Chapter 13
Principles of Asset Pricing

Asset pricing can be studied in two different settings: absolute pricing and relative pricing. Absolute pricing tries to explain the prices in terms of fundamental macroeconomic variables, applying utility functions and preferences. Relative pricing tries to explain the prices of a group of assets given the prices of a more fundamental group of assets.

We concentrate on relative pricing. Derivatives are assets whose payoffs are defined in terms of the payoffs of some basis assets. For example, an European call option gives the right to buy the underlying asset at the given expiration time c013-math-001 at the given strike price c013-math-002. Thus, the payoff of the call option at time c013-math-003 is equal to

equation

where c013-math-004 is the value of the underlying asset. We want to find a “fair price” c013-math-005 for the call option, when c013-math-006 is a previous time.

Derivatives are traded in exchanges just like stocks, and the price of a derivative is determined in an exchange by supply and demand. It can be argued that the pricing of the market is typically efficient. However, it is of interest to try to find fair prices by statistical and probabilistic methods at least for the following two reasons. (1) Sometimes options are bought and sold over the counter and not in exchanges. In this case, there is no information provided by the markets. (2) It is possible that the market prices are irrational. This can certainly happen in illiquid markets. In this case, a market participant can profit from the knowledge of scientific methods of pricing.

Besides pricing of options, it is of equal importance to hedge options. In fact, our main emphasis will be on the pricing by quadratic hedging. In this approach, the price of an option will be the initial investment of a trading strategy, which minimizes the quadratic error

equation

where c013-math-007 is the wealth obtained by the hedging strategy, and c013-math-008 is the value of the derivative at the expiration. This approach will be developed more in detail in Chapter 16.

Section 13.1 studies general principles of pricing heuristically, discussing such concepts as absolute and relative pricing, arbitrage, the law of one price, and completeness of models. In addition, we introduce the idea of quadratic hedging.

Section 13.2 presents basics of mathematical asset pricing in discrete time. We describe the first and the second fundamental theorems of asset pricing (Theorems 13.1 and 13.3). The first fundamental theorem says that a market is arbitrage-free if and only if there exists an equivalent martingale measure. The second fundamental theorem says that every derivative can be replicated if and only if the martingale measure is unique. If every derivative can be replicated, then it is said that the market is complete. Theorem 13.2 states that the arbitrage-free prices of European options are expectations with respect to an equivalent martingale measure.

We give a proof of the first fundamental theorem of asset pricing. The proof is constructive: we construct an equivalent martingale measure for an arbitrage-free market. The most proofs of the first fundamental theorem of asset pricing found in the literature are not constructive, but apply abstract functional analysis. However, the construction of suitable equivalent martingale measures is useful for practical applications, because these measures lead to the collections of arbitrage-free prices. We do not give a proof of the second fundamental theorem of asset pricing. This is due to the fact that the general theory of complete markets seems to be less relevant from the practical point of view than the theory of incomplete markets, although the Black–Scholes model is useful in applications. Chapter 14 describes the theory of Black–Scholes pricing and Chapter 15 is devoted to incomplete models.

Section 13.3 discusses methods for the comparison of different pricing and hedging methods. The main method for the comparison is to use historical simulation to generate trajectories of prices, hedge the derivative through the trajectories, and then compute the sample mean of the squared hedging errors.

13.1 Introduction to Asset Pricing

Section 13.1.1 discusses absolute pricing with the help of coin tossing games. These examples show that utility functions can be useful in determining reasonable prices. Section 13.1.2 discusses how the principle of excluding arbitrage and the law of one price can be applied in relative pricing. The one period binary model is introduced. This model will be used to derive the Black–Scholes prices in Chapter 14. Section 13.1.3 discusses relative pricing in cases where arbitrage cannot be applied. The one period ternary model is an example of such case. In these cases, a fair price can be defined by minimizing the mean squared hedging error, for example.

13.1.1 Absolute Pricing

Let us consider a coin tossing game where a participant receives 1 € when heads occur and 0 € when tails occurs. The probability of getting heads is 1/2 and the probability of obtaining tails is 1/2. What is the fair price to participate in this game? It can be argued that the fair price is the expected gain:

equation

The fairness of the price can be justified by the law of large numbers. The law of large numbers implies that the gain from repeated independent repetitions of the game with price 0.5 € converges to zero with probability 1. A larger price than 0.5 € would give an almost sure profit to the organizer of the game in the long run and a smaller price than 0.5 € would give an almost sure profit to the player of the game in the long run.

It does not seem as clear what the price should be if we change the game so that a participant receives 1 million € when heads occur and 0 € when tails occur. Only few people would be willing to invest half a million € in order to participate in this game. The law of large numbers cannot be applied to justify a price because the probability of a bankruptcy is quite large when a player repeats the game.1

It can be argued that the price of the game should be equal to the expected utility: Let c013-math-009 be the random variable with c013-math-010 and c013-math-011, where c013-math-012 million €. Then the expected utility is c013-math-013, where c013-math-014 is a utility function.

The St. Petersburg paradox can be used to suggest that a utility function should be used. In the St. Petersburg paradox, the banker flips the coin until the heads come out the first time. The player receives c013-math-015 coins when there are c013-math-016 tosses of the coin (1 coin if the heads come out in the first toss, 2 coins if the heads come out in the second toss, 4 coins if the heads come out in the third toss, and so on). What is the fair entrance fee to the game? We can calculate the expected gain. The probability that there are c013-math-017 tosses is c013-math-018. Thus the expected payoff is

equation

Thus, it would seem that the entrance fee could be arbitrarily high. However, applying common sense, it does not seem reasonable to pay a high entrance fee. The paradox can be solved by using a utility function to measure the utility of the wealth. For example, the logarithmic utility function c013-math-019 gives the expected utility of the game

equation

which would give the price of two coins for the game.2

The St Petersburg paradox suggests that we could use the expected utility instead of the expected monetary payoff to determine fair prices. Utility maximization will be discussed in Section 15.2.5, as a method for derivative pricing, but otherwise we do not study further this approach.

13.1.2 Relative Pricing Using Arbitrage

Sometimes relative pricing can be done solely by applying the principle of excluding arbitrage. We illustrate this type of relative pricing using a coin tossing example.3 After that, arbitrage and the law of one price are discussed more generally.

13.1.2.1 Pricing in a One Period Binary Model

Let us consider two games related to the same tossing of a coin. The first game is such that the player receives c013-math-023 € when heads occur and c013-math-024 € when tails occurs, where c013-math-025. The participation to this game can be compared to buying a stock. We denote with c013-math-026 the random variable with c013-math-027 and c013-math-028.

The second game is such that the player receives 1 € when heads occur and 0 € when tails occurs. The participation to this game can be compared to buying a derivative. Indeed, the second game can be considered as a derivative because the payoff in the second game is random variable c013-math-029 for c013-math-030, where c013-math-031 and c013-math-032. Random variable c013-math-033 has the distribution c013-math-034 and c013-math-035. The third asset is a bond with value c013-math-036. The price of bond is 1 and the price of stock is denoted with c013-math-037. We want to find the price of the derivative.

