trading strategy is given in (13.8).

There are other ways to write the hedging coefficient. We have from the formula (14.21) of the one-period model that

equation

where c014-math-335, c014-math-336, and c014-math-337. Here c014-math-338 and c014-math-339 mean conditional variance and covariance, conditional on c014-math-340, and under probability measure c014-math-341, which is the equivalent martingale measure.

We can also write the hedging coefficient as

Let us prove (14.33). We have from (14.31) that

equation

Thus,

equation

We have shown (14.33).

An additional formula for the hedging coefficient is

14.34 equation

In fact, c014-math-344. Thus,

equation

14.2.3 Asymptotics of the Multiperiod Binary Model

We start with the asymptotic normality of the logarithmic returns in the multiperiod binary model, then show the convergence of the arbitrage-free prices in the multiperiod binary model to the Black–Scholes prices, and finally show the convergence of the hedging coefficients in the multiperiod binary model to the Black–Scholes hedging coefficients.

14.2.3.1 Choice of the Parameters

Let

equation

We consider asymptotics when c014-math-345. We choose the up and down factors as

where c014-math-347. We choose the probabilities of the up and down movements as

equation

where c014-math-348. With these choices the logarithmic return c014-math-349 converges in distribution to the normal distribution c014-math-350, where c014-math-351. Note that when we choose c014-math-352 and c014-math-353 as in (14.35), then the probability c014-math-354 of the up movement in the risk-neutral distribution is

14.36 equation

14.2.3.2 Asymptotic Normality in the Multiperiod Binary Model

We show that the distribution of the stock price in the multiperiod binary model converges in distribution to a log-normal distribution, as the number of steps increases.

In the c014-math-356-period binary model the stock price c014-math-357, c014-math-358, can be written as

14.37 equation

where c014-math-360 are such i.i.d. random variables that c014-math-361 when c014-math-362 is a result of an up-movement and c014-math-363 when c014-math-364 is a result of a down-movement, so that

equation

It holds that

as c014-math-366, where

equation

We show a slightly more general result: For c014-math-367

as c014-math-369, where c014-math-370 is such that c014-math-371, as c014-math-372. In particular, we can choose c014-math-373. Now (14.38) follows as a special case when we choose c014-math-374 and c014-math-375.

We denote below c014-math-376. Let us denote c014-math-377. We can write

equation

where

equation

with c014-math-378. Now it holds that c014-math-379. We have that

  1. 1. c014-math-380,
  2. 2. c014-math-381,
  3. 3. c014-math-382,

as c014-math-383. Because c014-math-384, the claim (14.39) follows from items 1–3.

To prove item 1, we note that

equation

Thus,

equation

By the choice of c014-math-385, it holds c014-math-386 as c014-math-387, and thus

equation

as c014-math-388. Thus, item 1 follows by the central limit theorem.

To prove item 2, we use the fact that c014-math-389, which implies that c014-math-390 as c014-math-391. Thus, the weak law of the large numbers implies item 2.

To prove item 3, we note that

equation

Item 2 implies that c014-math-392 and since c014-math-393, we have that c014-math-394, so that c014-math-395.

14.2.3.3 Convergence of the Price

The arbitrage-free prices (14.27) in the multiperiod binomial model are called the Cox–Ross–Rubinstein prices. We show that the Cox–Ross–Rubinstein put and call prices converge to the Black–Scholes put and call prices, as c014-math-396. This is done in two steps. First, we show that the put and call prices in the multiperiod binary model converge to the expected values of the option pay-offs, when the expectation is taken with respect to a log-normal distribution. Second, we calculate closed-form expressions for the expected values.

Bounded Continuous Payoff Functions

A fundamental theorem about weak convergence states that if

  1. 1. c014-math-397,
  2. 2. c014-math-398 is bounded and continuous,

then

equation

as c014-math-399; see Billingsley (2005, Theorem 25.8, p. 335).

  1. 1. First, we apply the weak convergence in (14.38) to obtain
  2. where the stock price c014-math-401 is a random variable taking values c014-math-402 with probabilities
    equation
  3. and c014-math-403.
  4. 2. Second, we have noted in (14.29) that a Cox–Ross–Rubinstein price satisfies
  5. where the expectation is with respect to the risk-neutral measure c014-math-405, and c014-math-406 is a random variable taking values c014-math-407, defined by (14.25).

Let c014-math-408 be such function that c014-math-409 Then the option payoff can be written as

equation

If function c014-math-410 is continuous and bounded, then (14.40) and (14.41) imply that

14.42 equation

where the distribution of c014-math-412 is defined by

and we used the fact that c014-math-414, because c014-math-415.

Unbounded Continuous Payoff Functions

The fundamental theorem about weak convergence applies for any sequence converging weakly. However, in our case we are interested in the special case of the convergence of a binomial distribution toward a log-normal distribution. In this special case Föllmer and Schied (2002, Proposition 5.39, p. 265) notes that the convergence of expectations can be proved also when we relax the condition of the boundedness. Let c014-math-416 be measurable, almost everywhere continuous, and

14.44 equation

Then,

equation

as c014-math-418, where c014-math-419 satisfies (14.40) and c014-math-420 is distributed as (14.43).

Put Prices

The payoff function of a put option is

equation

where c014-math-421 is the strike price. Function c014-math-422, c014-math-423, is bounded and continuous. Thus, the Cox–Ross–Rubinstein price c014-math-424 of an European put option converges to the expectation:

The right hand side is equal to the Black–Scholes put price, as shown in (14.54).

Call Prices

The payoff function of a call option is

equation

where c014-math-426 is the strike price. Function c014-math-427, c014-math-428, is continuous but not bounded. Thus, the convergence of the Cox–Ross–Rubinstein price to the expected value cannot be inferred similarly as in the case of put options. However, we can apply the put-call parity in (14.8) to conclude that the Cox–Ross–Rubinstein prices have to satisfy

equation

where c014-math-429 is the Cox–Ross–Rubinstein call price, and c014-math-430 the put price. Since we have shown (14.45), it holds that

The right hand side is equal to the Black–Scholes call price, as shown in (14.52).

Calculation of the Expectations

We have proved in (14.45) and (14.46) that the arbitrage-free put and call prices in the multiperiod binary model approach the expected values of the option payoffs, when the expectations are with respect to the equivalent martingale measure (the risk neutral log-normal model). Thus, we want to calculate the expected values of the put and call payoffs. The expected values are the Black–Scholes prices, which we denote

14.47 equation

and

14.48 equation

where the expectations are taken with respect to the distribution of c014-math-434 defined by

where c014-math-436 and

equation

Note that we have discussed the log-normal distribution in (3.50).

The density of the standard normal distribution is c014-math-437, where c014-math-438. Then,

equation

where

equation

By writing c014-math-439 we have

equation

Thus,

where c014-math-441 is the distribution function of the standard normal distribution. Since c014-math-442,

where

equation

because c014-math-444 and c014-math-445. This leads to the call price

Similarly,

which leads to the put price

We have calculated explicit expressions for the call and put prices at time c014-math-449, when the stock price is c014-math-450. The Black–Scholes call and put prices at time c014-math-451, when the stock price is c014-math-452, are given by

and

where the expectations are taken with respect to the distribution of c014-math-455 defined by

equation

where c014-math-456 and c014-math-457. The corresponding explicit expressions are given in (14.58) and (14.59), and the Black–Scholes prices are discussed in Section 14.3.1.

