Microeconomics Leopold S¨ogner c Department of Economics and Finance Institute for Advanced Studies Josefst¨adter Straße 39 1080 Vienna soegner@ihs.ac.at October, 2015 Course Outline (1) Microeconomics Learning Objectives: • This course covers key concepts of microeconomic theory. The main goal of this course is to provide students with both, a basic understanding and analytical traceability of these concepts. • The main concepts are discussed in detail during the lectures. In addition students have to work through the textbooks and have to solve problems to improve their understanding and to acquire skills to apply these tools to related problems. 1 Course Outline (2) Microeconomics Literature: • Andreu Mas-Colell, A., Whinston, M.D., Green, J.R., Microeconomic Theory, Oxford University Press, 1995. (MWG in the following) • Gravelle, H. and Rees, R., Microeconomics, 3rd edition, Prentice Hall, 2004. (GR in the following) Supplementary Literature: • Gilboa, I., Theory of Decision under Uncertainty, Cambridge University Press, 2009. • Gollier C., The Economics of Risk and Time, Mit Press, 2004. • Jehle G.A. and P. J. Reny, Advanced Microeconomic Theory, Addison-Wesley Series in Economics, Longman, Amsterdam, 2000. • Ritzberger, K., Foundations of Non-Cooperative Game Theory, Oxford University Press, 2002. 2 Course Outline (3) Microeconomics • The slides you find in the university information system contain (most of) the topics covered in MWG Chapters 1-6, 10 and 15 and 16. • For the slides I’m grateful to comments by my colleges Egbert Dierker and Martin Meier as well as to all former students in Bruno and Vienna. • For the course in Brno I’m grateful for comments from and discussions with Rostislav Stanˇek, Josef Menˇsik, by Ondˇrej Krˇc´al, and many others. 3 Course Outline (4) Microeconomics 1. Decision theory and the theory of the consumer: – Rationality, preference primitives and axioms, preference representations and utility (MWG 1-3, GR 2A-B). – Utility maximization, Walrasian demand and comparative statics (MWG, 2, 3D, GR 2C-D). – Indirect utility, expenditure function, Hicksian demand (MWG 3E,G, GR 3A). – Slutsky equation, substitution and wealth effect (MWG 3 G, GR 3B). 4 Course Outline (5) Microeconomics 2. Production and cost: – Production functions, returns to scale (GR 5). – Production set (MWG 5 B). – Cost minimization, conditional factor demands, cost function (MWG 5 C, GR 6.A,B,E). – Profit maximization, input demands, profit function, objectives of the firm (MWG 5 C,G, GR 7.A,C,D). 5 Course Outline (6) Microeconomics 3. General Equilibrium: – Introduction, Walrasian equilibrium (MWG 15, GR 12.A-D). – The Edgeworth box (MWG 15B, GR 12.E). – Welfare theorems (MWG16 A-D, GR 13). 4. Decisions under uncertainty: – Expected utility theorem, risk aversion (MWG 6A-C,GR 17 A-D). 6 Course Outline (7) Microeconomics • Winter Term 2015 (see university information system) – First block: October 22-23; 2015. – Second block: November 19-20, 2015. – Third block: December 17-18, 2015. • Contact hours (per semester): 12 units a 90 minutes. • Practice session will be organized by Rostislav Stanˇek. 7 Course Outline (8) Microeconomics Some more comments on homework and grading: • Final test (80%). • Homework and practice session (20%). • Final test and retakes: tba 8 Consumer Theory (1) Rationality and Preferences Microeconomics • Consumption set X • Rationality • Preference relations and utility • Choice correspondences and the weak axiom of revealed preference • Relationship between the axiomatic approach and the revealed preference approach MWG, Chapter 1. 9 Consumer Theory (1) Rationality (1) Microeconomics • We consider agents/individuals and goods that are available for purchase in the market. • Definition: The set X of all possible mutually exclusive alternatives (complete consumption plans) is called consumption set or choice set. • ”Simplest form of a consumption set”: We assume that each good, xl ∈ X, l = 1, . . . , L can be consumed in infinitely divisible units, i.e. xl ∈ R+. With L goods we get the commodity vector x in the commodity space RL +. 10 Consumer Theory (1) Rationality (2) Microeconomics • Approach I: describe behavior by means of preference relations; preference relation is the primitive characteristic of the individual. • Approach II: the choice behavior is the primitive behavior of an individual. 11 Consumer Theory 1 Rationality (3) Microeconomics • Consider the binary relation “at least good as“, abbreviated by the symbol . • For x, y ∈ X, x y implies that from a particular consumer’s point of view x is preferred to y or that he/she is indifferent between consuming x and y. • From we derive the strict preference relation : x y if x y but not y x and the indifference relation ∼ where x y and y x. 12 Consumer Theory 1 Rationality (4) Microeconomics • Often we require that pair-wise comparisons of consumption bundles are possible for all elements of X. • Completeness: For all x, y ∈ X either x y, y x or both. • Transitivity: For the elements x, y, z ∈ X: If x y and y z, then x z • Definition[D 1.B.1]: The preference relation is called rational if it is complete and transitive. • Remark: Reflexive x x follows from completeness [D 1.B.1]. 