2.3Expected utility

In this section, we focus on individual financial assets under the assumption that their payoff distributions at a fixed time are known, and without any regard to hedging opportunities in the context of a financial market model. Such asset distributions may be viewed as lotteries with monetary outcomes in some interval on the real line. Thus, we take M as a fixed set of Borel probability measures on a fixed interval S . In this setting, we discuss the paradigm of expected utility in its standard form, where the function u appearing in the von NeumannMorgenstern representation has additional properties suggested by the monetary interpretation. We introduce risk aversion and certainty equivalents, and illustrate these notions with a number of examples.

Throughout this section, we assume that M is convex and contains all point masses δx for x S. We assume also that each μ M has a well-defined expectation

Remark 2.31. For an asset whose (discounted) random payoff has a known distribution μ, the expected value m(μ) is often called the fair price of the asset. For an insurance contract where μ is the distribution of payments to be received by the insured party in dependence of some random damage within a given period, the expected value m(μ) is also called the fair premium. Typically, actual asset prices and actual insurance premiums will be different from these values. In many situations, such differences can be explained within the conceptual framework of expected utility, and in particular in terms of risk aversion.

Definition 2.32. A preference relation on M is called monotone if

The preference relation is called risk averse if for μ M

It is easy to characterize these properties within the class of preference relations which admit a von NeumannMorgenstern representation.

Proposition 2.33. Suppose the preference relation has a von NeumannMorgenstern representation

Then:

(a) is monotone if and only if u is strictly increasing.

(b) is risk averse if and only if u is strictly concave.

Proof. (a): Monotonicity is equivalent to

(b): If is risk-averse, then

holds for all distinct x, y S and α (0, 1). Hence,

i.e., u is strictly concave. Conversely, if u is strictly concave, then Jensens inequality implies risk aversion:

with equality if and only if μ = δm(μ).

Remark 2.34. In view of the monetary interpretation of the state space S, it is natural to assume that the preference relation is monotone. The assumption of risk aversion is more debatable, at least from a descriptive point of view. In fact, there is considerable empirical evidence that agents tend to switch between risk aversion and risk seeking behavior, depending on the context. In particular, they may be risk averse after prior gains, and they may become risk seeking if they see an opportunity to compensate prior losses. Tversky and Kahneman [272] propose to describe such a behavioral pattern by a function u of the form

where c is a given benchmark level, and their experiments suggest parameter values λ around 2 and γ slightly less than 1. Nevertheless, one can insist on risk aversion from a normative point of view, and in the sequel we explore some of its consequences.

Definition 2.35. A function u : S is called a utility function if it is strictly concave, strictly increasing, and continuous on S. A von Neumann-Morgenstern representation

in terms of a utility function u is called an expected utility representation.

Any increasing concave function u : S is necessarily continuous on every interval (a, b] S; see Proposition A.7. Hence, the condition of continuity in the preceding definition is only relevant if S contains its lower boundary point. Note that any utility function u(x) decreases at least linearly as x inf S. Therefore, u cannot be bounded from below unless inf S > .

From now on, we will consider a fixed preference relation on M which admits a von NeumannMorgenstern representation

in terms of a strictly increasing continuous function u : S . The intermediate value theorem applied to the function u yields for any μ M a unique real number c(μ) for which

It follows that

i.e., there is indifference between the lottery μ and the sure amount of money c(μ).

Definition 2.36. The certainty equivalent of the lottery μ M with respect to u is defined as the number c(μ) of (2.11), and

is called the risk premium of μ.

If u is a utility function, risk aversion implies that

for every lottery μ with μ δm(μ). Hence, monotonicity yields that

In particular, the risk premium ϱ(μ) associated with a utility function is always nonnegative, and it is strictly positive as soon as the distribution μ carries any risk.

Remark 2.37. The certainty equivalent c(μ) can be viewed as an upper bound for any price of μ which would be acceptable to an economic agent with utility function u. Thus, the fair price m(μ) must be reduced at least by the risk premium ϱ(μ) if one wants the agent to buy the asset distribution μ. Alternatively, suppose that the agent holds an asset with distribution μ. Then the risk premium may be viewed as the amount that the agent would be ready to pay for replacing the asset by its expected value m(μ).