14.2.3.4 Convergence of the Hedging Coefficient

We show that the hedging coefficients of puts and calls in the multiperiod binary model converge to the Black–Scholes hedging coefficients. The Black–Scholes hedging coefficients are called deltas, and they are obtained by differentiating the Black–Scholes prices with respect to the stock price.

Let the initial time and the initial stock price be

equation

Let c014-math-458 be either the Black–Scholes call price or the put price at time c014-math-459, when the stock price at time c014-math-460 is c014-math-461, and the expiration is at time c014-math-462. These prices are written as expectations in (14.55) and (14.56), and in a more explicit form in (14.58) and (14.59).

Let us consider step c014-math-463 of the multiperiod binary model. Let c014-math-464 be one of the possible stock prices at step c014-math-465, where c014-math-466. These prices were defined in (14.23). Let c014-math-467 be such that

equation

as c014-math-468. In particular, we can choose c014-math-469, where c014-math-470 is the largest integer c014-math-471. Choose

equation

Then

equation

as c014-math-472.

The hedging coefficient of the multiperiod binary model was written in (14.32) as

equation

where c014-math-473 and c014-math-474 are the two possible prices of the derivative at step c014-math-475, when the value of the stock at step c014-math-476 is c014-math-477.7 Now it holds that

where c014-math-479 is the derivative of the Black–Scholes price with respect to stock price.

To show (14.57), note that from (14.46) and (14.45), and the continuity of the functions c014-math-480 we have that

equation

where c014-math-481 means that c014-math-482. Thus,

equation

Also, c014-math-483, c014-math-484, and c014-math-485, as c014-math-486.

14.2.3.5 Rate of Convergence

Figure 14.2 illustrates the convergence of (a) the ratio of the Cox–Ross–Rubinstein call price to the Black–Scholes call price, and (b) the ratio of the Cox–Ross–Rubinstein call hedging coefficient to the Black–Scholes call hedging coefficient. The ratios are plotted as a function of the number c014-math-487 of the steps in the multiperiod binary model, where c014-math-488. The moneyness c014-math-489 is 0.9 (green), 0.95 (black), and 1 (red). The annualized volatility is c014-math-490, the interest rate is c014-math-491, and the time to maturity is 1 month: c014-math-492. We see that the convergence of the at-the-money options is faster than the convergence of the out-of-the-money options.

Illustration of Convergence of the Cox-Ross-Rubinstein price and hedging coefficient.

Figure 14.2 Convergence of the Cox–Ross–Rubinstein price and hedging coefficient. Plotted are (a) the ratios of the Cox–Ross–Rubinstein call price to the Black–Scholes price and (b) the ratios of the Cox–Ross–Rubinstein call hedging coefficient to the Black–Scholes hedging coefficient as a function of c014-math-493. The three curves show the cases where the moneyness is c014-math-494 (green), c014-math-495 (black), and c014-math-496 (red).

14.2.3.6 Asian and Knock-Out-Options

We have proved the weak convergence at one point: Let c014-math-497 and c014-math-498. Then

equation

where c014-math-499 is the price of the stock at step c014-math-500 in the multiperiod binomial model, and c014-math-501 is the price of the stock at time c014-math-502 in the geometric Brownian motion model. If the option payoff is c014-math-503, then the price in the multiperiod binomial model converges to the price of the option whose payoff is c014-math-504 in the geometric Brownian motion model. Asian options and knock-out options depend on values of stocks in more than one point.

Asian Options

In the case of Asian options, the option payoffs can be written as

equation

where c014-math-505 are steps, and c014-math-506 are fixed time points. For example, in the case of an Asian call option

equation

The corresponding prices converge for a suitable c014-math-507, when

equation
Knock-Out Options

The payoffs of knock-out options depend on the trajectory of the prices through

equation

Let us divide the time interval c014-math-508 into c014-math-509 subintervals. Denote the c014-math-510 boundaries of the intervals by

equation

Points c014-math-511 fill the interval c014-math-512 asymptotically. We can define a continuous time process c014-math-513, c014-math-514, by linearly interpolating c014-math-515: define

equation

when c014-math-516. Geometric Brownian motion c014-math-517 is defined in (5.62). We can show that process c014-math-518 converges weakly to the geometric Brownian motion

equation

where c014-math-519 is the standard Brownian motion. The weak convergence

equation

as c014-math-520, happens in the metric space c014-math-521 of the continuous functions on c014-math-522. The prices of the options whose payoffs are

equation

converge for suitable c014-math-523.

14.2.4 American Put Options

An American put option has a different price than the corresponding European put option, and the price of an American put option does not have a closed-form expression in the multiperiod binary model. However, an American call option has the same price as the corresponding European call option, when the stock does not pay dividends (see Section 14.1.3). Thus, the American call options can be priced similarly as the European call options using the Black–Scholes prices or the recombining binary trees.

The American put options have to be priced by taking into account the possibility of an early exercise. We can use the recombining binary tree to price the American put options. At every node of the tree we consider whether it is better to exercise or to keep the option for a future exercise. We are not able to obtain a closed-form formula for the price of the American put options, but we obtain an algorithm for the computation of the price. First, the single period binary model is studied. Second, the multiperiod binary model leads to the final algorithm.

14.2.4.1 American Put Options in the One-Period Binary Model

In the one-period binary model, the American put option can be exercised at time c014-math-524 or at time c014-math-525. Let us denote with c014-math-526 the value of the American put option at time c014-math-527 and let us denote with c014-math-528 the value of the European put option at time c014-math-529. An arbitrage argument shows that

equation

The value of an European put option can be obtained from (14.17) as

equation

where

equation

with

equation

14.2.4.2 American Put Options in the Multiperiod Binary Model

The price of an American put option is determined in the c014-math-530-step binomial model by recursion. Remember that in the c014-math-531-step binomial model the possible prices at step c014-math-532, c014-math-533, are

equation

with

equation

The recursive steps are the following.

  1. 1. At time c014-math-534 the prices of the American put option are given by
    equation

    c014-math-535.

  2. 2. At time c014-math-536, c014-math-537, when the stock has price c014-math-538, we know from the previous steps of the algorithm that the two possible prices for the derivative at time c014-math-539 are c014-math-540 and c014-math-541. We can use the single period model to calculate the price at time c014-math-542. We get the price from the one-step binary model as
    equation
  3. where
    equation
  4. and
    equation
  5. with c014-math-543.

14.3 Black–Scholes Pricing

First we describe the properties of Black–Scholes call and put prices, second we discuss implied volatility, third we describe various ways to derive the Black–Scholes prices, and fourth we give Black–Scholes formulas for options on forwards, for fixed income options, and for currency options.

14.3.1 Call and Put Prices

The Black–Scholes price of the call option at time c014-math-544, with strike price c014-math-545, and with the maturity date c014-math-546, is equal to

where c014-math-548 is the stock price at time c014-math-549, c014-math-550 is the annualized risk-free rate,

equation

and c014-math-551 is the distribution function of the standard Gaussian distribution. The time c014-math-552 to expiration is expressed in fractions of a year. The put price is equal to

Note that it can be convenient to write

equation

The Black–Scholes price is derived under the assumption of a log-normal distribution of the stock price: It is assumed that at time c014-math-554

equation

where c014-math-555, c014-math-556 is the drift, and c014-math-557 is the volatility. Note that under the risk-neutral measure c014-math-558. The volatility c014-math-559 is the only unknown parameter that need to be estimated, since c014-math-560 does not appear in the price formula.

