Pedantic Notes on the Risk Premium and Risk Aversion

Similar documents
Notes on the Risk Premium and Risk Aversion

Differentiating an Integral: Leibniz Rule

Calculus and Maximization I

Convex Analysis and Economic Theory AY Elementary properties of convex functions

Choice under Uncertainty

Division of the Humanities and Social Sciences. Supergradients. KC Border Fall 2001 v ::15.45

REAL ANALYSIS II: PROBLEM SET 2

Summary Notes on Maximization

Concavity and Lines. By Ng Tze Beng

Theory of Value Fall 2017

Recitation 7: Uncertainty. Xincheng Qiu

Proof. We indicate by α, β (finite or not) the end-points of I and call

Expectation is a positive linear operator

Mathematical Preliminaries for Microeconomics: Exercises

Technical Results on Regular Preferences and Demand

Department of Mathematics

2019 Spring MATH2060A Mathematical Analysis II 1

September Math Course: First Order Derivative

Convex Functions and Optimization

We consider the problem of finding a polynomial that interpolates a given set of values:

Problem List MATH 5143 Fall, 2013

u xx + u yy = 0. (5.1)

6 Lecture 6: More constructions with Huber rings

Classical regularity conditions

Tijmen Daniëls Universiteit van Amsterdam. Abstract

Convex Analysis and Economic Theory Winter 2018

arxiv: v1 [q-fin.mf] 25 Dec 2015

SPRING 2006 PRELIMINARY EXAMINATION SOLUTIONS

Laplace s Equation. Chapter Mean Value Formulas

Convex Analysis and Economic Theory Winter 2018

Topological properties

II. Analysis of Linear Programming Solutions

Weak* Axiom of Independence and the Non-Expected Utility Theory

MATH 1A, Complete Lecture Notes. Fedor Duzhin

Follow links for Class Use and other Permissions. For more information send to:

Notes on the Implicit Function Theorem

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

A New Class of Non Existence Examples for the Moral Hazard Problem

Convex Analysis and Economic Theory Winter 2018

THE INVERSE FUNCTION THEOREM

Moral hazard in teams

1 Uncertainty. These notes correspond to chapter 2 of Jehle and Reny.

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

Mathematical Economics. Lecture Notes (in extracts)

Course 212: Academic Year Section 1: Metric Spaces

Introduction to Topology

Non-Life Insurance: Mathematics and Statistics

Incremental Risk Vulnerability

3 Intertemporal Risk Aversion

Robust Comparative Statics of Non-monotone Shocks in Large Aggregative Games

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

Division of the Humanities and Social Sciences. Sums of sets, etc.

Introduction to Real Analysis Alternative Chapter 1

Infinite series, improper integrals, and Taylor series

Chapter 1 - Preference and choice

MAXIMA AND MINIMA CHAPTER 7.1 INTRODUCTION 7.2 CONCEPT OF LOCAL MAXIMA AND LOCAL MINIMA

The general programming problem is the nonlinear programming problem where a given function is maximized subject to a set of inequality constraints.

CONSUMER DEMAND. Consumer Demand

Review of Optimization Methods

Iowa State University. Instructor: Alex Roitershtein Summer Homework #5. Solutions

SOME MEASURABILITY AND CONTINUITY PROPERTIES OF ARBITRARY REAL FUNCTIONS

Existence Theory: Green s Functions

Existence and monotonicity of solutions to moral hazard problems

SOLUTIONS TO ADDITIONAL EXERCISES FOR II.1 AND II.2

Exercise Solutions to Functional Analysis

7.5 Partial Fractions and Integration

LECTURE 15: COMPLETENESS AND CONVEXITY

Existence and uniqueness: Picard s theorem

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

ARE202A, Fall Contents

Sequences and Series of Functions

Math 341: Convex Geometry. Xi Chen

Ordinary Differential Equation Introduction and Preliminaries

Connectedness. Proposition 2.2. The following are equivalent for a topological space (X, T ).

Competitive Equilibria in a Comonotone Market

Increases in Risk Aversion and Portfolio Choice in a Complete Market

2 Sequences, Continuity, and Limits

Appendix B Convex analysis

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

14.12 Game Theory Lecture Notes Theory of Choice

MATH The Chain Rule Fall 2016 A vector function of a vector variable is a function F: R n R m. In practice, if x 1, x n is the input,

MORE ON CONTINUOUS FUNCTIONS AND SETS

CHAPTER 2: CONVEX SETS AND CONCAVE FUNCTIONS. W. Erwin Diewert January 31, 2008.