The derivative can be replicated with the bond and the stock: Consider the portfolio with c013-math-038 bonds and c013-math-039 stocks. We choose

equation

The replicating portfolio is c013-math-040. Indeed, we have that c013-math-041, because

equation

By the law of one price, to exclude the possibility of arbitrage, the price of the derivative has to be equal to the price of the portfolio:4

equation

The price of the portfolio is

equation

Thus, the price of the derivative is

The price of the derivative is in general not equal to 0.5. If c013-math-049, then the price of the derivative is c013-math-050. If c013-math-051, then the price of the derivative satisfies c013-math-052.5

We have given the price of the derivative in (13.1) in terms of the price of the stock. This is an example of relative pricing: a price of an asset is expressed in terms of the prices of another asset.

13.1.2.2 Arbitrage

Arbitrage is both a term of everyday language and a technical term used in mathematical finance.

Arbitrage is used in everyday language to denote a financial operation where one obtains a profit with probability one by a simultaneous selling and buying of assets. We give two examples of this type of arbitrage.

  1. 1. The stock of Daimler is listed both in Frankfurt and Stuttgart stock exchanges. If the stock can be bought in Frankfurt with the price of 10 € and sold in Stuttgart with the price of 11 €, we obtain a risk free profit of 1 € (minus the transaction costs).
  2. 2. Suppose the price of a stock is 10 € and a call option with strike price c013-math-053 € with the expiration time in 1 week can be bought with the price of 1 €. Then, we can sell the stock short and buy the call option. The profit of the operation will be c013-math-054 € (buying the call costs 1 €, selling the stock short gives 10 €, and exercising the option costs 8 €).

    In general, we have a lower bound c013-math-055 for the price of a call option, where c013-math-056 is the price of the stock at the time of buying the option, and c013-math-057 is the strike price. See (14.9) for a more precise lower bound.

In mathematical finance, an arbitrage is a financial operation whose payoff is always nonnegative and sometimes positive, that is, the probability of a nonnegative payoff is one and the probability of a positive payoff is greater than zero. More formally, arbitrage portfolio c013-math-058 is such that its value at time c013-math-059 satisfies c013-math-060 but its value c013-math-061 at a later time c013-math-062 satisfies c013-math-063 and c013-math-064. A reasonable system of prices should be such that arbitrage is excluded, so that there does not exist an arbitrage portfolio.

The absence of arbitrage implies the law of one price.

13.1.2.3 The Law of One Price and the Monotonicity Theorem

The law of one price states that if two financial assets have the same payoffs then they have the same price: If two portfolios satisfy

equation

then their prices are equal at a previous time c013-math-065:

equation

The absence of arbitrage implies that the law of one price holds. Indeed, consider the case where the law of one price does not hold. Then we have two assets with different prices at time c013-math-066, say c013-math-067, and the prices of the assets are the same with probability 1 at a later time c013-math-068: c013-math-069. Then we can by the cheaper asset at time c013-math-070 and sell the more expensive asset at time c013-math-071 to obtain the amount c013-math-072. This amount can be put into a bank account. At time c013-math-073 the two assets have the same price, and thus we have locked the profit of time c013-math-074. We have shown that there exists an arbitrage opportunity. Thus we have shown that the absence of arbitrage implies that the law of one price holds.

The monotonicity theorem states that if two financial assets satisfy

equation

then their prices satisfy

equation

at time c013-math-075. Furthermore, if c013-math-076, then their prices satisfy c013-math-077 at time c013-math-078. This formulation of the monotonicity theorem is similar to the formulation of Blyth (2014, p. 48).

The law of one price implies the linearity of the pricing function. Let

equation

be a portfolio, and let c013-math-079 be the prices of the basis assets at time c013-math-080. Then the price of the portfolio at time c013-math-081 is

13.1.2.4 Pricing using The Law of One Price

The law of one price can be used to price linear assets by replication.6 Furthermore, the law of one price can be used to price all assets in complete markets. By a market we mean a collections of tradable assets together with assumptions about the probability distributions of the asset values. A complete market is such that any possible payoff can be obtained by a portfolio of assets. That is, assume that the market has tradable assets c013-math-083. Assume that an arbitrary payoff c013-math-084 can be obtained, so that c013-math-085. The law of one price implies that price of this payoff is

equation

where we applied the linearity in (13.2).

Futures are linear derivatives, and thus the law of one price can be used to price futures; see Section 14.1. Futures can be priced by the law of one price because futures can be defined as a portfolio of the underlying asset and a bond: the payoff of a futures contract is a linear combination of the payoffs of the underlying asset and a bond.

The payoff of an option is not a linear function of the payoff of the underlying. Thus options cannot be priced as easily as futures. The law of one price can be used to price options in the Black–Scholes model, because the Black–Scholes model is a complete model for the markets, so that all derivatives can be replicated (linearly).

The law of one price can be used to derive the put-call parity, which says that the prices of two options satisfy an equation. The law of one price can also be used to give bounds to option prices without assuming the Black–Scholes model, or any other market model. See Section 14.1.2 for the derivation of the put-call parity.

13.1.3 Relative Pricing Using Statistical Arbitrage

We have derived the price of the derivative in (13.1) using the replication of the derivative with a stock and a bond. The exact replication is possible only under special circumstances. It suffices to move from the binary model to a ternary model to make exact replication impossible, so that only approximate replication is possible. We use the term “statistical arbitrage” to mean quadratic hedging (variance optimal hedging), quantile hedging, and other similar methods for approximate replication.

13.1.3.1 Pricing in a One Period Ternary Model

Let us have two games related to the same tossing of a dice. The first game is such that the player receives c013-math-086 € when the dice shows 1 or 2, c013-math-087 € when the dice shows 3 or 4, and c013-math-088 € when the dice shows 5 or 6, where c013-math-089. The participation to this game is an analogy to buying a stock and we denote with c013-math-090 the random variable with c013-math-091.

The second game is such that the player receives 0 € when the dice shows 1, 2, 3, or 4 and 1 € when the dice shows 5 or 6. The participation to this game is an analogy to buying a derivative and we denote with c013-math-092 the random variable c013-math-093, where c013-math-094 is defined by c013-math-095 when c013-math-096 and c013-math-097 when c013-math-098. Now c013-math-099 and c013-math-100. The third asset is a bond with value c013-math-101. The price of the bond is 1 and the price of the stock is denoted with c013-math-102. We want to find the price c013-math-103 of the derivative.

The derivative cannot be replicated with the bond and the stock: Consider the portfolio with c013-math-104 bonds and c013-math-105 stocks. The portfolio is c013-math-106. We have c013-math-107 when c013-math-108 and c013-math-109 satisfy

equation

We can typically not find such c013-math-110 and c013-math-111 because, in general, two parameters cannot satisfy three equations simultaneously. To obtain an approximate replication we could choose c013-math-112 and c013-math-113 so that c013-math-114 is minimized. We have that

equation

Since the probabilities are all equal to c013-math-115, we get the least squares solution for c013-math-116:

equation

where

equation

The solution is

equation

We set the price of the derivative to be equal to the price of the approximately replicating portfolio:

equation

If c013-math-117, then c013-math-118.