14.3.1.1 Computation of Black–Scholes Prices

For the application of the Black–Scholes formula the time c014-math-561 is taken as the time in fractions of year. For example, when the time to expiration is 20 trading days, then c014-math-562. Alternatively, when time to expiration is 20 calendar days, then c014-math-563.

The risk-free rate c014-math-564 is expressed as the annualized rate.

The only unknown parameter c014-math-565 has to be estimated. Let c014-math-566 be an equally spaced sample of stock prices and let us denote c014-math-567 for c014-math-568. We assume that c014-math-569 c014-math-570, are i.i.d. c014-math-571, so that the stock prices satisfy (3.50). We can estimate c014-math-572 with the sample variance

equation

where c014-math-573. Then an estimator of c014-math-574 is

For example, if we sample stock prices daily, then c014-math-576 and c014-math-577.8 If we sample stock prices monthly, then c014-math-579 and c014-math-580 The normalized sample standard deviation in (14.60) is called the annualized sample standard deviation.

14.3.1.2 Characteristics of Black–Scholes Prices

We study the qualitative behavior of the Black–Scholes prices as a function of five parameters c014-math-581, c014-math-582, c014-math-583, c014-math-584, and c014-math-585.

The prices of calls and puts increase as c014-math-586 increases. We have

equation

and

equation

which are the bounds derived from the put–call parity in (14.9). The prices of calls and puts increase as the time to maturity c014-math-587 increases. The price of a call increases as c014-math-588 increases and the price of a call decreases as c014-math-589 increases, but for puts the relations reverse. The price of a call increases as the interest rate c014-math-590 increases but the price of a put decreases as the interest rate c014-math-591 increases.

Figure 14.3 shows Black–Scholes prices for calls and puts as a function of the call moneyness c014-math-592. The call prices increase and the put prices decrease as a function of moneyness. Panel (a) shows the cases of annualized volatility c014-math-593 (black), c014-math-594 (red), and c014-math-595 (green). The time to maturity is 20 trading days and the interest rate is c014-math-596. Panel (b) shows the cases of interest rates c014-math-597 (black), c014-math-598 (red), and c014-math-599 (green). The time to maturity is 20 trading days and the annualized volatility is c014-math-600. We see from panel (a) that increasing volatility increases the prices, both for call and puts. The effect of increasing time to maturity is similar. We see from panel (b) that for calls increasing interest rates increases prices, but for puts increasing interest rates decreases prices.

Illustration of Call and put prices as a function of call moneyness S/K.

Figure 14.3 Call and put prices as a function of call moneyness c014-math-601. The call prices increase and the put prices decrease as a function of moneyness. (a) The annualized volatility is c014-math-602 (black), c014-math-603 (red), and c014-math-604 (green); (b) the interest rate is c014-math-605 (black), c014-math-606 (red), and c014-math-607 (green).

Figure 14.4 shows that the call and put prices are not symmetric. The ratios of call prices to put prices are shown as a function of the call moneyness c014-math-608: We show the functions

equation

where c014-math-609 is the current stock price. The strike prices for the calls take values c014-math-610, and the corresponding strike prices for the puts are c014-math-611. Panel (a) shows the cases of annualized volatility c014-math-612 (black), c014-math-613 (red), and c014-math-614 (green). The interest rate is c014-math-615. Panel (b) shows the cases of interest rates c014-math-616 (black), c014-math-617 (red), and c014-math-618 (green). The annualized volatility is c014-math-619. The time to maturity is 20 trading days in both panels. We see that the call prices are higher than the put prices for out-of-the-money options. This is related to the asymmetry of the log-normal distribution (see Figure 3.11). For at-the-money options the difference between the call and the put prices is not large. Increasing the volatility makes the ratios of the call and put prices closer to one, and decreasing the interest rate makes the ratios of the call and put prices closer to one.

Illustration of Ratios of call prices to put prices as a function of moneyness S/K.

Figure 14.4 Ratios of call prices to put prices as a function of moneyness c014-math-620. (a) Annualized volatility is c014-math-621 (black), c014-math-622 (red), and c014-math-623 (green); (b) interest rate is c014-math-624 (black), c014-math-625 (red), and c014-math-626 (green).

14.3.1.3 Black–Scholes Prices and Volatility

To apply Black–Scholes prices we need to estimate the volatility. We study how the Black–Scholes prices change when the volatility estimate changes. We apply the data of S&P 500 daily prices, described in Section 2.4.1.

Figure 14.5 shows time series of Black–Scholes call prices. The volatility is equal to the sequential annualized sample standard deviation. Panel (a) shows call prices with moneyness c014-math-627 and panel (b) shows call prices with moneyness c014-math-628. The time to expiration is either 20 trading days (black curves) or 30 trading days (red curves). The risk-free rates are deduced from 1-month Treasury bill rates. We can see that the time series of Black–Scholes prices follow closely to the time series of sequential standard deviations in Figure 7.5.

Illustration of Time series of Black-Scholes call prices.

Figure 14.5 Time series of Black–Scholes call prices. (a) Moneyness c014-math-629 and (b) moneyness c014-math-630. Call prices are computed using the sequential sample standard deviation. The time to expiration is 20 trading days (black) or 30 trading days (red).

Figure 14.6 studies Black–Scholes prices when the volatility is equal to the sequentially estimated GARCH(c014-math-631) volatility. Panel (a) shows time series of call prices with moneyness c014-math-632 (black) and c014-math-633 (green). The time to expiration is 20 trading days. Panel (b) shows kernel density estimates of the distributions of the prices. The horizontal and the vertical lines show the Black–Scholes prices when the volatility is the annualized sample standard deviation computed from the complete sample. The risk-free rate is zero.

Illustration of Black-Scholes prices with GARCH-volatility.

Figure 14.6 Black–Scholes prices with GARCH-volatility. (a) Time series of prices; (b) kernel estimates of the density of the collection of prices. The moneyness is c014-math-634 (black) and c014-math-635 (green).

14.3.1.4 The Greeks

The greeks are defined by differentiating the option price with respect to the price of the underlying, the time to the expiration, the interest rate, or the volatility. In this section, we denote the Black–Scholes call and put prices by

equation

where c014-math-636 is the current time, c014-math-637 is the current price of the underlying, c014-math-638 is the strike price, c014-math-639 is the time of the expiration, c014-math-640 is the volatility, and c014-math-641 is the interest rate. Sometimes we leave out some of the arguments and denote, for example, c014-math-642.

The Delta

The delta is the derivative of the price function with respect to the underlying. The delta is the hedging coefficient in the Black–Scholes hedging, as discussed in Section 14.4. The call and put delta are

We have that

equation

In fact, if c014-math-644 increases, then the price of the call increases and the price of the put decreases: c014-math-645 and c014-math-646. The absolute value of the change in the value of a call or a put cannot exceed the absolute value of the change in the underlying: c014-math-647 and c014-math-648.