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Chapter 4: Asymptotic Properties of the MLE

Set, functions and Euclidean space. Seungjin Han

Risk Aversion over Incomes and Risk Aversion over Commodities. By Juan E. Martinez-Legaz and John K.-H. Quah 1

Consumer theory Topics in consumer theory. Microeconomics. Joana Pais. Fall Joana Pais

x y More precisely, this equation means that given any ε > 0, there exists some δ > 0 such that

HOMEWORK ASSIGNMENT 6

Calculus 221 worksheet

FINANCIAL OPTIMIZATION

LECTURE 6. CONTINUOUS FUNCTIONS AND BASIC TOPOLOGICAL NOTIONS

NOTES ON CALCULUS OF VARIATIONS. September 13, 2012

Integral Jensen inequality

The Non-Existence of Representative Agents

The Process of Mathematical Proof

A SECOND ORDER STOCHASTIC DOMINANCE PORTFOLIO EFFICIENCY MEASURE

f(x) f(z) c x z > 0 1

1 Lyapunov theory of stability

Transcription:

Division of the Humanities and Social Sciences Pedantic Notes on the Risk Premium and Risk Aversion KC Border May 1996 1 Preliminaries The risk premium π u (w, Z) for an expected utility decision maker with Bernoulli utility function u is defined implicitly by the equation u ( w + E Z π u (w, Z) ) = E u(w + Z). ( ) It is the most that they would be willing to pay to completely insure against the risk Z to their initial wealth w. The first obvious question is, how do we know that such π exists, and if it does exist how do we know that it is unique? Clearly, in order for equation ( ) to make sense, we must place some restrictions on the random variable Z. Let us say that a random variable Z is admissible for u at w if w + Z takes on values in the domain of u almost surely, both E Z and E u(w + Z) are finite, and w + E Z belongs to the domain of u. If the domain is an interval, this latter requirement follows from the others. Let us say that (w, Z) is an admissible pair for u if Z is admissible for u at w. 1 Lemma Let D be an interval of the real line. Let u be a continuous strictly increasing real function on D. Suppose (w, Z) is an admissible pair for u. Then equation ( ) is satisfied by a unique value π u (w, Z). Proof : Since u is continuous, the Intermediate Value Theorem (e.g., [2, Theorem 3.8, p. 144]) implies that the range of u is an interval. Moreover, the range contains the number E u(w +Z). It then follows that there is some π so that equation ( ) is satisfied. If u is strictly increasing, then π must also be unique. So henceforth we shall require: 2 Assumption Assume that the utility function is continuous, strictly increasing, and defined on an interval of R. I wrote these notes to reproduce some of the results in Pratt s [9] paper in my own style, since I was never facile with Landau s o and O notation. It also bothered me that Pratt assumed that the utility functions are sufficiently regular to justify the proofs (p. 123, line 2 from bottom). I wish my referees would let me make assumptions like that. 1

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 2 Recall that a concave function is automatically continuous on the interior of its domain, so if u is concave, strictly increasing, and defined on an open interval, then π u (w, Z) is unique. A point worth noting is that subjecting the utility u to a positive affine transformation does not affect the validity of equation ( ), which is a good thing because it means that the risk premium is a property of the preference relation, not the particular Bernoulli utility function. Another point worth mentioning is that if Z is admissible for u at w, then Z E Z is admissible for u at w + E Z, and π u (w, Z) = π u (w + E Z, Z E Z), so without loss of generality we need only consider admissible pairs (w, Z) where E Z = 0. It is immediate that strictly increasing functions are one-to-one and so have inverses. The next easy lemma gives simple conditions for the continuity of the inverse. 3 Lemma Let u be a strictly increasing real function on the domain D in R. If D is either closed or an interval, then u 1 is continuous on the range of u. Proof : To see this, suppose u(x n ) u(x). We wish to show that x n x. By passing to a subsequence we may assume without loss of generality that either u(x n ) u(x) or u(x n ) u(x). Let s consider the case u(x n ) u(x). Since u is strictly increasing, the sequence x n is increasing and bounded above by x. It thus has a limit y satisfying x 1 y x. If D is either closed or an interval, then y D. Now if y < x, then u(x n ) u(y) < u(x) for all n, which contradicts u(x n ) u(x). Therefore x n y = x, which shows the continuity of u 1. The case of u(x n ) u(x) is disposed of similarly. 3 u 1 (x) 2 2 u(x) 1 1 ) 0 ) 1 2 3 x 0 1 2 x Figure 1. Discontinuity of the inverse.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 3 4 Example (Continuous and discontinuous inverses) Note that the result above does not require continuity of u, but it can fail if the domain D is neither closed nor an interval. For instance, set D = [0, 1) [2, 3] and define u by u(x) = x for x in [0, 1) and u(x) = x 1 for x in [2, 3]. Then u is continuous, strictly increasing, and its range is the interval [0, 2]. Nevertheless u 1 is not continuous, it jumps up at 1. See Figure 1. Now flip the picture over and set v = u 1. Then v is strictly increasing but not continuous, and its domain is the closed interval [0, 2]. Observe that its inverse u is continuous on the range [0, 1) [2, 3] of v. We shall be interested in comparing different utility functions u and v defined on the same domain D. The next lemma will be extremely useful. It says that we can write either utility as a strictly increasing function of the other. Recall that a function is of class C k on D if it is continuous on D and has a continuous k th -order derivative on the interior of D. 5 Lemma Let u and v be strictly increasing real functions defined on the common interval D. Then the function g defined on the range of v by is the unique function satisfying g(x) = u ( v 1 (x) ). u = g v. Furthermore g is strictly increasing. If in addition, u is continuous, then g is continuous too. If v is continuous, then the range of v, and hence the domain of g, is an interval. If u and v are both C k and v is never zero, then g is C k on its domain. Proof : Since v is strictly increasing, it has an inverse v 1, which is also strictly increasing, so g exists and is the unique function satisfying u = g v. Clearly g is strictly increasing, being the composition of two strictly increasing functions. Since v is strictly increasing, then by Lemma 3 v 1 is continuous on the range of v. So if u is continuous too, then g(x) = u ( v 1 (x) ) is continuous. If v is continuous, as we remarked above, the Intermediate Value Theorem implies that its range is an interval. If v is C k and v is never 0, the Inverse Function Theorem (see, e.g., [7, Theorem 1, p. 230]) implies that v 1 is also C k. Thus g as the composition of two C k functions is also C k. 2 Global results The main theorem in the area is due to Pratt [9]. The next few theorems refine the published results. 6 Theorem Let D be an interval of the real line. Let u and v be continuous strictly increasing real functions on D. Then the following statements are equivalent. (1) For all w in D and random variables Z that are admissible for u and v at w and satisfy E Z = 0, π u (w, Z) π v (w, Z).