13.1.3.2 Statistical Arbitrage and the Law of Approximate Price

Statistical arbitrage is a financial operation where a profit is obtained with a high probability. The principle of excluding statistical arbitrage is a pricing principle, which can be used when the principle of excluding arbitrage does not apply. However, the concept of statistical arbitrage can be defined in many ways. Let us compare the principle of excluding arbitrage to the idea of excluding statistical arbitrage.

  1. 1. Excluding arbitrage. The value of a derivative is c013-math-119 at time c013-math-120. Let us have another asset whose value is c013-math-121 at time c013-math-122. Assume that the values are equal with probability 1: c013-math-123. Then it should hold that the value of the derivative and the other asset are equal at all previous times: c013-math-124 for all previous times c013-math-125. Otherwise, there would be an arbitrage opportunity: sell the more expensive instrument and buy the cheaper instrument to obtain a risk free profit at time c013-math-126.
  2. 2. Excluding statistical arbitrage. The value of a derivative is c013-math-127 at time c013-math-128. Let us have an other asset whose value is c013-math-129 at time c013-math-130. If the random variables c013-math-131 and c013-math-132 are “close,” then the prices c013-math-133 and c013-math-134 should be close at all previous times c013-math-135. The closeness of random variables can be defined in many ways. For example, we can say that two random variables c013-math-136 and c013-math-137 are close when c013-math-138 is small. A derivative can be priced with statistical arbitrage if we can construct an asset, which replicates the payoff of the derivative with high probability.

Pricing with statistical arbitrage requires that we define the best approximation c013-math-139 to a random payoff c013-math-140. As an example, we can consider a call option written at time c013-math-141, with the strike price c013-math-142 and with the expiration time c013-math-143. The payout of the option at the expiration time is c013-math-144, where c013-math-145 is the price of the underlying instrument at time c013-math-146. The best constant approximation of random variable c013-math-147 in the sense of mean squared error is its expectation:7

equation

where the minimization is taken with respect to all real numbers. Thus, expectation c013-math-149 can give a first approximation to the price of c013-math-150. We can use the underlying asset to provide a better approximation. The best approximation of c013-math-151 with a function c013-math-152 of c013-math-153 is the conditional expectation:

equation

where the minimization is taken with respect to functions c013-math-154, and function c013-math-155 takes values c013-math-156. Thus, the conditional expectation c013-math-157 could be a candidate for the fair price. However, the conditional expectation is typically not a tradable asset, and we will make a further restriction to find such function c013-math-158, which is tradable, which leads to linear approximations.

13.2 Fundamental Theorems of Asset Pricing

Our intention is to describe the basic mathematical terminology and fundamental theorems of asset pricing in discrete time models. Our presentation follows Shiryaev (1999) and Föllmer and Schied (2002). The mathematics of asset pricing is a fascinating topic with elegant results and we hope that the presentation will inspire readers to study the subject in a greater detail.

The first fundamental theorem of asset pricing says that a market is arbitrage-free if and only if there exists an equivalent martingale measure. Furthermore, these martingale measures define the collection of arbitrage-free prices for a derivative. In a complete model, there is exactly one equivalent martingale measure, but in an incomplete model there are many equivalent martingale measures. Thus, the main problem will be to choose the martingale measure for pricing from a collection of available martingale measures. Our emphasis will be in incomplete models.

13.2.1 Discrete Time Markets

Let c013-math-159 be the time series of prices of a riskless bond (bank account). Let c013-math-160 be the vector time series of prices of risky assets, where c013-math-161. The price vector that contains both the bond and the risky assets is denoted by

equation

13.2.1.1 Filtered Probability Spaces

The underlying probability space c013-math-162 is accompanied with a filtration of sigma-algebras c013-math-163.8 The price process of stocks is adapted with respect to the filtration: c013-math-171 is measurable with respect to c013-math-172.9 The price process of the bond is predictable with respect to the filtration: c013-math-177 is measurable with respect to c013-math-178, c013-math-179, and c013-math-180 is measurable with respect to c013-math-181. Thus, the value of c013-math-182 is known at time c013-math-183, which makes c013-math-184 locally riskless. The prices are assumed to be nonnegative. We assume that10

equation

Thus, c013-math-187 and elements of c013-math-188 are constants (with probability 1).

13.2.1.2 Trading Strategies

A trading strategy is

equation

The values c013-math-189 and c013-math-190 express the quantity of the bond and the c013-math-191th asset held between c013-math-192 and c013-math-193. The trading strategy is predictable: c013-math-194 and c013-math-195 are measurable with respect to c013-math-196. This means that c013-math-197 and c013-math-198 are determined at time c013-math-199, using the information available at time c013-math-200.

13.2.1.3 Examples

Let us give examples of the locally riskless bond c013-math-201. We can take c013-math-202, where c013-math-203 is a constant, or c013-math-204, where c013-math-205. In addition, we can take c013-math-206, and

equation

for c013-math-207, where c013-math-208 is predictable. We can also take c013-math-209 and c013-math-210, where c013-math-211 is predictable.

Consider the two period binary model as an example of adaptability and predictability. Now c013-math-212 and c013-math-213. The initial stock price is c013-math-214. The next price is c013-math-215 and the final price is c013-math-216, where c013-math-217 and c013-math-218 are random variables. Random variable c013-math-219 satisfies c013-math-220 and c013-math-221 for c013-math-222, where c013-math-223 is a fixed constant. Random variable c013-math-224 has the same distribution as c013-math-225, and is independent of c013-math-226. Choose

equation

where c013-math-227 and c013-math-228 refer to the upwards and downwards movements of the stock. Set c013-math-229 describes all possible trajectories of the process. Now,

equation

Let c013-math-230. It follows that

equation

In order for the stock prices to be adapted to the filtration we need

equation

We have that

equation

It follows that in order for the stock prices to be adapted to the filtration we need

equation

Let the initial bond price be c013-math-231. Let the next bond price be c013-math-232, where c013-math-233 is a constant. Let the final bond price be c013-math-234, where c013-math-235. Now bond prices are predictable with respect to the filtration. Bond price at time c013-math-236 depends only on the stock price at time c013-math-237. Thus, the bond price at time c013-math-238 is a random variable which is known at time c013-math-239.

13.2.2 Wealth and Value Processes

We define the wealth and value processes. The value process is obtained from the wealth process by dividing with the bond price. After that we derive an expression for the wealth and value processes under the assumption of self-financing.

13.2.2.1 The Wealth and Value Processes

We use the following notation for the inner product:

equation

At time 0 the initial wealth is c013-math-240 and after that,

equation

Indeed, the portfolio vector c013-math-241 is chosen at time c013-math-242 and hold during the period c013-math-243.

We assume that c013-math-244 for all c013-math-245 and choose the bond as a numéraire. The discounted price process is defined by

equation

We denote

The value process is defined as c013-math-247 and

equation

13.2.2.2 The Wealth Process under Self-financing

We assume in most cases that the trading strategy is self-financing. The local quadratic hedging without self-financing in Section 16.2.3 is a case where self-financing is not assumed.