The call delta is equal to

and the put delta equal to

Let us calculate the call delta. The delta of a call is given in (14.62) because

equation

and finally9

equation

Figure 14.7 shows the call delta as a function of moneyness c014-math-652. In panel (a) the annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month. In panel (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The solid lines show the case of interest rate c014-math-653 and the dashed lines show the case of interest rates c014-math-654. The more interesting part is the moneyness c014-math-655. In this region the deltas are c014-math-656 in both panels. Panel (a) shows that when moneyness is c014-math-657, then a larger volatility leads to a larger delta. Panel (b) shows that increasing time to maturity has a similar qualitative effect as increasing volatility.

Graph for Call deltas.

Figure 14.7 Call deltas. Call deltas are shown as a function of moneyness c014-math-658. (a) The annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month; (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The interest rate is c014-math-659 (solid lines) and c014-math-660 (dashed lines).

Figure 14.8 shows a time series of Black–Scholes deltas using the data of S&P 500 daily prices, described in Section 2.4.1. The daily prices are used to create a time series with the sampling frequency of 20 and 30 trading days. Panel (a) shows call prices with moneyness c014-math-661 and panel (b) shows call prices when moneyness c014-math-662. The time to expiration is either 20 trading days (black curves) or 30 trading days (red curves). The volatility is equal to the sequential sample standard deviation. The risk-free rates are deduced from 1-month Treasury bill rates. The corresponding time series of Black–Scholes prices is given in Figure 14.5.

Illustration of Time series of Black-Scholes call deltas.

Figure 14.8 Time series of Black–Scholes call deltas. (a) Moneyness c014-math-663 and (b) moneyness c014-math-664. Call deltas are computed using the sequential sample standard deviation. The time to expiration is 20 trading days (black) or 30 trading days (red).

The Gamma

The gamma is the second derivative of the price function with respect to the underlying:

equation

The price functions are convex with respect to c014-math-665 and thus

equation

The call gamma and the put gamma are given by

equation

Figure 14.9 shows the call gamma as a function of moneyness c014-math-666. In panel (a) the annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month. In panel (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The solid lines show the case of interest rate c014-math-667 and the dashed lines show the case of interest rates c014-math-668.

Illustration of Call gammas.

Figure 14.9 Call gammas. Call gammas are shown as a function of moneyness c014-math-669. (a) The annualized volatility is 10% (black) and 20% (red) with the time to expiration 1 month; (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The interest rate is c014-math-670 (solid lines) and c014-math-671 (dashed lines).

Theta

The theta is the derivative of the price function with respect to time:

equation

As c014-math-672 increases the value of the option decreases (everything else being equal) and thus

equation

The call theta is equal to

equation

and the put theta is equal to

equation

Figure 14.10 shows the call theta as a function of moneyness c014-math-673. In panel (a) the annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month. In panel (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The solid lines show the case of interest rate c014-math-674 and the dashed lines show the case of interest rates c014-math-675.

Illustration of Call thetas.

Figure 14.10 Call thetas. Call thetas are shown as a function of moneyness c014-math-676. (a) The annualized volatility is 10% (black) and 20% (red) with the time to expiration 1 month; (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The interest rate is c014-math-677 (solid lines) and c014-math-678 (dashed lines).

Vega

The vega is the derivative of the price function with respect to volatility:

equation

The call vega and the put vega are equal to

equation

Figure 14.11 shows the call vega as a function of moneyness c014-math-679. In panel (a) the annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month. In panel (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The solid lines show the case of interest rate c014-math-680 and the dashed lines show the case of interest rates c014-math-681.

Illustration of Call vega.

Figure 14.11 Call vega. Call vegas are shown as a function of moneyness c014-math-682. (a) The annualized volatility is 10% (black) and 20% (red) with the time to expiration 1 month; (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The interest rate is c014-math-683 (solid lines) and c014-math-684 (dashed lines).

Rho

The rho is the derivative of the price function with respect to interest rate:

equation

The call rho is equal to

equation

and the put rho is equal to

equation

Figure 14.12 shows the call rho as a function of moneyness c014-math-685. In panel (a) the annualized volatility is 10% (black) and 20% (red) with time to expiration 1 month. In panel (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The solid lines show the case of interest rate c014-math-686 and the dashed lines show the case of interest rates c014-math-687.

Illustration of Call rho.

Figure 14.12 Call rho. Call rhos are shown as a function of moneyness c014-math-688. (a) The annualized volatility is 10% (black) and 20% (red) with the time to expiration 1 month; (b) the time to expiration is 1 month (black) and 3 months (red) with volatility 10%. The interest rate is c014-math-689 (solid lines) and c014-math-690 (dashed lines).

14.3.2 Implied Volatilities

Implied volatilities can be derived both from call prices or from put prices. Let c014-math-691 be a Black–Scholes call price. The mapping

is bijective and can be inverted. Let us denote by c014-math-693 the inverse of mapping (14.64). When we observe a market price c014-math-694 for a call option, then

equation

is the implied volatility of the option. The implied volatility of a put option can be defined similarly.

14.3.2.1 Quoting Option Prices

It is helpful to quote option prices using implied volatilities. Option prices and implied volatilities are in a bijective correspondence, but it is easier to compare the prices of options with different maturities and strike prices using the implied volatilities than using the market prices. This is analogous to expressing the bond prices with annualized rates instead of quoted prices.10 The implied volatilities can be used to quote prices even when we do not think that the Black–Scholes prices are fair prices, similarly as the bond rates can be defined using various conventions.

14.3.2.2 The Volatility Surface

If the Black–Scholes model describes the true distribution of the asset prices, and if the market prices coincide with the Black–Scholes prices, then the implied volatilities of options with different strike prices and with different maturities are all equal. However, in practice the implied volatilities are different for the options with different strike prices and with different maturities.

The volatility surface gives for each strike price and for each maturity the corresponding implied volatility. Let c014-math-700 be the market prices of call (put) options with strike prices c014-math-701 and expiration dates c014-math-702, where c014-math-703, c014-math-704. The options are otherwise similar. The volatility surface is the function

equation

where c014-math-705 is the time to the expiration.

The volatility surface is typically not a constant function. Instead, for a fixed maturity c014-math-706, function

equation

is typically u-shaped (smile) or skew (one sided smile). Instead of the strike c014-math-707 one may take as the argument the moneyness c014-math-708 (or c014-math-709), or the delta of the options.

Options written on equity indices yield often skews. This might be due to the fact that a crash in stock markets leads to increased volatility, whereas a rise in stock markets is not usually involved with an increased volatility. Options on various interest rates yield more monotonous one sided smiles than the equity indices.

Currency markets yield often symmetric smiles. Currency markets are more symmetric than stock markets because a big movement in either direction results in an increased volatility (it is always a crash for one of the currencies).

The smile tends to flatten out with maturity. This might be due to the better Gaussian approximation when the maturity is longer.

14.3.2.3 Pricing of Options Using Implied Volatilities

Out-of-the-money options can be priced in the following way. (1) Find the implied volatility of at-the-money options. (2) Adjust the implied volatility (using experience) to get a new volatility. (3) Calculate the price of the out-of-the-money option using the new volatility.

14.3.2.4 VIX Index

The VIX index of CBOE (Chicago Board Options Exchange) uses prices of options to derive the volatility that is expected by the markets. Section 6.3.1 contains a discussion of the VIX index and Figure 6.5 shows a time series of the VIX index. Let us derive the formula of the VIX index.