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 4 (2) There exists a concave strictly increasing function g defined on the range of v satisfying u = g v. Proof : (1) = (2): Lemma 5 gives the existence of the strictly increasing function g = u v 1, so the crux is to prove that g is concave. That is, for v 1, v 2 in the range of v, g ( λv 1 +(1 λ)v 2 ) λg(v 1 ) + (1 λ)g(v 2 ) for 0 λ 1. Note that Lemma 5 implies that the domain of g is an interval and hence a convex set, so g is actually defined for each λv 1 + (1 λ)v 2. To this end, let w 1, w 2 in D satisfy and set v(w 1 ) = v 1 and v(w 2 ) = v 2, w = λw 1 + (1 λ)w 2. Define the random variable Z so that { (1 λ)(w1 w Z = 2 ) with probabilityλ λ(w 2 w 1 ) with probability1 λ. Then E Z = 0 and w + Z = { w1 with probabilityλ w 2 with probability1 λ. Consequently, Z is admissible for u and v at w, and and u ( w π u (w, Z) ) = E u(w + Z) = E g ( v(w + Z) ) = λg(v 1 ) + (1 λ)g(v 2 ) u ( w π v (w, Z) ) = g (v ( w π v (w, Z) )) = g ( E v(w + Z) ) = g ( λv 1 + (1 λ)v 2 ). But u is increasing and π u (x, Z) π v (x, Z), so u ( w π v (w, Z) ) u ( w π u (w, Z) ), so we conclude that g ( λv 1 + (1 λ)v 2 ) λg(v1 ) + (1 λ)g(v 2 ) which proves that g is indeed concave. (2) = (1): This is clearly a job for Jensen s inequality. Let Z be admissible for u and v at w and satisfy E Z = 0. Observe that u ( w π u (w, Z) ) = E u(w + Z) = E g ( v(w + Z) ) g ( E v(w + Z) ) = g (v ( w π v (w, Z) )) = u ( w π v (w, Z) ),

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 5 where the inequality follows from Jensen s inequality and the concavity of g, and the other equalities are either from the definition of the risk premium or the assumption that statement (2) holds. Since u is strictly increasing it follows that π u (w, Z) π v (w, Z). The above proof can be trivially modified to show the following result. 7 Theorem Let D be an interval of the real line. Let u and v be continuous strictly increasing real functions on D. Then the following statements are equivalent. 1. For all w in D and all nondegenerate random variables Z that are admissible for u and v at w and satisfy E Z = 0, π u (w, Z) > π v (w, Z). 2. There exists a strictly concave strictly increasing function g on the range of v satisfying u = g v. 3 Local results When u is twice differentiable, the (Arrow Pratt de Finetti) coefficient of (local) risk aversion 1 r u is defined by r u (w) = u (w) u (w). Note that this coefficient is invariant under positive affine transformations of u, so it really is a property of the preferences. 8 Theorem Let D be an interval of the real line. Let u and v be continuous strictly increasing functions on D that are twice differentiable with strictly positive derivatives everywhere on the interior of D. Then the following statements are equivalent. (1) For all w in D and all random variables Z that are admissible for u and v at w and satisfy E Z = 0, π u (w, Z) π v (w, Z). (2) There exists a concave strictly increasing function g defined on the range of v satisfying (3) For all w in the interior of D, u = g v. u (w) u (w) v (w) v (w). 1 This is usually referred to as the Arrow Pratt coefficient of risk aversion. Pratt [9, p. 123n] points out that the importance of this coefficient had been independently and earlier discovered by Arrow and by Schlaiffer, but he gives no references. This coefficient (or rather twice its inverse, 2u /u ) had been earlier discussed by de Finetti [4, p. 700] who essentially stated the relation used in the proof of (1) = (3) below. See also the review article by Montesano [8].