Let us describe trading under the condition of self-financing. At time 0 the initial wealth is c013-math-248. The wealth is allocated among the available assets: the quantities c013-math-249 are chosen so that

equation

The prices change from c013-math-250 to c013-math-251, and the wealth changes accordingly from c013-math-252 to c013-math-253. After that, wealth c013-math-254 is allocated among the available assets. We obtain

equation

We continue in this way to obtain

equation

The final wealth is

equation

At time c013-math-255 we need not do the reallocation, because it is the last time instance.

We have described a process of trading, which is self-financing. We say that a trading strategy c013-math-256 is self-financing if

When the trading strategy is self-financing, then no external funds are received, and no funds are reserved for consumption.

Under the assumption (13.4) of self-financing, the change of wealth can be written as

equation

for c013-math-258. Thus, the wealth at time c013-math-259 can be written as11

13.5 equation

where c013-math-263.

13.2.2.3 The Value Process Under Self-Financing

Let us assume that the rebalancing is made respecting the condition of self-financing: c013-math-264 is chosen so that the wealth c013-math-265 is allocated among the assets. Equation c013-math-266 sets a linear constraint on the vector c013-math-267. It is convenient to write the wealth process so that the quantity c013-math-268 of bonds is eliminated, and this can be done using the value process.

The self-financing condition in (13.4) implies that the discounted price process c013-math-269 in (13.3) satisfies

Similarly as for the wealth process, it holds that

Under the condition (13.6) of self-financing, an increment of the value process can be written as

equation

where the last equality follows because the first element of c013-math-272 is 1 for all c013-math-273. Thus, the value at time c013-math-274 can be written as

13.8 equation

Note that the value process is written in terms of the quantity c013-math-276 of stocks. The quantity c013-math-277 of bonds is obtained from the equations

13.9 equation

which follow from self-financing equations (13.7).

The gains process is defined as

equation

For a self-financing strategy

The gains process is a discrete stochastic integral. The gains process is a transformation of c013-math-280 by means of c013-math-281.12

The value process can be used to derive some expressions for the wealth. For example, when c013-math-291, then13

equation

13.2.3 Arbitrage and Martingale Measures

An arbitrage opportunity is a self-financing trading strategy c013-math-294 so that its value process satisfies

equation

This means that with an initial investment of zero it is possible to get a final wealth, which is always nonnegative and sometimes positive.

A martingale is a stochastic process c013-math-295 on a filtered probability space c013-math-296 if14

  1. 1. c013-math-301 it is adapted (c013-math-302 is c013-math-303 measurable),
  2. 2. c013-math-304 for c013-math-305,
  3. 3. c013-math-306 for c013-math-307.

A martingale difference satisfies conditions 1 and 2, but condition 3 takes the form c013-math-308 for c013-math-309. Thus, a martingale difference is a martingale if c013-math-310.

A probability measure c013-math-311 on c013-math-312 is called a martingale measure, or a risk neutral measure, if the discounted price process c013-math-313 is a c013-math-314-dimensional martingale. Then,

equation

for c013-math-315, and

equation

c013-math-316-almost surely for c013-math-317, where c013-math-318.

An equivalent martingale measure is a martingale measure, which is equivalent to the original measure c013-math-319. Measures c013-math-320 and c013-math-321 are equivalent, if c013-math-322 if and only if c013-math-323. The equivalence of measures is denoted by c013-math-324. Let c013-math-325 be the set of equivalent martingale measures:

equation

where c013-math-326 is the underlying probability measure of the market model.

The first fundamental theorem of asset pricing states that a market model is arbitrage-free if and only if there exists an equivalent martingale measure.

Theorem 13.1 was proved in Harrison and Kreps (1979) and Harrison and Kreps (1981) in the case of finite c013-math-328. Dalang et al. (1990) proved it for arbitrary c013-math-329. A proof of Theorem 13.1 can be found in Föllmer and Schied (2002, Theorem 5.17) and in Shiryaev (1999, p. 413). We proof Theorem 13.1 by first showing that the existence of an equivalent martingale measure implies no-arbitrage. After that, an equivalent martingale measure is constructed for an arbitrage-free model.

13.2.3.1 The Existence of a Risk Neutral Measure Implies No-Arbitrage

We think that it is instructive to prove the result first for the case c013-math-330, and after that for the general case c013-math-331.

A Proof in the One Period Model

Assume that there exists an equivalent martingale measure c013-math-332. The martingale measure satisfies

equation

for c013-math-333. The value process is

equation

Take a portfolio such that c013-math-334. Then,

equation

Thus, we cannot have c013-math-335. Thus, we cannot have c013-math-336, and we cannot have c013-math-337, and c013-math-338 cannot be an arbitrage opportunity.

A Proof in the Multiperiod Model

A proof can be found in Shiryaev (1999, p. 417). We assume that there exists a martingale measure c013-math-339, which is equivalent to c013-math-340 and such that c013-math-341 is a c013-math-342-dimensional martingale with respect to c013-math-343, where c013-math-344. We noted in (13.10) that the value process satisfies

equation

where

equation

Note that since c013-math-345 is a martingale with respect to c013-math-346, then sequence c013-math-347 is a martingale transformation with respect to c013-math-348, when a martingale transformation is defined by (13.11).

Let c013-math-349 be a strategy with c013-math-350, and c013-math-351, so that c013-math-352, and c013-math-353. Let us assume that c013-math-354 for c013-math-355, where c013-math-356 is a constant. Then c013-math-357 is a martingale, and c013-math-358. Since c013-math-359, then c013-math-360, which implies c013-math-361, and c013-math-362.

The case of unbounded c013-math-363 is handled in Shiryaev (1999, p. 98, Chapter II §1c)15 and in Föllmer and Schied (2002, Theorem 5.15, p. 229).

13.2.3.2 A Construction of an Equivalent Martingale Measure

We have taken the proof from Shiryaev (1999, p. 413), which follows the ideas of Rogers (1994). The construction of equivalent martingale measures is based on the Esscher conditional transformations. The Esscher transforms were used also in Gerber and Shiu (1994) to construct an equivalent martingale measure. Note that the most proofs found in literature are not constructive, but apply a separation theorem in finite-dimensional Euclidean spaces, for example.16 It is instructive to first consider the case of the one period model with one risky asset, second consider the case of the multiperiod model with one risky asset, and third consider the general case.

A Martingale Measure in the One Period Model with One Risky Asset

Let us consider the one period model (c013-math-371) with one risky asset (c013-math-372). We assume for simplicity that c013-math-373. Let

equation

The absence of arbitrage implies that17

equation

We need to construct measure c013-math-376 so that

equation

Let

equation

for c013-math-377, where

equation

We can assume that c013-math-378 for each c013-math-379 such that c013-math-380.18 In addition, c013-math-382 and c013-math-383. We define the probability measure

equation

Now c013-math-384. We have that c013-math-385, and thus c013-math-386 is strictly convex on c013-math-387. Let

equation

We have to prove that

If (13.12) holds, then we define

equation

Now c013-math-389 is the required measure because c013-math-390 and

equation

Let us prove (13.12). Let us assume that (13.12) does not hold and derive a contradiction. Let c013-math-391 be a sequence such that

Then c013-math-393 or c013-math-394. Otherwise, we can choose a convergent subsequence, the minimum is attained at a finite point, and (13.12) holds. Let

equation

We have that

equation

Thus there exists c013-math-395 such that

equation

Thus,

equation

as c013-math-396. For sufficiently large c013-math-397 we have

equation

which contradicts (13.13).