It is assumed that the stock price follows a geometric Brownian motion, as defined in (5.62). That is,

equation

where c014-math-710 is the standard Brownian motion, c014-math-711, and c014-math-712. Under the equivalent martingale measure c014-math-713, where c014-math-714 is the yearly risk-free interest rate. Thus, for the equivalent martingale measure we have

equation

We can solve for c014-math-715 to get

equation

A Taylor expansion gives

equation

where c014-math-716. Thus,

equation

Theorem 13.2 implies that when the expectation is taken with respect to a risk-neutral measure, then

equation

where c014-math-717 and c014-math-718 are the arbitrage-free prices of the call and the put. Under the risk-neutral measure

equation

where c014-math-719 is the futures price, as given in (14.2). Thus,

equation

We arrive at the variance formula

The variance formula (14.65) was derived in Demeterfi et al. (1999). CBOE uses the approximation

equation

CBOE:s VIX index is defined as11

14.66 equation

where c014-math-722 is the midpoint of the bid-ask spread for the option, and c014-math-723. The calculation is done for two expiration dates, and the final index value is a weighted average of these. Note that the put–call parity (14.8) gives the equality c014-math-724.

14.3.3 Derivations of the Black–Scholes Prices

We have derived the Black–Scholes prices as the limits of the prices in the multiperiod binary model. Now we describe shortly the martingale derivation of the Black–Scholes prices, derivation of the prices using the Black–Scholes differential equation, and the derivation of the prices using the put–call parity.

14.3.3.1 Martingale Derivation

The second fundamental theorem of asset pricing says that an arbitrage-free market model is complete if and only if there exists exactly one equivalent martingale measure. We have stated this theorem for the discrete time model in Theorem 13.3.

The Black–Scholes prices can be derived as a corollary of the second fundamental theorem of asset pricing: If the Black–Scholes market model is arbitrage-free and complete, then the Black–Scholes prices are the discounted expectations with respect to the equivalent martingale measure.

Shiryaev (1999, p. 710) states in a continuous time framework that if there exists a unique equivalent martingale measure, then the unique arbitrage-free price of the option is the discounted expected value of the payoff with respect to the unique equivalent martingale measure.

We give a sketch of some elements of the derivation of the Black–Scholes prices directly from the second fundamental theorem of asset pricing. Details can be found in Shiryaev (1999, p. 739).

The Black–Scholes model assumes that the stock price follows the geometric Brownian motion, as defined in (5.62). The stock price satisfies

14.67 equation

and the bank account satisfies

equation

where c014-math-726.12 Thus,

equation

The Girsanov's theorem was stated in (5.64). We apply Girsanov's theorem with the constant function c014-math-727. Then,

is a Brownian motion with respect to measure c014-math-729, defined by c014-math-730, where

equation

Measure c014-math-731 is the unique martingale measure that is equivalent to c014-math-732; see Shiryaev (1999, p. 708). Thus, the price of the call option c014-math-733 is

equation

We need to find the distribution of c014-math-734 under c014-math-735. From (14.68) we obtain that

equation

Thus,

equation

14.3.3.2 The Black–Scholes Differential Equation

Black and Scholes (1973) and Merton (1973) derived the Black–Scholes price by solving a differential equation. The Black–Scholes partial differential equation is

where c014-math-737 is the value of the option at time c014-math-738, c014-math-739 is the theta of the option, c014-math-740 is the delta of the option, c014-math-741 is the gamma of the option, and c014-math-742. When c014-math-743 is the value of a call option, then the solution is found under the boundary condition

equation

The price of the option is c014-math-744. The differential equation is solved, for example, in Shiryaev (1999, p. 746).

The Black–Scholes partial differential equation can be derived heuristically in the following way. Itô's lemma (5.61) applied to the function c014-math-745 gives

Assume that value c014-math-747 of the option is replicated by the portfolio

equation

that is,

equation

The assumptions c014-math-748 and c014-math-749 imply

Equating (14.70) and (14.71) gives c014-math-751, that is,

equation

This means that we want a perfect replication of c014-math-752. Choose c014-math-753, which is called delta hedging. This makes c014-math-754 to disappear and leads to the differential equation (14.69). Delta hedging has removed all uncertainty, and we have obtained a perfect hedge.

14.3.3.3 Derivation of the Prices Using the Put–Call parity

We can derive the Black–Scholes prices for the calls and puts using the put–call parity given in Section 14.1.2. The basic idea is that if we are willing to assume that the prices are expectations with respect to a log-normal distribution, then the put–call parity implies that the log-normal distribution should be the risk-neutral log-normal distribution.

We assume that the distribution of the stock price c014-math-755 is defined by (14.49). Let us denote by c014-math-756 the price of the call option at time c014-math-757. We assume that the price of the call option is equal to

equation

and the value of the put option is equal to

equation

Thus, using (14.50) and (14.53),

equation

because c014-math-758 for all c014-math-759. The put–call parity (14.8) implies that we have to take

Inserting (14.72) to (14.50) and (14.53) leads to (14.58) and (14.59). This derivation was noted in Derman and Taleb (2005).

14.3.4 Examples of Pricing Using the Black–Scholes Model

We give some examples of Black–Scholes prices. The examples include pricing functions of options on a forward, caplets, swaptions, options on a foreign currency, and barrier options.

14.3.4.1 Options on a Forward

Let the underlying be a futures contract c014-math-761 with the maturity c014-math-762. Consider a call option with the expiration time c014-math-763. The payoff is c014-math-764, where c014-math-765 is the strike price. The price of the call option is obtained by replacing c014-math-766 in the Black–Scholes formula (14.58) by c014-math-767. This gives the price

where

equation

The volatility c014-math-769 is the volatility of the stock. This is called Black's formula, and it was introduced in Black (1976).

Why have we replaced c014-math-770 with c014-math-771? This is due to the fact that under the risk-neutral measure the stock price has the distribution

equation

The pricing formula of the futures contract in (14.2) gives c014-math-772. Thus, the distribution of c014-math-773 under the risk-neutral measure is

equation

The distribution of c014-math-774 is

equation

where c014-math-775. The price of the call option is

equation

and this expectation is calculated similarly as in (14.51).

The formula (14.73) holds when the option is subject to the stock type settlement. If the option is subject to the futures type settlement, then set c014-math-776 in (14.73). The futures type settlement means that the gains and losses are realized daily, whereas in the stock type settlement the gain or loss is realized at the time of liquidation.

14.3.4.2 Caplets

Caplets are discussed in Section 18.3.1. A caplet is a call option on the Libor rate c014-math-777. The payoff of a caplet at time c014-math-778 is

equation

where c014-math-779 is the principal, and c014-math-780 is the strike. We assume that under the risk-neutral measure

where c014-math-782 and c014-math-783 is the forward rate, defined in (18.14). Note that at time c014-math-784 the forward Libor rate is equal to the spot Libor rate: c014-math-785. The Black's formula for the price of the caplet is

14.75 equation

where

equation

The caplet price can be written as the expectation

equation

where the expectation is with respect to the risk-neutral distribution in (14.74). The expectation is calculated similarly as in (14.51). The caplet price c014-math-787 is obtained from the Black–Scholes price of a call option on a stock when c014-math-788 is replaced by c014-math-789 and c014-math-790 is replaced by c014-math-791.