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 6 Proof : The equivalence of (1) and (2) has already been established. (1) = (3): Fix w in the interior of D. Then there is an interval A containing zero such that ε A implies w ± ε D. For each ε in A let Z ε be a random variable that takes on each of the values ε and ε with probability 1 2. Then E Z ε = 0 and Z ε is admissible for u and v at w. To simplify notation, define the real function p u on A by p u (ε) = π u (w, Z ε ), and similarly for p v. Note that p u (0) = 0, p u (ε) = p u ( ε), and by equation ( ), u ( w p u (ε) ) = 1 2 u(w + ε) + 1 2u(w ε) (1) for all ε in A. Defining p v in an analogous manner, we note that p u p v on A. Note that (1) implies that the function p u is twice differentiable on A. To see this, define the function f : A (D w) R by f(ε, p) = u ( w p ) 1 2 u(w + ε) 1 2u(w ε) (2) f(0, 0) and note that f is twice differentiable, f(0, 0) = 0, and = u (w) < 0. Therefore p by the Implicit Function Theorem (see, e.g., [7, Theorem 2, p. 235]) there is a unique twice differentiable function defined on a neighborhood of zero giving p as a function of ε to satisfy equation (2). 2 This function must be our p u. (Naturally the same is true of p v too.) For the time being we shall drop the u subscript and simply write p for p u. Since (1) holds everywhere in A, we may differentiate both sides to get As a special case, Differentiating a second time yields u ( w p(ε) ) p (ε) = 1 2 u (w + ε) 1 2 u (w ε). p (0) = 0. u ( w p(ε) )( p (ε) ) 2 p (ε)u ( w p(ε) ) = 1 2 u (w + ε) + 1 2 u (w ε). Using p(0) = p (0) = 0 we have p (0) = u (w) u (w) = r u(w). (So that s where that expression comes from.) Having shown earlier that p is twice differentiable on A, for ε > 0 we can apply Taylor s Theorem [5, p. 290] to write p(ε) = p(0) + εp (0) + ε2 ( p (0) + R(ε) ) 2 = ε2 ( p (0) + R(ε) ), (3) 2 2 This conclusion may look a bit different from the statement of the theorem, which says that if u is k times continuously differentiable then p is k times continuously differentiable. What we have assumed is continuous once differentiability and everywhere twice differentiability (not necessarily continuously). But the formula for the derivative of p that is given below can be deduced from the continuous differentiability. But clearly the derivative is differentiable.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 7 where lim ε 0 R(ε) = 0. 3 Now we have to start reattaching subscript us and vs. Using equation (3) and p u p v, we conclude p u(0) + R u (ε) p v(0) + R v (ε) for all ε > 0 in A. Letting ε 0, we conclude that p u(0) p v(0). That is, u (w) u (w) v (w) v (w). Since w is arbitrary, this completes the proof of the implication. (3) = (2): As above, Lemma 5 gives the existence of strictly increasing continuous g, so to prove statement (2) we need to show that g is concave. It follows from Lemma 5 that g is twice differentiable on the interior of its domain, so it suffices to prove that g 0. Since u(w) = g ( v(w) ) for all w, we may differentiate both sides to conclude I should put a reference here. and u (w) = g ( v(w) ) v (w) (4) u (w) = g ( v(w) )( v (w) ) 2 + g ( v(w) ) v (w). (5) Dividing (5) by (4) yields u ( ) (w) u (w) = g v(w) g ( v(w) ) v (w) + v (w) v (w). Thus r u r v implies g ( v(w) ) g ( v(w) ) v (w) 0, but u > 0 and v > 0 imply g ( v(w) ) ( ) = u v(w) v > 0, (w) so g ( v(w) ) 0. Thus g is concave. 9 Remark (More nitpicking) Note that if D is open and u is concave and strictly increasing, then the assumption that u > 0 is redundant, since 0 u(w) implies that w is a global maximum of u, which cannot occur (since D is open and u is strictly increasing). Again we may strengthen some of the inequalities. 10 Theorem Let D be an interval of the real line. Let u and v be continuous strictly increasing functions on D that are twice differentiable with strictly positive derivatives everywhere on the interior of D. Consider following statements. (1) For all w in D and all nondegenerate random variables Z that are admissible for u and v at w and satisfy E Z = 0, π u (w, Z) > π v (w, Z). 3 This form of Taylor s Theorem, given by Hardy [5], requires only twice differentiability at 0, not twice continuous differentiability on a neighborhood of 0.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 8 (2) There exists a strictly concave strictly increasing function g on the range of v satisfying u = g v. (3) For all w in the interior of D, u (w) u (w) > v (w) v (w). Then (3) = (2) = (1). I do not claim that (1) = (3), since adapting the arguments in the proof of Theorem 8 only yield a weak inequality. Pratt himself makes the slightly weaker claim that (1) implies the following: (3 ): For all w in the interior of D, u (w) u (w) v (w) v (w), and for every subinterval of D there is some w with u (w) u (w) v (w) v (w). Varian [12, pp. 182 183] makes the stronger claim, so I need to do some work here. 4 A more general interpretation As an aside we mention another interpretation of (3). The variance of Z ε is ε 2. So pu(ε) is the ε 2 fraction of the variance that someone with utility u would be willing to pay to insure against Z ε. The limit of this fraction as ε 0 is then 1 2 r u(w). In fact, this generalizes to more general admissible random variables with variance ε > 0, but the only proof I can think of is subtle. 11 Theorem Let D be an interval of the real line. Let u and v be continuous strictly increasing functions on D that are twice differentiable with strictly positive derivatives everywhere on the interior of D. Fix w in the interior of D, and let η > 0 satisfy (w η, w+η) D. For 0 < ε < η, let X ε be an admissible random variable for u at w satisfying E X ε = 0 and E Xε 2 = ε 2. Assume in addition that X ε 0. Then ε 0 π u (w, X ε ) lim ε 0 ε 2 = 1 2 r u(w). Proof : To simplify notation, set p(ε) = π u (w, X ε ). Then by definition, u ( w p(ε) ) = u ( w + X ε (s) ) dp (s). (6) S