A Martingale Measure in the Multiperiod Model with One Risky Asset

Let us consider the multiperiod model (c013-math-398) with one risky asset (c013-math-399). We assume for simplicity that c013-math-400. Let

equation

where c013-math-401. The absence of arbitrage implies that19

c013-math-409-almost surely, for c013-math-410. We need to construct measure c013-math-411 so that

equation

Then c013-math-412 is a martingale difference with respect to c013-math-413, and c013-math-414 is a martingale with respect to c013-math-415. Let

equation

for c013-math-416, We can assume that c013-math-417 is finite.20 For a fixed c013-math-421 function c013-math-422 is strictly convex, as follows from (13.14). There exists a unique finite

equation

such that c013-math-423 is attained at c013-math-424, which can be shown similarly as (13.12). We can show that c013-math-425 is c013-math-426-measurable.21 Let c013-math-430 and

equation

for c013-math-431. Now c013-math-432, c013-math-433 are c013-math-434-measurable, and they form a martingale:

equation

c013-math-435-almost surely. We define the probability measure

equation

Now c013-math-436, c013-math-437, and c013-math-438 for c013-math-439.

A Martingale Measure in the Multiperiod Model with Several Risky Assets

Let c013-math-440 and c013-math-441. We assume for simplicity that c013-math-442. Let

equation

where c013-math-443. Now c013-math-444 are vectors of length c013-math-445. The portfolio vector c013-math-446 is a c013-math-447-dimensional c013-math-448-measurable vector. The components are bounded, so that c013-math-449 for c013-math-450 and c013-math-451. The absence of arbitrage implies that

equation

c013-math-452-almost surely, for c013-math-453. We need to construct measure c013-math-454 so that

equation

Then c013-math-455 is a martingale difference with respect to c013-math-456, and c013-math-457 is a martingale with respect to c013-math-458. Let

equation

for c013-math-459. There exists a unique finite c013-math-460 such that c013-math-461 is attained at c013-math-462, and c013-math-463 is c013-math-464-measurable.22 Let c013-math-469 and

equation

for c013-math-470. We define the probability measure

equation

and c013-math-471 is the required equivalent martingale measure.

13.2.3.3 Estimation of an Equivalent Martingale Measure

We estimate the Esscher martingale measures using S&P 500 daily data, described in Section 2.4.1. We consider both a one period model and a two period model.

A Martingale Measure for S&P 500: One Period

We consider the one period model where the period consists of 20 days. Let

equation

be the price increment. Our S&P 500 data provides a sample of identically distributed observations of c013-math-472: we use data c013-math-473, where c013-math-474 and c013-math-475 is the gross return over the period of 20 days. We use nonoverlapping increments. The risk free rate is c013-math-476.

The density c013-math-477 of the Esscher martingale measure with respect to underlying physical measure of c013-math-478 can be estimated as

equation

where c013-math-479, c013-math-480 is the sample average of c013-math-481, and c013-math-482 is the minimizer of c013-math-483 over c013-math-484. The underlying physical density c013-math-485 of c013-math-486 with respect to the Lebesgue measure can be estimated using the kernel estimator c013-math-487. The kernel density estimator is defined in (3.43). The density c013-math-488 of the martingale measure with respect to the Lebesgue measure can be estimated as

equation

Figure 13.1(a) shows the estimate c013-math-489 of the density of the martingale measure with respect to the physical measure (red). The blue curve shows the density of the risk neutral log-normal density with respect to the estimated physical measure. We see that the measures put more probability mass on the negative increments than the physical measure. Fitting of a log-normal distribution is discussed in the connection of Figure 3.11.23 Panel (b) shows the kernel estimate c013-math-495 of the density of the physical measure as a red curve, and the estimate c013-math-496 of the density of the Esscher martingale measure with respect to the Lebesgue measure as a red dashed curve. We apply the standard normal kernel and the normal reference rule to choose the smoothing parameter. The blue curves show the corresponding densities in the log-normal model.

Illustration of Esscher martingale measure: One Period.

Figure 13.1 Esscher martingale measure: One Period. (a) The density of the Esscher martingale measure (red) and the density of the risk neutral log-normal measure (blue). The densities are with respect to the physical measure. (b) The kernel density estimate of the physical measure (red), and the corresponding Esscher martingale measure (red dashed). The log-normal physical measure (blue), and the corresponding risk neutral log-normal density (dashed blue).

Figure 13.2 shows density ratios. Panel (a) shows the ratio

equation

and panel (b) shows the ratios

equation

where c013-math-497 is an estimate of the log-normal physical measure, c013-math-498 is an estimate of the log-normal risk neutral measure, and c013-math-499.

Illustration of Esscher martingale measure: Density ratios in one period.

Figure 13.2 Esscher martingale measure: Density ratios in one period. (a) The density estimate of the Esscher martingale measure divided by the density estimate of the log-normal martingale measure. (b) The density estimate of the physical measure divided by the estimate log-normal physical measure (solid). The dashed line shows the ratio of the corresponding risk neutral densities.

A Martingale Measure for S&P 500: Two Periods

Let us estimate the Esscher martingale measure for the two period model with two periods of 10 days. Let

equation

be the price increments. Our S&P 500 data provides a sample of identically distributed observations of c013-math-500. The observations are

equation

where

equation

where c013-math-501. We use nonoverlapping increments. Let

equation

where c013-math-502 is the sample average of c013-math-503, and c013-math-504 is the minimizer of c013-math-505 over c013-math-506. Let

equation

where c013-math-507 is the minimizer of c013-math-508 over c013-math-509, and c013-math-510 is a regression estimate evaluated at c013-math-511, when the response variable is c013-math-512 and the explanatory variable is c013-math-513. We apply a kernel regression estimate of (6.20) and (6.21) to define

equation

where c013-math-514, c013-math-515, are the observation of c013-math-516,

equation

are the kernel weights, c013-math-517 is the Gaussian kernel function and c013-math-518 is the smoothing parameter, chosen by the normal reference rule.

The density c013-math-519 of the martingale measure with respect to the underlying physical measure of c013-math-520 can be estimated as

We can also assume the independence of the increments and estimate c013-math-522 by

Figure 13.3 shows estimates of the density of the Esscher measure with respect to the physical measure. In panel (a) we show estimate (13.15), which does not assume independence, and in panel (b) we show estimate (13.16), which assumes independence.

Illustration of Esscher martingale measure: Two Periods.

Figure 13.3 Esscher martingale measure: Two Periods. Estimates of the density of the Esscher measure with respect to the physical measure. (a) Increments are assumed to be dependent. (b) Increments are assumed to be independent.