Caps are defined in Section 18.3.2. Let c014-math-792 be the time points for the caplets on the Libor rates c014-math-793, where c014-math-794. A cap is priced by

equation

where c014-math-795 are the prices of the caplets.

14.3.4.3 Swaptions

Swaptions are discussed in Section 18.3.3. Let c014-math-796 be the current time, c014-math-797 be the expiry time, and c014-math-798 be the maturity time of the swaption. The payoff of an European call option on a swap is given in (18.27) as

equation

where c014-math-799 is the equilibrium swap rate, c014-math-800 is the strike,

equation

c014-math-801 is the principal, and c014-math-802 is a zero-coupon bond. We assume that under the risk-neutral measure

equation

where c014-math-803. The Black's formula for the price of the swaption is

equation

where

equation

The swaption price can be written as the expectation

equation

where the expectation is with respect to the risk-neutral distribution in (14.74). The expectation is calculated similarly as in (14.51). The caplet price c014-math-804 is obtained from the Black–Scholes price of a call option on a stock when c014-math-805 is replaced by c014-math-806 and c014-math-807 is replaced by c014-math-808.

Note that the simultaneous log-normality for all caplets and all swaptions is not consistent because a swap rate is a linear combination of forward rates and cannot be log-normal if the underlying forward rates are.

14.3.4.4 Options on a Foreign Currency

Let c014-math-809 be the price of a foreign currency in the domestic currency units. The payoff of an European call option on a foreign currency is given by

equation

where c014-math-810 is the strike, and c014-math-811 is the expiration time. According the Garman–Kohlhagen model, under the risk-neutral measure,

equation

where c014-math-812. Then the price of the call option is

equation

where

equation

where c014-math-813 is the risk-free rate in the foreign currency and c014-math-814 is the risk-free rate in the domestic currency. The put price is

equation

The price of the call option can be written as

equation

where the expectation is with respect to the risk-neutral measure. The expectation is calculated similarly as in (14.51).

The call price c014-math-815 on the exchange rate is obtained from the Black–Scholes price of a call option on a stock when c014-math-816 is replaced by the exchange rate c014-math-817 and c014-math-818 is replaced by c014-math-819.

14.3.4.5 Down-and-Out Call

A down-and-out call on stock c014-math-820 has the payoff

equation

where c014-math-821 is the strike price and c014-math-822 is the barrier. We assume that the stock price has a log-normal distribution

equation

under the risk-neutral measure, where c014-math-823. The price of the down-and-out call is

equation

where c014-math-824 is the Black–Scholes price of the vanilla call and

equation

where

equation

14.4 Black–Scholes Hedging

The hedging coefficients of calls and puts are given in (14.61). The Black–Scholes hedging coefficients are equal to the deltas of options. A delta is the derivative of the price of the option with respect to the price of the underlying: the call and put deltas are

equation

This is shown in Section 14.2.3; see (14.57).13 For the Black–Scholes pricing functions the call delta and the put delta are shown in (14.62) and (14.63) to be equal to

14.76 equation

and

14.77 equation

In this section, our purpose is to illustrate how hedging can be used to approximately replicate options, and to study how hedging frequency, expected stock returns, and the volatility of stock returns affect the replication. The study is made using Black–Scholes hedging, since Black–Scholes hedging provides a benchmark for comparing various hedging methods.

We take the purpose of hedging to be to make the probability distribution of the hedging error of the writer of the option as concentrated around zero as possible. We consider S&P 500 options and estimate the distribution of the hedging error of the writer using the S&P 500 daily data of Section 2.4.1.

Section 14.4.1 reviews historical simulation for the estimation of the distribution of the hedging error. Section 14.4.2 studies the effect of hedging frequency to the distribution of the terminal wealth. Section 14.4.3 studies the effect of the strike price, Section 14.4.4 studies the effect of the mean return, and Section 14.4.5 studies the effect of the return volatility.

14.4.1 Hedging Errors: Nonsequential Volatility Estimation

We have discussed the estimation of the hedging error using historical simulation in Section 13.3.1. Let us write the formulas for hedging error again, this time taking into account the varying hedging frequencies.

14.4.1.1 Hedging Errors

The hedging error c014-math-836 of the writer of the option is obtained from (13.10) by the formula

where

equation

when we take the risk-free rate c014-math-838. Here c014-math-839 is the price of the option, c014-math-840 is the terminal value of the option, c014-math-841 are the hedging coefficients, and c014-math-842 are the stock prices. In (14.78) the current time is denoted by c014-math-843, the time to expiration is c014-math-844 days, and hedging is done daily. The hedging can be done with a lesser frequency. When hedging is done c014-math-845 times during the period c014-math-846, then

14.79 equation

where

equation

14.4.1.2 Historical Simulation

Let us denote the time series of observed historical daily prices by c014-math-848. Let us denote the time to the expiration by

equation

(The interpretation is that the last observation c014-math-849 is made at time c014-math-850, and we are interested to write an option whose expiration time is c014-math-851.) We construct c014-math-852 sequences of prices:

equation

where

equation

for c014-math-853. Each sequence has length c014-math-854 and the initial prices are always c014-math-855. We estimate the distribution of the hedging error c014-math-856 from the observations

equation

where c014-math-857 is computed from the prices c014-math-858. For the estimation of the density we use both the histogram estimator and the kernel density estimator, defined in Section 3.2.2.

We consider hedging of call options. The Black–Scholes prices are given in (14.58) and the hedging coefficients are given in (14.62). The volatility c014-math-859 is estimated using the annualized sample standard deviation, computed from the complete data. We study the effect of volatility estimation to hedging in Section 14.5, where sequential (out-of-sample) volatility estimation is studied. In this section, we use in-sample volatility estimation.

More precisely, the c014-math-860th hedging error is

14.80 equation

where c014-math-862 is the Black–Scholes price from (14.58), computed with the stock price c014-math-863, time to expiration c014-math-864, and volatility c014-math-865. Here c014-math-866 is the annualized sample standard deviation computed from return data

equation

where

equation

c014-math-867. Thus, the volatility is estimated in-sample. The hedging coefficient c014-math-868 is the Black–Scholes delta from (14.62), computed with the stock price c014-math-869, time to expiration c014-math-870, and volatility c014-math-871. The terminal value is c014-math-872.

14.4.2 Hedging Frequency

In this section, we illustrate how a change in the hedging frequency changes the distribution of the hedging error. We study hedging of S&P 500 call options, using daily data of Section 2.4.1. The time to expiration is 20 days. The volatility in the Black–Scholes formula is the non-sequential annualized sample standard deviation.

14.4.2.1 Moneyness c014-math-873

Figure 14.13 shows time series of hedging errors. Panel (a) shows the case when there is no hedging, so that c014-math-874. Panel (b) shows the case when the hedging is done daily. Moneyness is c014-math-875.

Illustration of Hedging frequency: Time series of hedging errors.

Figure 14.13 Hedging frequency: Time series of hedging errors. (a) There is no hedging; (b) hedging is done daily.

Figure 14.14 shows tails plots of the hedging errors. Panel (a) shows the left tail plots and panel (b) shows the right tail plots. We show cases of no hedging (red), hedging once (black), hedging twice (blue), and hedging 20 times (green).

Illustration of Hedging frequency: Tail plots.