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 9 Using Taylor s Theorem write the left-hand side of (6) as u ( w p(ε) ) = u(w) u ( w δ(ε) ) p(ε) where for each ε, δ(ε) [0, p(ε)]. Now us the stochastic version of Taylor s Theorem (Theorem 12 below) to write the right-hand side of (6) as u ( w + X ε (s) ) dp (s) S { = u(w) + u (w)x ε (s) + 1 2 u ( w + ξ ε (s) ) } Xε 2 (s) dp (s), S where each ξ ε is a measurable function and for each s, ξ ε (s) [0, X ε (s)]. Dividing both terms by ε 2 and simplifying (using E X ε = 0), we can rewrite (6) as Now the left-hand side satisfies u ( w δ(ε) )p(ε) ε 2 = 1 u ( w + ξ ε (s) )X2 ε (s) 2 S ε 2 dp (s). u ( w δ(ε) )p(ε) ε 2 ε 0 ( u (w) lim ε 0 ) p(ε) ε 2. Since u is continuous, and ξ ε X ε, for each η > 0 there is some ε 0 < 0 such ε 0 that for all ε < ε 0 we have u ( w + ξ ε (s) ) u (w) < η for almost every s. So for such ε, since S X 2 ε (s) ε 2 Combining yields S u ( w + ξ ε (s) )X2 ε (s) ε 2 dp (s) = 1. Therefore S u ( w + ξ ε (s) )X2 ε (s) ε 2 dp (s) u (w) X2 ε (s) S ε 2 p(ε) lim ε 0 ε 2 = 1 u (w) 2 u (w) = 1 2 r u(w). dp (s) < η, dp (s) ε 0 u (w). (7) The above proof of Theorem 11 relies on the following stochastic version of Taylor s Theorem. 12 Stochastic Taylor s Theorem Let h: [a, b] R be continuous and possess a continuous n th -order derivative on (a, b). Fix c [a, b] and let X be a random variable on the probability space (S, S, P ) such that c + X [a, b] almost surely. Then there is a (measurable) random variable ξ satisfying ξ(s) [0, X(s)] for all s (where [0, X(s)] is the line segment joining 0 and X(s), regardless of the sign of X(s)), and h ( c + X(s) ) n 1 1 = h(c) + k! h(k) (c)x k (s) + 1 n! h(n)( c + ξ(s) ) X n (s). k=1

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 10 Proof : Taylor s Theorem asserts that there is such a ξ(s) for each s, the trick is to show that there is a measurable version. To this end define the correspondence φ: S R by φ(s) = [0, X(s)]. It follows from [1, Theorem 18.5, p. 595] that φ is measurable and it clearly has compact values. Set g(s) = h ( c+x(s) ) h(c) n 1 k=1 1 k! h(k) (c)x k (s), f(s, x) = 1 n! h(n)( c+x ) X n (s). Then g is measurable and f is a Carathéodory function. (See sections 4.10 and 18.1 of [1] for the definitions of Carathéodory functions and measurable correspondences.) By Filippov s Implicit Function Theorem [1, Theorem 18.17, p. 603] there is a measurable function ξ such that for all s, ξ(s) φ(s) and f ( s, ξ(s) ) = g(s), and we are done. 5 Some odd examples 13 Example To see the need for the assumption X ε 0 in Theorem 11, consider the following simple example, which shows how (7) can fail. Let 1 with probability ε ε+1 X ε = ε with probability 1 ε+1 Then EX ε = 0 and EXε 2 = ε, so X ε 2 0, but X ε = 1. The risk premium p(ε) for X ε is defined by u ( (w p(ε) ) = ε 1 u(w + 1) + u(w ε). (8) ε + 1 ε + 1 Proceeding more along the lines of Pratt, we use Taylor s Theorem to write u(w p) = u(w) p ( u (w) + R 1 (p) ) and u(w ε) = u(w) εu (w) + 1 2 ε2( u (w) + R 2 (ε) ), and clearly Therefore (8) can be written u(w) p(ε) ( u ( )) (w) + R 1 p(ε) ( = u(w) + ε ε + 1 or regrouping, u(w + 1) = u(w) + u(w + 1) u(w) u u (w). (w) ) u(w + 1) u(w) u u (w) + 1 ( u(w) εu (w) + 1 (w) ε + 1 2 ε2( u (w) + R 2 (ε) )). p(ε) ( u ( )) (w) + R 1 p(ε) ( ε u(w + 1) u(w) = ε + 1 u (w) ) 1 u (w) + 1 2 ε2( u (w) + R 2 (ε) ).