13.2.3.4 Examples of Equivalent Martingale Measures

We calculate the class of equivalent martingale measures in the one period binary model, in the one period ternary model, and in the one period model with a finite amount of states, which generalizes the two previous models.

The One Period Binary Model

Let us have two assets: bond c013-math-524 and stock c013-math-525. The value of the bond at time 1 is c013-math-526, where c013-math-527. The value of the stock at time 1 is c013-math-528 with probability c013-math-529 and c013-math-530 with probability c013-math-531, where c013-math-532 and c013-math-533. That is,

equation

Let the price of the bond be c013-math-534 and the price of the stock be c013-math-535. Let us consider probability measure c013-math-536 which is defined by

equation

where c013-math-537. If c013-math-538 is a martingale measure, then it satisfies

equation

This holds if

Thus, there exists an equivalent martingale measure if and only if

Thus, the market is arbitrage-free if and only if (13.18) holds. The martingale measure is unique. The calculation will be repeated in (14.18), where derivative pricing is discussed.24 Note that the pricing in the binary model was already studied in (13.1).

The One Period Ternary Model

Let us have two assets: bond c013-math-549 and stock c013-math-550. The value of the bond at time 1 is c013-math-551, where c013-math-552. The value of the stock at time 1 is c013-math-553 with probability c013-math-554, c013-math-555 with probability c013-math-556, and c013-math-557 with probability c013-math-558, where c013-math-559, and c013-math-560. That is,

equation

Let the price of the bond be c013-math-561 and the price of the stock be c013-math-562. Let us consider probability measure c013-math-563, which is defined by

equation

where c013-math-564 and c013-math-565. If c013-math-566 is a martingale measure, then it satisfies

equation

This holds if c013-math-567, where

equation

We can write

equation

Thus, there exists an equivalent martingale measure if and only if

Thus, the market is arbitrage-free if and only if (13.20) holds. There are several martingale measures.25

A Finite Amount of States

Let us consider the one period model with a finite amount of states. Now the probability space c013-math-574 has a finite number of elements. We have c013-math-575 basic securities and c013-math-576 possible states. In the binary model c013-math-577. In the ternary model c013-math-578 and c013-math-579.

The c013-math-580th risky asset c013-math-581 takes c013-math-582 values, corresponding to the c013-math-583 different states. Let c013-math-584 be the c013-math-585 matrix whose elements are c013-math-586, where c013-math-587 is the c013-math-588th state and c013-math-589 is the c013-math-590th risky asset.

Let c013-math-591 be the c013-math-592 vector of the probabilities of the c013-math-593 states. Let c013-math-594 be the c013-math-595 vector of the prices of the risky assets at time 0. Let c013-math-596 be the vector of the risky assets.

Let c013-math-597 be a c013-math-598 vector of probabilities of the c013-math-599 states. Vector c013-math-600 is a martingale measure if

We can assume that c013-math-602 and c013-math-603, because the redundant basic assets can be removed. A redundant asset would correspond to a column of c013-math-604 which could be expressed as a linear combination of the other columns.

When c013-math-605, then there exists a unique equivalent martingale measure, and this is the solution to (13.21):

When c013-math-607, then the system (13.21) of c013-math-608 linear equations with c013-math-609 variables has many solutions.

13.2.4 European Contingent Claims

13.2.4.1 The Definition of an European Continent Claim

We use the terms “contingent claim” and “derivative” interchangeably. However, these terms can have a different meaning.

  1. 1. A European contingent claim is a nonnegative random variable c013-math-610 defined on the probability space c013-math-611.
  2. 2. A derivative of the underlying assets c013-math-612 is a European contingent claim, which is measurable with respect to the c013-math-613-algebra c013-math-614 generated by the price process c013-math-615, c013-math-616.

We assume that c013-math-617, and thus in our case the two definitions lead to the same concept. Time c013-math-618 is called the maturity, or the expiration date.26 The examples of European contingent claims include the following.

  1. 1. A call and a put on the c013-math-620th asset are defined by
    equation
  2. where c013-math-621 is the strike price and c013-math-622.
  3. 2. An Asian call and put option are defined by
    equation
  4. where
    equation
  5. c013-math-623, and c013-math-624 is the cardinality of c013-math-625.
  6. 3. A knock-out option on the c013-math-626th asset is defined by
    equation
  7. where c013-math-627 is the strike price, and c013-math-628 is the barrier.

13.2.4.2 Arbitrage-Free Prices of European Continent Claims

A European contingent claim c013-math-629 is attainable (replicable, redundant), if there exists a self-financing trading strategy c013-math-630 whose terminal portfolio value is equal to c013-math-631:

equation

The trading strategy c013-math-632 is called a replicating strategy for c013-math-633. A contingent claim is attainable if and only if the discounted claim

equation

has the form

equation

for a self-financing trading strategy c013-math-634 with value process c013-math-635. Now it is natural to take the initial value

equation

to be the fair price of c013-math-636, because a different price would lead to an arbitrage opportunity. The corresponding arbitrage-free price of the contingent claim c013-math-637 is

equation

We need to define an arbitrage-free price also for those contingent claims which are not attainable. In fact, in typical market models a contingent claim cannot be replicated. Föllmer and Schied (2002, p. 238) formulate the following definition. An arbitrage-free price of a discounted claim c013-math-638 is a real number c013-math-639, if there exists an adapted stochastic process c013-math-640 such that

  1. 1. c013-math-641,
  2. 2. c013-math-642 for c013-math-643,
  3. 3. c013-math-644, and
  4. 4. the enlarged market model with price process c013-math-645 is arbitrage-free.

According to this definition, an arbitrage-free price of a contingent claim is such that trading with this price at time 0 does not allow an arbitrage opportunity. A corresponding arbitrage-free price of the continent claim c013-math-646 is then

equation

We can express the class of arbitrage-free prices with the help of equivalent martingale measures. Föllmer and Schied (2002, Theorem 5.30, p. 239) formulate the following theorem.

The price of contingent claim c013-math-673 can now be written as

In the Black–Scholes model we use continuous compounding, where c013-math-675, and c013-math-676, so that for a call option c013-math-677; see (14.47). In many cases we denote by c013-math-678 the time of writing the option and then c013-math-679, so that c013-math-680.

13.2.4.3 Pricing Kernel

The pricing kernel (discount factor) c013-math-681, related to the martingale measure c013-math-682, is defined as the discounted density of c013-math-683 with respect to the physical measure c013-math-684:

equation

The price of c013-math-685 is given in (13.23) as c013-math-686. Now the price of derivative c013-math-687 can be written as

equation
The One Period Binary Model

In the one period binary model, the martingale measure was defined as the measure

equation

where the probability c013-math-688 is defined in (13.17). The pricing kernel is function c013-math-689, defined by

equation

Let c013-math-690 be a derivative, where c013-math-691. The price of c013-math-692 is

equation
A Finite Amount of States

Let us continue to study the one period model with a finite amount of states. Let c013-math-693 be the c013-math-694 vector of the probabilities of the c013-math-695 states. Let c013-math-696 be the c013-math-697 vector of the prices of the risky assets at time 0. Let c013-math-698 be the vector of the risky assets. Let c013-math-699 be the c013-math-700 matrix whose elements are the values c013-math-701 of the c013-math-702th asset at the c013-math-703th state.