Example 2.38 (St. Petersburg Paradox). Consider the lottery

which may be viewed as the payoff distribution of the following game. A fair coin is tossed until a head appears. If the head appears on the nth toss, the payoff will be 2n1. Up to the early 18th century, it was commonly accepted that the price of a lottery should be computed as the fair price, i.e., as the expected value m(μ). In the present example, the fair price is given by m(μ) = ,but it is hard to find someone who is ready to pay even 20. In view of this paradox, posed by Nicholas Bernoulli in 1713, Gabriel Cramer and Daniel Bernoulli [26] independently introduced the idea of determining an acceptable price as the certainty equivalent with respect to some utility function. For the two utility functions

proposed, respectively, by G. Cramer and by D. Bernoulli, these certainty equivalents are given by

and this is within the range of prices people are usually ready to pay. Note, however, that for any utility function which is unbounded from above we could modify the payoff in such a way that the paradox reappears. For example, we could replace the payoff 2n by u1(2n) for n 1000, so that The choice of a utility function that is bounded from above would remove this difficulty, but would create others; see the discussion on pp. 8589.

Exercise 2.3.1. Prove (2.12).

Given the preference order on M, we can now try to determine those distributions in M which are maximal with respect to . As a first illustration, consider the following simple optimization problem. Let X be an integrable random variable on some probability space (Ω,F, P) with nondegenerate distribution μ M. We assume that X is bounded from below by some number a in the interior of S. What is the best mix

of the risky payoff X and the certain amount c, which also belongs to the interior of S? If we evaluate X λ by its expected utility E [ u(Xλ) ] and denote by μλ the distribution of X λ under P, then we are looking for a maximum of the function f on [0, 1] defined by

If u is a utility function, f is strictly concave and attains its maximum in a unique point λ [0, 1].

Proposition 2.39. Let u be a utility function.

(a) We have λ = 1 if E[ X ] c, and λ > 0 if c c(μ).

(b) If u is differentiable, then

and

Proof. (a): Jensens inequality yields that

with equality if and only if λ = 1. It follows that λ = 1 if the right-hand side is increasing in λ, i.e., if E[ X ] c.

Strict concavity of u implies

with equality if and only if λ {0, 1}. The right-hand side is increasing in λ if c c(μ), and this implies λ > 0.

(b): Clearly, we have λ = 0 if and only if the right-hand derivative of f satisfies see Appendix A.1 for the definition of and Note that the difference quotients

satisfy P-a.s.

and that they converge to

as λ 0. By Lebesgues theorem, this implies

If u is differentiable, or if the countable set {x | +(x) (x)} has μ-measure 0, then we can conclude

i.e., if and only if

In the same way, we obtain

If u is differentiable at c, then we can conclude

This implies (1) < 0, and hence λ < 1, if and only if E[ X ] > c.

Exercise 2.3.2. As above, let X be a random variable with a nondegenerate distribution μ M. Show that for a differentiable utility function u we have

Example 2.40 (Demand for a risky asset). Let S = S1 be a risky asset with price π = π1. Given an initial wealth w, an agent with utility function u C1 can invest a fraction (1 λ)w into the asset and the remaining part λw into a risk-free bond with interest rate r. The resulting payoff is

The preceding proposition implies that there will be no investment into the risky asset if and only if

In other words, the price of the risky asset must be below its expected discounted payoff in order to attract any risk averse investor, and in that case it will indeed be optimal for the investor to invest at least some amount. Instead of the simple linear profiles X λ, the investor may wish to consider alternative forms of investment. For example, this may involve derivatives such as max{S, K} = K + (S K)+ for some threshold K. In order to discuss such nonlinear payoff profiles, we need an extended formulation of the optimization problem; see Section 3.3 below.