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 11 Dividing by ε ( u (w) + R 1 ( p(ε) )) gives p(ε) = p(ε) var X ε ε = 1 ( ) u(w + 1) u(w) ε + 1 u 1 (w) u (w) u ( ) 1 ε u (w) + R 2 (ε) (w) + R 1 p(ε) 2 ε + 1 u ( ) (w) + R 1 p(ε) 1 u(w + 1) u(w) u (w) asε 0. This is not in general half the coefficient of risk aversion. If u is concave, then the supergradient inequality implies 1 u(w+1) u(w) u (w) > 0. To be more concrete, consider the quadratic utility u(w) = (a w) 2 on the interval [0, a]. Then u (w) = 2(a w) and u (w) = 2, so the risk aversion coefficient is r u (w) = (a w) which is positive on [0, a]. Fix w and let b = a w. Then (8) becomes (b + p) 2 = ε ε + 1 (b 1)2 1 (b + ε)2 ε + 1 Expanding and simplifying yields p 2 + 2bp ε = 0, so p(ε) = b 2 + ε b. Now p(ε)/ε p (0) = 1/2b as ε 0, whereas 1 2 r u(a b) = b/2. Figure 2 shows plots of p and p/ε 2 for b = 2 (i.e., w = a 2). 14 Example For a more dramatic example, let { 1/ε with probability ε 3 /(1 + ε 3 ) X ε = ε 2 with probability 1/(1 + ε 3 ) Then EX ε = 0 and EXε 2 = ε, so X ε 2 0, but X ε = 1/ε. The risk premium p(ε) for X ε is defined by u ( (w p(ε) ) = ε3 1 + ε 3 u(w + 1 ε ) + 1 1 + ε 3 u(w ε2 ). (9) By Taylor s Theorem, u(w p) = u(w) p ( u (w) + R 1 (p) ), u(w ε 2 ) = u(w) ε 2 u (w) + 1 2 ε4( u (w) + R 2 (ε 2 ) ),

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 12 0.2 0.15 0.1 0.05 0.25 0.248 0.246 0.244 0.242 0.238 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 Risk premium p(ε). 0.236 p(ε)/var X ε Figure 2. The risk premium p(ε) for the quadratic utility u(w) = (a w) 2 facing the risk X ε of Example 13 at wealth level w = a 2. (So 1 2r(w) = 1.) and also Therefore (9) can be written u(w) p(ε) ( u (w) + R 1 ( p(ε) )) or regrouping, u(w + 1/ε) = u(w) + p(ε) ( u (w) + R 1 ( p(ε) )) Dividing by ε ( u (w) + R 1 ( p(ε) )) gives Thus p(ε) = p(ε) var X ε ε = ε2 1 + ε 3 = = u(w + 1/ε) u(w) u u (w). (w) ε 3 ( 1 + ε 3 + 1 1 + ε 3 u(w) + ) u(w + 1/ε) u(w) u u (w) (w) ( u(w) ε 2 u (w) + 1 2 ε4( u (w) + R 2 (ε 2 ) )), ε 3 ( u(w + 1/ε) u(w) 1 + ε 3 u (w) + 1 ( 1 1 + ε 3 2 ε4( u (w) + R 2 (ε 2 ) )) ( u(w + 1/ε) u(w) u 1 ) u (w) (w) ε u ( ) 1 (w) + R 1 p(ε) 2 lim p(ε)/ε = 1 ε 0 u (w) lim ε 0 ε2 u(w + 1/ε). ε 3 1 ε ) u (w) 1 + ε 3 u (w) + R 2 (ε) u (w) + R 1 ( p(ε) ). If u is bounded, then this limit is zero. If u is unbounded but concave, then by the supergradient inequality, u(w) + u (w)/ε u(w + 1/ε), so lim ε 0 ε2 u(w + 1/ε) lim ε 2( u(w) + u (w)/ε ) = 0. ε 0