Figure 14.14 Hedging frequency: Tail plots. (a) Left tail plots and (b) right tail plots. We show cases of no hedging (red), hedging once (black), hedging twice (blue), and hedging 20 times (green).

Figure 14.15 shows (a) histograms of hedging errors and (b) kernel density estimates of the distribution of hedging errors. Panel (a) shows cases of no hedging (red) and daily hedging (green). Panel (b) shows additionally the cases of hedging once (black) and hedging twice (blue). Note that the red kernel density estimate is very inaccurate, because the underlying distribution is such that a large part of the probability mass is concentrated at one point.

Illustration of Hedging frequency: Density estimates of hedging errors.

Figure 14.15 Hedging frequency: Density estimates of hedging errors. (a) Histograms; (b) kernel density estimates. Panel (a) shows cases of no hedging (red) and hedging 20 times (green). Panel (b) shows additionally the cases of hedging once (black) and hedging twice (blue).

14.4.2.2 Moneyness c014-math-876

Figure 14.16 shows tails plots of the hedging errors. Panel (a) shows the left tail plots and panel (b) shows the right tail plots. We show cases of no hedging (red), hedging once (black), hedging twice (blue), and hedging 20 times (green).

Image described by caption and surrounding text.

Figure 14.16 Hedging frequency: Tail plots with moneyness c014-math-877. (a) Left tail plots; (b) right tail plots. We show cases of no hedging (red), hedging once (black), hedging twice (blue), and hedging 20 times (green).

Figure 14.17 shows (a) histograms of hedging errors and (b) kernel density estimates of the distribution of hedging errors when the moneyness is c014-math-878. Panel (a) shows cases of no hedging (red) and daily hedging (green). Panel (b) shows additionally the cases of hedging once (black) and hedging twice (blue). The red kernel density estimate is very inaccurate, because the underlying distribution is such that a large part of the probability mass is concentrated at one point.

Image described by caption and surrounding text.

Figure 14.17 Hedging frequency: Density estimates of hedging errors with moneyness c014-math-879. (a) Histograms; (b) kernel density estimates. Panel (a) shows cases of no hedging (red) and hedging 20 times (green). Panel (b) shows additionally the cases of hedging once (black) and hedging twice (blue).

14.4.2.3 Expected Utility

Figure 14.18 shows the estimated expected utility as a function of the hedging frequency. In panel (a) the moneyness is c014-math-880. In panel (b) c014-math-881. We apply the exponential utility function c014-math-882, where the risk-aversion parameter takes values c014-math-883 (black), c014-math-884 (red), and c014-math-885 (blue). The expected utilities are estimated using the sample averages. Increasing hedging frequency clearly increases the expected utility. We can see that when the risk aversion is small, then the increase in the expected utility as a function of hedging frequency is much smaller than the increase when the risk aversion is large.

Illustration of Expected utility as a function of hedging frequency.

Figure 14.18 Expected utility as a function of hedging frequency. (a) The moneyness is c014-math-886; (b) c014-math-887. The risk aversion parameter takes values c014-math-888 (black), c014-math-889 (red), and c014-math-890 (blue).

14.4.3 Hedging and Strike Price

In this section, we illustrate how a change in the strike price changes the distribution of the hedging error. We study hedging of S&P 500 call options, using daily data of Section 2.4.1. The time to expiration is 20 days and hedging is done daily. The volatility in the Black–Scholes formula is the nonsequential annualized sample standard deviation.

Figure 14.19(a) shows tail plots of hedging errors and panel (b) shows kernel density estimates of the distribution of the hedging error. The strikes prices of calls are c014-math-891 (red), c014-math-892 (black), and c014-math-893 (blue), when the stock price is c014-math-894. For in-the-money options the distributions have a larger spread than for out-of-the-money options. Also, the distributions for at-the-money options have a center that is located to the right from the centers for out-of-the-money options.

Illustration of Several strike prices. (a) Tail plots; (b) kernel density estimates.

Figure 14.19 Several strike prices. (a) Tail plots; (b) kernel density estimates. The strike prices are c014-math-895 (red), c014-math-896 (black), and c014-math-897 (blue), when the stock price is c014-math-898.

14.4.4 Hedging and Expected Return

We study the effect of the mean return to the distribution of the hedging error. This is done by manipulating the S&P 500 data. We change observations so that the new net returns are

equation

where c014-math-899 are the observed net returns, c014-math-900 is the sample mean of c014-math-901, and c014-math-902 is a value for the expected return that we choose. The new prices are obtained from the net returns by

equation

For the S&P 500 returns the annualized mean return is about c014-math-903. We try the annualized expected returns c014-math-904 and c014-math-905. Thus, c014-math-906 and c014-math-907.

Figure 14.20 shows histogram estimates of the distribution of the hedging error when the expected annualized return is c014-math-908. In panel (a) there is no hedging and in panel (b) there is daily hedging. That is, panel (a) shows a histogram made from realizations of the random variable c014-math-909, where c014-math-910, and c014-math-911 is the Black–Scholes price. Panel (b) shows a histogram made from realizations of the random variable c014-math-912. The time to the expiration of the call option is c014-math-913 days, and the strike price is c014-math-914 with the initial stock price c014-math-915. Since c014-math-916 is large the call option gives a profit to its owner with a large probability, as can be seen from panel (a). However, large c014-math-917 does not change much the distribution of the hedging error when hedging is done daily, as can be seen from panel (b). In fact, the corresponding distribution of the hedging error when the expected return is moderate is shown in Figure 14.15.

Image described by caption and surrounding text.

Figure 14.20 No hedging versus daily hedging with a large positive drift. (a) No hedging: A histogram from realizations of c014-math-918. (b) Daily hedging: A histogram from realizations of c014-math-919.

Figure 14.21 shows the setting of Figure 14.20 when the annualized expected return is c014-math-920, instead of c014-math-921. Panel (a) shows the distribution of the hedging error of the writer when no hedging is done and panel (b) shows the hedging error when delta hedging is done daily. Since the expected return is negative, the writer of the call option gets a profit with a large probability, but this does not affect much the hedging error when the hedging is done daily, as can be seen from panel (b).

Illustration of No hedging versus daily hedging with a negative drift.

Figure 14.21 No hedging versus daily hedging with a negative drift. (a) No hedging: A histogram from realizations of c014-math-922. (b) Daily hedging: A histogram from realizations of c014-math-923.

(1) When the drift is larger than the risk-free rate, then the Black–Scholes price c014-math-924 is smaller than the expectation c014-math-925. The possibility of hedging makes c014-math-926 smaller than c014-math-927. The expectation c014-math-928 increases when the drift increases but the possibility of hedging makes the price independent of the drift. (2) When the drift is equal to the risk-free rate, then c014-math-929 is close to c014-math-930, but hedging reduces the risk of the writer of the option, because it changes the wealth distribution of the writer of the option. (3) When the drift is negative, then c014-math-931 is smaller than c014-math-932, and hedging reduces the expected profit of the writer of the option. However, the hedging reduces also the risk of the writer of the option, and thus hedging is reasonable even in the case of negative drift.