Example 2.41 (Demand for insurance). Suppose an agent with utility function u C1 considers taking at least some partial insurance against a random loss Y, with 0 Y w and P[ Y E[ Y ] ] > 0, where w is a given initial wealth. If insurance of λY is available at the insurance premium λπ, the resulting final payoff is given by

By Proposition 2.39, full insurance is optimal if and only if π E [ Y ]. In reality, however, the insurance premium π will exceed the fair premium E[ Y ]. In this case, it will be optimal to insure only a fraction λY of the loss, with λ [0, 1). This fraction will be strictly positive as long as

Since the right-hand side is strictly larger than E[ Y ] due to (2.13), risk aversion may create a demand for insurance even if the insurance premium π lies above the fair price E[ Y ]. As in the previous example, the agent may wish to consider alternative forms of insurance such as a stop-loss contract, whose payoff has the nonlinear structure (Y K )+ of a call option.

Let us take another look at the risk premium ϱ(μ) of a lottery μ. For an approximate calculation, we consider the Taylor expansion of a sufficiently smooth and strictly increasing function u(x) at x = c(μ) around m := m(μ), and we assume that μ has finite variance var(μ). On the one hand,

On the other hand,

where r(x) denotes the remainder term in the Taylor expansion of u. It follows that

Thus, α(m(μ)) is the factor by which an economic agent with von Neumann-Morgenstern preferences described by u weighs the risk, measured by in order to determine the risk premium he or she is ready to pay.

Definition 2.42. Suppose that u is a twice continuously differentiable and strictly increasing function on S. Then

is called the ArrowPratt coefficient of absolute risk aversion of u at level x.

Example 2.43. The following classes of utility functions u and their corresponding coefficients of risk aversion are standard examples.

(a) Constant absolute risk aversion (CARA): α(x) equals some constant α > 0. Since α(x) = (log )ʹ(x), it follows that u(x) = a b · eαx. Using an affine transformation, u can be normalized to

(b) Hyperbolic absolute risk aversion (HARA): α(x) = (1 γ)/x on S = (0, ) for some γ < 1. Up to affine transformations, we have

Sometimes, these functions are also called CRRA utility functions, because their relative risk aversion (x) is constant. Of course, these utility functions can be shifted to any interval S = (a,). The risk-neutral limiting case γ = 1 would correspond to an affine function u.

Exercise 2.3.3. Compute the coefficient of risk aversion for the S-shaped utility function in (2.9). Sketch the graphs of u and its risk aversion for λ = 2 and γ = 0.9.

Proposition 2.44. Suppose that u and are two strictly increasing functions on S which are twice continuously differentiable, and that α and are the corresponding ArrowPratt coefficients of absolute risk aversion. Then the following conditions are equivalent.

(a) α(x) (x) for all x S.

(b) u = F ◦ for a strictly increasing concave function F.

(c) The respective risk premiums ϱ and associated with u and satisfy ϱ(μ) (μ) for all μ M.

Proof. (a)(b): Since is strictly increasing, we may define its inverse function, w.

Then F (t) := u w(t)is strictly increasing, twice differentiable, and satisfies u = F ◦ .

For showing that F is concave we calculate the first two derivatives of w:

Now we can calculate the first two derivatives of F:

and

This proves that F is concave.

(b)(c): Jensens inequality implies that the respective certainty equivalents c(μ) and (μ) satisfy

Hence, ϱ(μ) = m(μ) c(μ) m(μ) (μ) = (μ).

(c)(a): If condition (a) is false, there exists an open interval O S such that (x) > α(x) for all x O. Let := (O), and denote again by w the inverse of . Then the function F (t) = u (w(t)) will be strictly convex in the open interval by (2.15). Thus, if μ is a measure with support in O, the inequality in (2.16) is reversed and is even strict unless μ is concentrated at a single point. It follows that ϱ(μ) < (μ), which contradicts condition (c).

As an application of the preceding proposition, we will now investigate the structure of those continuous and strictly increasing functions u on whose associated certainty equivalents have the following translation property:

where the translation μt of μ M by t is defined by

Here we also assume that M is closed under translation, i.e., μt M for all μ M and t .

Lemma 2.45. Suppose the certainty equivalent associated with a continuous and strictly increasing function u : satisfies the translation property (2.17). Then u belongs to C().

Proof. Let λ denote the Lebesgue measure on [0, 1]. Then

and this implies f C1() with

Thus, u(x) = f (x c(λ)) is in C1(), which implies that C1() by (2.19), hence f C2(). Iterating the argument we get u C().