Let c013-math-704 be an equivalent martingale measure, which is a solution to the equation in (13.21). In the case c013-math-705 we may obtain c013-math-706 from (13.22) as c013-math-707. Let

equation

for c013-math-708. Let c013-math-709 be a derivative. The price of c013-math-710 is

equation

The Arrow–Debreu securities take value 1 in one state and value 0 in other states: for the c013-math-711th Arrow–Debreu security c013-math-712 it holds that c013-math-713, and c013-math-714 for c013-math-715. Then c013-math-716 is the c013-math-717 identity matrix. Then c013-math-718 and c013-math-719.27

13.2.5 Completeness

The second fundamental theorem of asset pricing says that every European contingent claim can be attained (replicated) if and only if there exists a unique equivalent martingale measure. If every contingent claim can be attained, then every contingent claim has a unique arbitrage-free price and every contingent claim can be hedged perfectly. The case that there is only one equivalent martingale measure occurs never in practice, but it is possible to be close to this situation.

An arbitrage-free market model is called complete if every European contingent claim is attainable. Now we state the second fundamental theorem of asset pricing.

A proof can be found in Föllmer and Schied (2002, Theorem 5.38, p. 245), where an additional statement is proved: In a complete market, the number of atoms in c013-math-734 is bounded above by c013-math-735.28

Let us give examples related to completeness.

13.2.5.1 The One Period Binary Model

In the one period binary model, we have two assets: bond c013-math-743 and stock c013-math-744. The bond satisfies c013-math-745, where c013-math-746. The stock satisfies

equation

where c013-math-747 and c013-math-748. Let the price of the bond be c013-math-749 and the price of the stock be c013-math-750. The space of attainable payoffs is

equation

Let us consider contingent claim c013-math-751, where c013-math-752. To replicate the contingent claim, we need to choose c013-math-753 and c013-math-754 so that

equation

This leads to equations

equation

We have two equations and two free variables. The model is complete.

13.2.5.2 The One Period Ternary Model

Let us have two assets: bond c013-math-755 and stock c013-math-756. The bond satisfies c013-math-757, where c013-math-758. The stock satisfies

equation

where c013-math-759, and c013-math-760. Let the price of the bond be c013-math-761 and the price of the stock be c013-math-762. The space of attainable payoffs is

equation

Let us consider contingent claim c013-math-763, where c013-math-764. To replicate the contingent claim, we need to choose c013-math-765 and c013-math-766 so that

equation

This leads to equations

equation

We have three equations and two free variables. The model is not complete.

13.2.5.3 A Finite Amount of States

Let us continue to study the one period model with a finite amount of states. Let c013-math-767 be the c013-math-768 vector of the probabilities of the c013-math-769 states. Let c013-math-770 be the c013-math-771 vector of the prices of the risky assets at time 0. Let c013-math-772 be the vector of the risky assets. Let c013-math-773 be the c013-math-774 matrix whose elements are the values c013-math-775 of the c013-math-776th asset at the c013-math-777th state.

We can assume that c013-math-778 and c013-math-779, because the redundant basic assets can be removed. A redundant asset would correspond to a column of c013-math-780 which could be expressed as a linear combination of the other columns. A redundant basic asset could be considered as a derivative.

A derivative security is random variable c013-math-781 which takes c013-math-782 possible values. Let those values be in c013-math-783 vector c013-math-784. To replicate c013-math-785, we need to find c013-math-786 vector c013-math-787 so that c013-math-788. This leads to the matrix equation

equation

When c013-math-789, then

equation

When c013-math-790, then we do not always have a solution, because there are c013-math-791 free variables and c013-math-792 equations.

We can choose an approximate replication by minimizing the sum of squared replication errors. Let the replication error be

equation

where c013-math-793 is the Euclidean norm in c013-math-794. The minimizer is

equation

which is the same formula as the formula for the least squares coefficients in the linear regression c013-math-795. Note that c013-math-796 matrix c013-math-797 has rank c013-math-798, when c013-math-799 has rank c013-math-800, and thus c013-math-801 is invertible.

The Arrow–Debreu securities provide an example of derivatives. An Arrow–Debreu security has price 1 in one state and 0 in the other states. There are as many Arrow–Debreu securities as there are states. When c013-math-802, the columns of c013-math-803 give the portfolio weights for replicating the Arrow–Debreu securities.

13.2.6 American Contingent Claims

An American contingent claim is defined as a non-negative adapted process

equation

on the filtered space c013-math-804. The random variable c013-math-805 is the payoff if the American option is exercised at time c013-math-806. For example, in the case of the American call option with strike price c013-math-807, c013-math-808, where c013-math-809 is the price of the underlying asset at time c013-math-810.

The buyer of an American contingent claim has the right to choose the exercise time c013-math-811. The buyer receives the amount c013-math-812 at time c013-math-813.

A stopping time is a random variable c013-math-814 taking values in c013-math-815 such that c013-math-816 for c013-math-817. An exercise strategy is a stopping time taking values in c013-math-818. The payoff obtained by using c013-math-819 is equal to c013-math-820. We denote with c013-math-821 the set of exercise strategies.

13.2.6.1 European and Bermudan Options

An European contingent claim is obtained as a special case of an American contingent claim, when we choose c013-math-822 for c013-math-823. The value of the American option is larger or equal to the value of the corresponding European option.

A Bermudan option can be exercised at times c013-math-824. Formally we can define a Bermudan contingent claim as a non-negative adapted process c013-math-825, c013-math-826, on the filtered space c013-math-827. A Bermudan option can be obtained as an American option with c013-math-828 for c013-math-829.

On the other hand, an American option can be considered as a special case of a Bermudan option with c013-math-830. Also, from the point of view of the continuous time model with time space c013-math-831, an American option in a discrete time model could be considered as a Bermudan option whose possible exercise times are c013-math-832.

13.2.6.2 The Set of Arbitrage-Free Prices

Let c013-math-833 be a discounted American claim and let c013-math-834 be the payoff which is obtained for a fixed exercise strategy c013-math-835. Now c013-math-836 can be considered as a discounted European contingent claim, whose set of arbitrage-free prices is given in Theorem 13.2 as

equation

Föllmer and Schied (2002, Definition 6.31) give the following definition for an arbitrage-free price. A number c013-math-837 is called an arbitrage-free price of a discounted American claim c013-math-838 if

  1. 1.

    There exists some c013-math-839 and c013-math-840 such that c013-math-841.

    (The price c013-math-842 is not too high.)

  2. 2.

    There does not exist c013-math-843 such that c013-math-844 for all c013-math-845.

    (The price c013-math-846 is not too low.)

The set of arbitrage-free prices is characterized in Föllmer and Schied (2002, Theorem 6.33). It is assumed that c013-math-847 for all c013-math-848 and c013-math-849. The set c013-math-850 of arbitrage-free prices is an interval with endpoints c013-math-851 and c013-math-852, where

equation

The interval can be a single point, open, or half open.