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 13 This makes sense. The risk premium will always be smaller than the maximum loss, and the maximum loss divided by the variance is ε, which tends to zero. 6 Recovering the utility We can recover the utility function from the coefficient of risk aversion. 15 Theorem Let r : D R be a continuous function on an open interval D. Then there is a C 2 utility u: D R satisfying u (w) > 0 for all w D and u (w) u (w) = r(w) for all w D. The function u is unique up to positive affine transformation and is given by where a D, α > 0, and β is an arbitrary real number. u(w) = α w a R(x) = e R(x) dx + β x a r(t) dt, Proof : Observe that we are seeking a solution to the linear differential equation u (x) r(x)u (x) = 0. Letting f(x) denote u (x), this becomes the first order equation f (x) + r(x)f(x) = 0. It is well known, see Theorem 17 below, that for each a D and α R, this equation has a unique solution satisfying f(a) = α. It is given by f(x) = αe R(x), where R(x) = x a r(t) dt. (In Theorem 17, set P = r and Q = 0.) Any such solution is continuous, so by the same result (now with P = 0 and Q = f), the differential equation has a solution u given by u(x) = u (x) = f(x) = α x a x a f(t) dt + β e R(t) dt + β (as P = 0 implies A = 0 in Theorem 17). To guarantee that u (x) > 0, we need only take α > 0, as e R(x) > 0. Thus within the class of strictly increasing functions, given a, the utility u is unique up to the constants α > 0 and β.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 14 16 Example Suppose the risk aversion function is constant: r(x) = c for all x. Then assuming a = 0 is in the domain, we have R(x) = cx, so x u(x) = α e ct dt + β. 0 If c = 0, then x 0 e 0t dt = x 0 1 dt = x u(x) = αx + β. If c 0, then R(x) = cx, and the primitive of e cx = 1 c e cx, so x 0 e ct dt = 1 c e cx ( 1 c e c0 ) = 1 c ( e cx 1 ), so our formula gives u(x) = α ( e cx 1 ) + β. c We can choose particularly aesthetic values of α and β. For c = 0, choose α = 1 and β = 0, to get u(x) = x (c = 0). For c > 0, set β = 0 and α = c > 0 to get For c < 0, set β = 0 and α = c > 0, to get u(x) = 1 e cx (c > 0). u(x) = e cx 1 (c < 0). The following theorem is taken from Apostol [2, Theorem 8.3, p. 310]. See my notes [3] for an economic derivation and interpretation. 17 Theorem (First order linear differential equations) Let P and Q be continuous functions on an open interval D. Let a D and b R. Then there is exactly one function f satisfying f(a) = b and f (x) + P (x)f(x) = Q(x). It is given by where x f(x) = be A(x) + e A(x) Q(x)e A(x) dx, A(x) = x a z P (x) dx.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 15 7 Ross generalization Suppose initial wealth is random, or equivalently, you can only partially insure your wealth so there is still uncertainty after insurance. That is, after insurance the wealth is a random variable w, but without insurance it is w + z. It is natural to implicitly define the generalized risk premium ˆπ u ( w, z) by the equation E u ( w + E z ˆπ u ( w, z) ) = E u( w + z). ( ) But there are problems with this definition that are not present in the Pratt definition for the case of full insurance. One obvious problem is that if the domain of u is bounded below, for instance, if it is R ++, and if the support of w includes the lower bound then for any positive ˆπ, the quantity w ˆπ lies outside the domain of u with positive probability, so ( ) may have no solution. A more subtle problem is that even for concave u, the risk premium ˆπ may actually be negative. That is, z may already provide some insurance against fluctuations in w. A simple extreme example is where w assumes the values x ± ε equally likely and z = x w, so that w + z = x. Then ˆπ( w, z) < 0 when u is concave. The problem here is that in this case z does not represent a risk. In order to be classified as a risk, we want w + z to be riskier than z in the sense of Rothschild and Stiglitz [11]. This amounts to the criterion that E( z w) = 0. With these caveats in mind, let us say that a pair ( w, z) of random variables is admissible for u if w and w + z take on values in the domain of u almost surely, E( z w) = 0 (which implies E z = 0), and E w, E u( w), and E u( w + z) are finite, and ( ) admits a solution ˆπ. Let us say that the utility function u is strongly more risk averse or more risk averse in the sense of Ross than v if u and v have a common domain and ˆπ u ( w, z) ˆπ u ( w, z) for all admissible pairs ( w, z). Ross [10] provides the following characterization of strong risk aversion, similar to Pratt s theorem. 18 Theorem (Ross) For u, v twice differentiable with u, v > 0 and u, v < 0, (strictly increasing and strictly concave), defined on a common nondegenerate interval D, the following are equivalent. 1. There exists λ > 0 and a decreasing concave function H satisfying H 0, H 0 such that u = λv + H. 2. For all w, z such that E( z w) = 0, ˆπ u ( w, z) ˆπ v ( w, z). That is, u is strongly more risk averse than v. 3. There exists λ > 0 such that for all x, y in the interior of D, u (x) v (x) λ u (y) v (y).

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 16 19 Remark There is an importance difference between this result and Pratt s result. If g is strictly concave and strictly increasing, then g v is always strictly more risk averse in Pratt s sense than v. Ross theorem does not assert that if λ > 0 and H is a decreasing concave function, then λv + H is strongly more risk averse than v. Indeed, it may be that λv + H is decreasing. The theorem does say that if λv + H is strictly increasing and concave, then λv + H is strongly more risk averse than v. The Ross ordering thus admits maximal elements, which the Pratt ordering does not. In fact, if v (x) 0 as x, then v is maximally strongly risk averse. To see this, suppose u = λv + H is strictly increasing, where H is concave and nonincreasing. It follows that H is a constant. That is, u is a positive affine transformation of v, so that u and v are equivalent Bernoulli utilities. Proof of Theorem 18: (1) = (2): This part uses Jensen s inequality. E u( w ˆπ u ) = E u( w + z) (by ( )) = E λv( w + z) + E H( w + z) (by hypothesis) = E λv( w + z) + E [ E ( H( w + z) w )] E λv( w + z) + E H ( E( w + z w) ) (Jensen s inequality) = E λv( w + z) + E H( w) (since E( z w) = 0) = E λv( w ˆπ v ) + E H( w) (definition of ˆπ v ) E λv( w ˆπ v ) + E H( w ˆπ v ) (as H is decreasing) = E u( w ˆπ v ) (by ( )) Since u is increasing ˆπ u ˆπ v. (2) = (3): Let 1 > p > 0 and consider a random wealth w with the following distribution y with probability 1 p w = x with probability p and a risk z with the following distribution. 0 if w = y z = ε with probability 1 2 ε with probability 1 2 if w = x if w = x. Then E ( z w) = 0 and y with probability 1 p w + z = x + ε with probability p 2 x ε with probability p 2. Then ( ) becomes pu ( ) ( ) x ˆπ u + (1 p)u y ˆπu = 1 2 pu(x + ε) + 1 2pu(x ε) + (1 p)u(y). (10) This implicitly defines ˆπ u as a function of ε and p. Let us write π(ε, p) for ˆπ u ( w, z). Note that π(0, p) = 0