The following proposition implies in particular that a utility function u that satisfies the translation property (2.17) is necessarily a CARA utility function of exponential type as in part (a) of Example 2.43.

Proposition 2.46. Suppose the certainty equivalent associated with a continuous and strictly increasing function u : satisfies the translation property (2.17). Then u has constant absolute risk aversion and is hence either linear or an exponential function. More precisely, there are constants a and b, α > 0 such that u(x) equals one of the following three functions

Proof. For t let ut(x) := u(x + t), and denote by ct(μ) the corresponding certainty equivalent. For μ M we have

It follows that ct(μ) = c(μ) for all t and μ M. Therefore,

for all t . Since u is smooth by Lemma 2.45, we may apply Proposition 2.44 to conclude that the respective Arrow-Pratt coefficients αt(x) = utʹʹ (x)/utʹ (x) and α(x) = uʹʹ(x)/(x) are equal for all t and x. But αt(x) = α(x + t), and so α(x) does not depend on x. If α > 0, we see as in Example 2.43 that u is of the form u(x) = a beαx. If α = 0, u must be linear. And if α < 0, we must have u(x) = a + beαx.

Now we focus on the case in which u is a utility function and preferences have the expected utility representation (2.10). In view of the underlying axioms, the paradigm of expected utility has a certain plausibility on a normative level, i.e., as a guideline of rational behavior in the face of risk. But this guideline should be applied with care: If pushed too far, it may lead to unplausible conclusions. In the remaining part of this section we discuss some of these issues. From now on, we assume that S is unbounded from above, so that w + x S for any x S and w 0. So far, we have implicitly assumed that the preference relation on lotteries reflects the views of an economic agent in a given set of conditions, including a fixed level w 0 of the agents initial wealth. In particular, the utility function may vary as the level of wealth changes, and so it should really be indexed by w. Usually one assumes that uw is obtained by simply shifting a fixed utility function u to the level w, i.e., uw(x) := u(w + x). Thus, a lottery μ is declined at a given level of wealth w if and only if

Let us now return to the situation of Proposition 2.39 where μ is the distribution of an integrable random variable X on (Ω,F, P), which is bounded from below by some number a in the interior of S. We view X as the net payoff of some financial bet, and we assume that the bet is favorable in the sense that

Remark 2.47. Even though the favorable bet X might be declined at a given level w due to risk aversion, it follows from Proposition 2.39 that it would be optimal to accept the bet at some smaller scale, i.e., there is some γ > 0 such that

On the other hand, it follows from Proposition 2.49 below that the given bet X becomes acceptable at a sufficiently high level of wealth whenever the utility function is unbounded from above.

Sometimes it is assumed that some favorable bet is declined at every level of wealth. The assumption that such a bet exists is not as innocent as it may look. In fact it has rather drastic consequences. In particular, we are going to see that it rules out all utility functions in Example 2.43 except for the class of exponential utilities.

Example 2.48. For any exponential utility function u(x) = 1 e αx with constant risk aversion α > 0, the induced preference order on lotteries does not depend at all on the initial wealth w. To see this, note that

is equivalent to

Let us now show that the rejection of some favorable bet μ at every wealth level w leads to a not quite plausible conclusion: At high levels of wealth, the agent would reject a bet ν with huge potential gain even though the potential loss is just a negligible fraction of the initial wealth.

Proposition 2.49. If the favorable bet μ is rejected at any level of wealth, then the utility function u is bounded from above, and there exists A > 0 such that the bet

is rejected at any level of wealth.

Proof. We have assumed that X is bounded from below, i.e., μ is concentrated on [a,) for some a < 0, where a is in the interior of S. Moreover, we can choose b > 0 such that

is still favorable. Since u is increasing, we have

for any w 0, i.e., also the lottery is rejected at any level of wealth. It follows that

Let us assume for simplicity that u is differentiable; the general case requires only minor modifications. Then the previous inequality implies

where

due to the fact that is favorable. Thus,

for any w, hence

for any x in the interior of S. This exponential decay of the derivative implies u() := limx u(x) < . More precisely, if A := n(|a| + b) for some n, then

Take n such that γn 1/2. Then we obtain

i.e.,

for all x such that x A S.