13.2.6.3 Exercise Strategies for the Buyer

Let us call an exercise strategy c013-math-853 optimal if

where

equation

Thus, an exercise strategy is defined to be optimal, if it maximizes the expectation among the class c013-math-855 of the payoffs. Note that the definition can be generalized so that we maximize c013-math-856, where c013-math-857 is a utility function, and c013-math-858 is a probability measure, possibly different from the physical measure c013-math-859.

Föllmer and Schied (2002, Theorem 6.20) shows that the exercise strategy c013-math-860 is optimal, when we define

equation

where

It is assumed that c013-math-862 for c013-math-863. Process c013-math-864 is called a Snell envelope of c013-math-865 with respect to the measure c013-math-866. Föllmer and Schied (2002, Proposition 6.22) notes that any optimal exercise strategy c013-math-867 satisfies c013-math-868, so that c013-math-869 is the minimal optimal exercise strategy.

Föllmer and Schied (2002, Theorem 6.23) shows that c013-math-870 is also an optimal exercise strategy, when we define

equation

where c013-math-871 means c013-math-872. In addition, c013-math-873 is the largest optimal exercise strategy in the sense that any optimal exercise strategy c013-math-874 satisfies c013-math-875.

13.2.6.4 American Options in Complete Models

Let us assume that the market model is complete, so that there exists the unique equivalent martingale measure c013-math-876. We defined the optimal exercise time in (13.24) using the physical measure c013-math-877, and the construction of the optimal exercise time was made in (13.25) using the physical measure c013-math-878. Let us define the Snell envelope c013-math-879 using the equivalent martingale measure c013-math-880. The value c013-math-881 can be considered as the unique arbitrage-free price, because

equation

holds c013-math-882-almost surely, when c013-math-883 is optimal with respect to c013-math-884; see Föllmer and Schied (2002, Corollary 6.24).

13.3 Evaluation of Pricing and Hedging Methods

The evaluation of option pricing and hedging can be done either from the point of view of the writer or from the point of view of the buyer. The writer's point of view is to minimize the hedging error, or to optimize the return of the hedging portfolio. The buyer's point of view is to find fair prices for options. For example, the buyer could be interested whether the buying of the options leads to abnormal returns, as compared to the returns of the underlying.

13.3.1 The Wealth of the Seller

We assume that the seller (writer) of the option hedges the position, and thus the wealth of the seller of the option at the expiration is equal to the hedging error. The writer receives the option premium, makes self-financed trading to replicate the option, and pays the terminal value of the option. We consider the pricing to be fair and the hedging to be effective if the distribution of the hedging error (replication error) is as concentrated around zero as possible. However, we have to study separately the negative hedging errors (losses) and the positive hedging errors (gains).

13.3.1.1 Hedging Error

The hedging error c013-math-885 of the writer of the option is obtained from (13.10) as

equation

where

equation

Here the risk free rate is c013-math-886, c013-math-887 is the price of the option, c013-math-888 is the terminal value of the option, c013-math-889 are the hedging coefficients, c013-math-890 are the stock prices, the current time is denoted by 0, the time to expiration is c013-math-891 days, and hedging is done daily.

13.3.1.2 Historical Simulation

We denote the time series of observed historical daily prices by c013-math-892. We construct c013-math-893 sequences of prices:

equation

where

equation

for c013-math-894. Each sequence has length c013-math-895 and the initial price in each sequence is c013-math-896. We estimate the distribution of the hedging error c013-math-897 from the observations

equation

where c013-math-898 is computed from the prices c013-math-899.

An example of a computation of hedging errors is given in (14.81), where Black-Scholes hedging is applied with sequential sample standard deviations as the volatility estimates. See also (14.80), where Black-Scholes hedging is applied with non-sequential sample standard deviations as the volatility estimates.

13.3.1.3 Comparison of the Error Distributions

We estimate the distribution of the hedging error c013-math-900 using data c013-math-901. A graphical summary of the error distribution is obtained by using tail plots and kernel density estimation, for example.

The Mean of Squared Hedging Errors

In quadratic hedging the purpose is to minimize the mean squared hedging errors c013-math-902; see Section 15.1 and Chapter 16. Thus it is natural to estimate the quality of a hedging strategy by the sample mean of squared hedging errors

equation

We can decompose the mean into the mean over negative hedging errors and over positive hedging errors:

equation

where

equation

This composition is reasonable because the negative hedging errors are losses for the writer of the option and the positive hedging errors are gains for the writer of the option.

The Expected Utility of the Error Distribution

Even when the purpose of the hedging is typically to minimize the hedging error and not to maximize the wealth of the hedger, it is of interest to study the properties of the wealth distribution from the point of view of portfolio theory. This can be done by estimating the expected utility of the error distribution. We can use the exponential utility function c013-math-903, where c013-math-904 is the risk aversion parameter. Note that the hedging errors c013-math-905 can take any real value, and thus we cannot apply the power utility functions. The expected utility is estimated by

equation

See Section 9.2.2 for a discussion about expected utility.

13.3.2 The Wealth of the Buyer

We can ignore the hedging of the options and try to evaluate solely the fairness of the price. This approach can be considered to be the approach of the buyer of the option. We have at least the following possibilities.

  1. 1. Comparison with the market prices. The success of a pricing approach can be evaluated by testing whether the prices of the approach provide a good fit to the observed market prices of the derivatives.

    The comparison with the market prices is possible for liquid options. Note that a pricing method for illiquid options can be obtained by calibrating the parameters of the pricing method using liquid options. For example, the implied volatility of Black–Scholes pricing can be obtained from liquid options and then used as the volatility of the Black–Scholes formula to price illiquid options.

  2. 2. Sharpe ratios. We can estimate the Sharpe ratios of option strategies. The Sharpe ratios obtained by option buying should not be too far away from the Sharpe ratios of the underlying assets. This is illustrated in Section 17.2.3.
equation

and thus the expected gain is

equation

However, a typical price for a Finnish lottery of this type is 0.8 € and playing the game with this entrance price results in the expected loss of 0.67 €. Using the logarithmic utility gives a positive utility

equation

which would give the price of c013-math-022 €. Thus, the market price of a lottery can be justified by pointing out that the market price is equal to the expected utility, for some utility function. Intuitively, a lottery is attractive because it provides an opportunity to a dramatic improvement of wealth, with a negligible price.

If c013-math-288 is a martingale, then c013-math-289 is called a martingale transformation. In our case c013-math-290.

equation
equation

for each c013-math-300.

equation
equation

where c013-math-428 is the set of rational numbers and c013-math-429.

where c013-math-542 is the convex set of probability measures equivalent to c013-math-543 and c013-math-544. Now c013-math-545. If c013-math-546, then there is arbitrage. If c013-math-547, then arbitrage is obtained by borrowing with the risk-free rate and buying the stock. If c013-math-548, then the arbitrage is obtained by selling the stock short and investing in the risk-free rate.

equation

Let c013-math-725 and let c013-math-726 be the vector of c013-math-727 Arrow–Debreu securities Then, c013-math-728, and c013-math-729, where c013-math-730 is the c013-math-731th state, c013-math-732.

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

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