14.4.5 Hedging and Volatility

We study the effect of the return volatility to the distribution of the hedging error. We manipulate the S&P 500 data by changing observations so that the new net returns are

equation

where c014-math-933 are the observed net returns, c014-math-934 is the sample standard deviation of c014-math-935, c014-math-936 is the sample mean of c014-math-937, and c014-math-938 is a value of our choice for the volatility. The new prices are obtained from the net returns by

equation

For the S&P 500 returns the annualized sample standard deviation is about c014-math-939. We try the annualized standard deviation c014-math-940. Thus, c014-math-941.

Figure 14.22 shows histogram estimates of the distribution of the hedging error when the annualized volatility is c014-math-942. In panel (a) there is no hedging, and in panel (b) there is daily hedging. That is, panel (a) shows a histogram made from realizations of the random variable c014-math-943, where c014-math-944, and c014-math-945 is the Black–Scholes price. Panel (b) shows a histogram made from realizations of the random variable c014-math-946. The time to the expiration is c014-math-947 days, and the strike price is c014-math-948 with the initial stock price c014-math-949. We see that the larger volatility makes the dispersion of the probability distribution of the hedging error larger.

Image described by caption and surrounding text.

Figure 14.22 Large volatility. (a) No hedging: A histogram of realizations of c014-math-950. (b) Daily hedging: A histogram of realizations of c014-math-951.

14.5 Black–Scholes Hedging and Volatility Estimation

We continue to study the distribution of the hedging error c014-math-952, as in Section 14.4. In this section, our aim is to study how the volatility estimation affects the distribution of the hedging error. The Black–Scholes prices and the hedging coefficients depend on the annualized volatility c014-math-953. We compare the performance of GARCH(c014-math-954) and exponentially weighted moving averages for the estimation of c014-math-955. The performance of Black–Scholes hedging will be used as a benchmark.

14.5.1 Hedging Errors: Sequential Volatility Estimation

We have discussed in Section 13.3.1 the estimation of the distribution of the hedging error using historical simulation. We write again the formulas for the hedging error and historical simulation, adapting to the current setting.

14.5.1.1 Hedging Errors

The hedging error c014-math-956 of the writer of the option is obtained from (13.10) as

equation

where

equation

Here the risk-free rate is c014-math-957, c014-math-958 is the price of the option, c014-math-959 is the terminal value of the option, c014-math-960 are the hedging coefficients, c014-math-961 are the stock prices, the current time is denoted by 0, the time to expiration is c014-math-962 days, and hedging is done daily.

14.5.1.2 Historical Simulation

We denote the time series of observed historical daily prices by c014-math-963. We construct c014-math-964 sequences of prices:

equation

where

equation

for c014-math-965. Each sequence has length c014-math-966 and the initial price in each sequence is c014-math-967. We estimate the distribution of the hedging error c014-math-968 from the observations

equation

where c014-math-969 is computed from the prices c014-math-970.

More precisely, the c014-math-971th hedging error is

where c014-math-973 is the Black–Scholes price from (14.58), computed with the stock price c014-math-974, time to expiration c014-math-975, and volatility c014-math-976. Here c014-math-977 is the annualized volatility estimate computed using return data

equation

where

equation

Thus, the volatility estimation is done sequentially (out-of-sample). The hedging coefficient c014-math-978 is the Black–Scholes delta from (14.62), computed with the stock price c014-math-979, time to expiration c014-math-980, and volatility c014-math-981. The terminal value is c014-math-982.

For the estimation of the density we use both the histogram estimator and the kernel density estimator, defined in Section 3.2.2.

14.5.2 Distribution of Hedging Errors

In the following examples the time to expiration is c014-math-983 trading days. We start pricing and hedging after 4 years of data has been collected. The risk-free rate is equal to zero. We consider S&P 500 call options and use the daily S&P 500 data, described in Section 2.4.1.

As a summary of the results, we can note that the GARCH(c014-math-984) and the exponential moving average (with a suitable smoothing parameter) improve the distribution of the hedging error from the point of view of the writer of the option, when compared to the sequential sample standard deviation. GARCH(c014-math-985) and the exponential moving average lead to a distribution whose left tail is lighter: with these volatility estimators the losses of the writer of the option are smaller. On the other hand, GARCH(c014-math-986) and the exponential moving average lead to larger positive hedging errors. The positive hedging errors are gains for the writer of the option.

Figure 14.23 shows (a) the means of negative hedging errors and (b) the means of positive hedging errors as a function of the moneyness c014-math-987. The arithmetic means of positive and negative hedging errors are defined as

equation

where the hedging errors c014-math-988 are defined in (14.81). The hedging is done with sequentially computed sample standard deviation (red with “s”), GARCH(c014-math-989) volatility (blue with “g”), the exponential moving average with c014-math-990 (yellow with “1”), and with smoothing parameter c014-math-991 (green with “2”). We see that the sample standard deviation is the best for the positive hedging errors and the worst for the negative hedging errors. The performance of GARCH(c014-math-992) and the exponentially weighted moving average with c014-math-993 are close to each other. The exponentially weighted moving average with c014-math-994 has the performance between the sample standard deviation and the GARCH(c014-math-995). We can see that hedging errors are larger for the at-the-money options than for the out-of-the-money options.

Illustration of Means of hedging errors.

Figure 14.23 Means of hedging errors. The figure shows the means of (a) negative hedging errors as a function of moneyness and (b) positive hedging errors. The volatility is estimated by the sequentially computed sample standard deviation (red with “s”), GARCH(1,1) (blue with “g”), the exponential moving average with c014-math-996 (yellow with “1”), and with c014-math-997 (green with “2”).

Illustration of Distribution of hedging errors.

Figure 14.24 Distribution of hedging errors. The figure shows (a) tail plots of hedging errors and (b) kernel density estimates of the distribution of the hedging error. The volatility is estimated by the sequentially computed sample standard deviation (red), GARCH(c014-math-998) (blue), and the exponential moving average with c014-math-999 (green).

Illustration of Distribution of hedging errors:Moving averages.

Figure 14.25 Distribution of hedging errors: Moving averages. The figure shows (a) tail plots of hedging errors and (b) kernel density estimates of the distribution of the hedging error. The volatility is estimated by the sequentially computed standard deviation (red), and by the exponentially weighted moving average with c014-math-1000 (purple), c014-math-1001 (green), c014-math-1002 (blue), and c014-math-1003 (black).

Figure 14.24 shows (a) tail plots of hedging errors and (b) kernel density estimates of the distribution of the hedging error. The moneyness of call options is c014-math-1004. The volatility is estimated by the sequentially computed standard deviation (red), by GARCH(c014-math-1005) (blue), and by the exponentially weighted moving average with c014-math-1006 (green). 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 is chosen by the normal reference rule.

Figure 14.25 shows (a) tail plots of hedging errors and (b) kernel density estimates of the hedging error. The volatility is estimated by the sequentially computed standard deviation (red), and by the exponentially weighted moving average with c014-math-1007 (purple), c014-math-1008 (green), c014-math-1009 (blue), and c014-math-1010 (black).

equation

since c014-math-145 for the risk-neutral measure c014-math-146.

equation
equation
equation

where c014-math-827 is the number of stocks that are bought at time c014-math-828 to hedge the position until c014-math-829. The amount c014-math-830 is borrowed with the risk-free rate at time c014-math-831. If the hedging gives a position without risk, we should have c014-math-832, which gives

equation

when c014-math-833. This gives the instantaneous optimal hedging coefficient as

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

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