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 17 for any p. By the Implicit Function Theorem, since u is C 2, and u > 0, we see that π is C 2. We can use (10) to compute the partial derivatives π 2 π ε and. ε 2 Differentiating (10) with respect to ε yields ( pu (x π) + (1 p)u (y π) ) π ε = p ( u (x + ε) u (x ε) ). (11) 2 Evaluating this at ε = 0, since the right-hand side is zero and u > 0, we have π(0, p) ε Differentiating (11) with respect to ε yields ( pu (x π) + (1 p)u (y π) ) ( π ε ) 2 = 0. ( pu (x π) + (1 p)u (y π) ) 2 π ε 2 = 1 2 p( u (x + ε) + u (x ε) ). Evaluating at ε = 0, we have 2 π(0, p) ε 2 By Taylor s theorem, for ε 0, π(ε, p) = π(0, p) + }{{} =0 pu (x) = pu (x) + (1 p)u (y). (12) π(0, p) ε } {{ } =0 ε + 1 2 π ( δ(ε), p ) 2 ε 2 ε 2, (13) where δ(ε) is strictly between 0 and ε. Reattaching subscripts u and v in (13) we see that π u (ε, p) π v (ε, p) implies 2 π ( u δu (ε), p ) ε 2 2 π ( v δv (ε), p ) ε 2. Since each π is a twice differentiable function of ε, letting ε 0 yields 2 π ( ) u 0, p ε 2 2 π ( ) v 0, p ε 2. So by (12) pu (x) pu (x) + (1 p)u (y) pv (x) pv (x) + (1 p)v (y). Multiplying both sides by the positive quantity pu (x) + (1 p)u (y) pv (x) u (x) v (x) pu (x) + (1 p)u (y) pv (x) + (1 p)v (y). gives Since x and y are arbitrary, letting p 0 gives the desired conclusion. (3) = (1): Set H = u λv, where λ is given by statement (3). Then H = u λv, but λ u v so u λv 0. Therefore H 0. Now H = u λv. But also by statement (3) u v λ so u λv as v 0. Therefore H 0.

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 18 It follows from this theorem that even if u is more risk averse than v in the sense of Arrow and Pratt that u need not be strongly more risk averse than v. The following example due to Ross [10] drives this point home. 20 Example Consider the case where u and v have constant absolute risk aversion, and u is more risk averse than v. That is, For this case, (13) implies u(x) = e ax and v(x) = e bx with a > b > 0. lim lim p 0 ε 0 π u π v = u (y) u (x) v (y) v (x) = a b e (a b)(x y). Thus, if ε and p are sufficiently small, and x y is large enough, then π u (ε, p) < π v (ε, p) even though u is more risk averse in Pratt s sense. References [1] C. D. Aliprantis and K. C. Border. 2006. Infinite dimensional analysis: A hitchhiker s guide, 3d. ed. Berlin: Springer Verlag. [2] T. M. Apostol. 1967. Calculus, 2d. ed., volume 1. Waltham, Massachusetts: Blaisdell. [3] K. C. Border. 2016. The economics of first-order linear differential equations. http://people.hss.caltech.edu/~kcb/notes/diffeq.pdf [4] B. de Finetti. 1952. Sulla preferibilità. Giornale degli Economisti e Annali di Economia 11(11/12):685 709. http://www.jstor.org/stable/23236169 [5] G. H. Hardy. 1952. A course of pure mathematics, 10th. ed. Cambridge: Cambridge University Press. [6] M. J. Machina and W. S. Neilson. 1987. The Ross characterization of risk aversion: Strengthening and extension. Econometrica 55(5):1139 1149. http://www.jstor.org/stable/1911264 [7] J. E. Marsden. 1974. Elementary classical analysis. San Francisco: W. H. Freeman and Company. [8] A. Montesano. 2009. De Finetti and the Arrow Pratt measure of risk aversion. In M. C. Galavotti, ed., Bruno de Finetti Radical Probabilist, Texts in Philosophy, chapter 7, pages 115 127. College Publications. http://works.bepress.com/aldo_montesano/35/ [9] J. W. Pratt. 1964. Risk aversion in the small and in the large. Econometrica 32(1/2):122 136. http://www.jstor.org/stable/1913738 [10] S. A. Ross. 1981. Some stronger measures of risk aversion in the small and the large with applications. Econometrica 49(3):621 638. http://www.jstor.org/stable/1911515

KC Border Pedantic Notes on the Risk Premium and Risk Aversion 19 [11] M. Rothschild and J. E. Stiglitz. 1970. Increasing risk I: A definition. Journal of Economic Theory 2:225 243. http://cowles.econ.yale.edu/p/cp/p03a/p0341a.pdf [12] H. R. Varian. 1992. Microeconomic analysis, 3d. ed. New York: W. W. Norton & Co.