Example 2.50. For an exponential utility function u(x) = 1 eαxν , the bet defined in the preceding lemma is rejected at any level of wealth as soon as

Suppose now that the lottery μ M is played not only once but n times in a row. For instance, one can think of an insurance company selling identical policies to a large number of individual customers. More precisely, let (Ω,F, P) be a probability space supporting a sequence X1, X2, . . . of independent random variables with common distribution μ. The value of X i will be interpreted as the outcome of the ith drawing of the lottery μ. The accumulated payoff of n successive independent repetitions of the financial bet X1 is given by

and we assume that this accumulated payoff takes values in S; this is the case if, e.g., S = [0, ).

Remark 2.51. It may happen that an agent refuses the single favorable bet X at any level of wealth but feels tempted by a sufficiently large series X1, . . . , Xn of independent repetitions of the same bet. It is true that, by the weak law of large numbers, the probability

(for ε := m(μ)) of incurring a cumulative loss at the end of the series converges to 0 as n . Nevertheless, the decision of accepting n repetitions is not consistent with the decision to reject the single bet at any wealth level w. In fact, for Wk := w + Z k we obtain

i.e., the bet described by Z n should be rejected as well.

Let us denote by μn the distribution of the accumulated payoff Zn. The lottery μn has the mean m(μn) = n · m(μ), the certainty equivalent c(μn), and the associated risk premium ϱ(μn) = n · m(μ) c(μn). We are interested in the asymptotic behavior of these quantities for large n. Kolmogorovs law of large numbers states that the average outcome converges P-a.s. to the constant m(μ). Therefore, one might guess that a similar averaging effect occurs on the level of the relative certainty equivalents

and of the relative risk premiums

Does cn converge to m(μ), and is there a successive reduction of the relative risk premiums ϱn as n grows to infinity? Applying our heuristic (2.14) to the present situation yields

Thus, one should expect that ϱn tends to zero only if the ArrowPratt coefficient α(x) becomes arbitrarily small as x becomes large, i.e., if the utility function is decreasingly risk averse. This guess is confirmed by the following two examples.

Example 2.52. Suppose that u(x) = 1 eαx is a CARA utility function with constant risk aversion α > 0 and assume that μ is such that Then, with the notation introduced above,

Hence, the certainty equivalent of μn is given by

It follows that cn and ϱn are independent of n. In particular, the relative risk premiums are not reduced if the lottery is drawn more than once.

The second example displays a different behavior. It shows that for HARA utility functions the relative risk premiums will indeed decrease to 0. In particular, the lottery μn will become attractive for large enough n as soon as the price of the single lottery μ is less than m(μ).

Example 2.53. Suppose that μ is a nondegenerate lottery concentrated on (0, ), and that u is a HARA utility function of index γ [0, 1). If γ > 0 then and hence

If γ = 0 then u(x) = log x, and the relative certainty equivalent satisfies

Thus, we have

for any γ [0, 1). By symmetry,

see, e.g., part II of §20 in [20]. It follows that

Since u is strictly concave and since μ is nondegenerate, we get

i.e., the relative certainty equivalents are strictly increasing and the relative risk premiums ϱn are strictly decreasing. By Kolmogorovs law of large numbers,

Thus, by Fatous lemma (we assume for simplicity that μ is concentrated on [ε,) for some ε > 0 if γ = 0),

hence

Suppose that the price of μ is given by π c(μ), m(μ). At initial wealth w = 0, the agent would decline a single bet. But, in contrast to the situation in Remark 2.51, a series of n repetitions of the same bet would now become attractive for large enough n, since c(μn) = ncn > for

Remark 2.54. The identity (2.21) can also be written as

where An+1 = σ(Zn+1, Zn+2, . . . ). This means that the stochastic process n = 1,2. . . , is a backwards martingale, sometimes also called reversed martingale. In particular, Kolmogorovs law of large numbers (2.22) can be regarded as a special case of the convergence theorem for backwards martingales; see part II of §20 in [20].

Exercise 2.3.4. Investigate the asymptotics of cn in (2.20) for a HARA utility function with γ < 0.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.22.74.66