13 Consumer Theory 1 Rationality (5) Microeconomics • Based on these remarks it follows that: Proposition [P 1.B.1]: If is rational then, – is transitive and irreflexive. – ∼ is transitive, reflexive and symmetric. – If x y z then x z. 14 Consumer Theory 1 Utility (1) Microeconomics • Definition : A function X → R is a utility function representing if for all x, y ∈ X x y ⇔ u(x) ≥ u(y). [D 1.B.2] • Does the assumption of a rational consumer imply that the preferences can be represented by means of a utility function and vice versa? • Theorem: If there is a utility function representing , then must be complete and transitive. [P 1.B.2] • The other direction requires more assumptions on the preferences - this comes later & in Micro II! 15 Consumer Theory 1 Choice Rules (1) Microeconomics • Now we follow the behavioral approach (vs. axiomatic approach), where choice behavior is represented by means of a choice structure. • Definition - Choice structure: A choice structure (B, C(.)) consists: (i) of a family of nonempty subsets of X; with B ⊂ X. The elements of B are the budget sets B. and (ii) a choice rule C(.) that assigns a nonempty set of chosen elements C(B) ∈ B for every B ∈ B. • B is an exhaustive listing of all choice experiments that a restricted situation can pose on the decision maker. It need not include all subsets of X. 16 Consumer Theory 1 Choice Rules (2) Microeconomics • The choice rule assigns a set C(B) to every element B, i.e. it is a correspondence. • Example [1.C.1,(i)]: X = {x, y, z}. B = {B1, B2} with B1 = {x, y} and B2 = {x, y, z}. C(B1) = {x} and C(B2) = {x}. • Example [1.C.1,(ii)]: X = {x, y, z}. B = {B1, B2} with B1 = {x, y} and B2 = {x, y, z}. C(B1) = {x} and C(B2) = {x, y}. • Put structure on choice rule - weak axiom of revealed preference. 17 Consumer Theory 1 Choice Rules (3) Microeconomics • Definition - Weak Axiom of Revealed Preference: A choice structure (B, C(.)) satisfies the weak axiom of revealed preference if for some B ∈ B with x, y ∈ B we have x ∈ C(B), then for any B ∈ B with x, y ∈ B and y ∈ C(B ) , we must also have x ∈ C(B ). [D 1.C.1] • I.e. if we have a budget set where both x and y are available and x is chosen, then there cannot be a budget set containing both bundles, for which y is chosen and x is not. • Example: C({x, y}) = x, then C({x, y, z}) = y is not possible. 18 Consumer Theory 1 Choice Rules (4) Microeconomics • Definition - Revealed Preference Relation ∗ based on the Weak Axiom: Given a choice structure (B, C(.)), ∗ is defined by: x ∗ y ⇔ there is some B ∈ B such that x, y ∈ B and x ∈ C(B). [D 1.C.2] • In other words: If x is revealed at least as good as y, then y cannot be revealed preferred to x. • Example [1.C.1,(i)] satisfies the weak axiom, Example [1.C.1,(ii)] does not. 19 Consumer Theory 1 Choice Rules (5) Microeconomics • If a decision maker has a rational preference ordering , does the choice structure satisfy the weak axiom of reveal preferences? • Given a choice structure satisfying the weak axiom, does this result in a rational preference relation? • Definition: Choice structure implied by a rational preference relation : C∗ (B, ) = {x ∈ B|x y for every y ∈ B}. Assume that C∗ (B, ) is nonempty. • Proposition: If is a rational preference relation, then the choice structure C∗ (B, ) satisfies the weak axiom. [P 1.D.1] 20 Consumer Theory 1 Choice Rules (6) Microeconomics Proof: • Suppose that x, y in some B ∈ B and x ∈ C∗ (B, ), then (i) x y. • We have to show that if x, y are in some B and y ∈ C∗ (B , ), then also x ∈ C∗ (B , ). • Suppose that x, y in B ∈ B and y ∈ C∗ (B , ), then (ii) y z for all z ∈ B . • The rational preference relation is transitive. Hence (i), (ii) and transitivity imply that x y z. Therefore x ∈ C∗ (B, ) by the definition of this set. 21 Consumer Theory 1 Choice Rules (7) Microeconomics • Other direction. We want to know whether the choice rule C(.) satisfying the weak axiom is equal to C∗ (., ) for some rational preference relation . • Definition: Given a choice rule C(.). rationalizes the choice rule if C(B) = C∗ (B, ) for all B ∈ B. [D 1.D.1] • Proposition: If a choice structure (B, C(.)) satisfies the weak axiom of revealed preference and B includes all subsets of X of up to three elements, then there is a rational preference relation that rationalizes C(.) relative to B. That is, C(B) = C∗ (B, ) for all B ∈ B. The rational preference relation is the unique preference relation that does so. [P 1.D.2] 22 Consumer Theory 1 Choice Rules (8) Microeconomics • Example [1.D.1]: Counterexample that choice structure cannot be rationalized if B does not contain all subsets up to three elements. X = {x, y, z}. B = {B1, B2, B3} with B1 = {x, y}, B2 = {y, z} and B3 = {x, z}. C(B1) = {x}, C(B2) = {y} and C(B3) = {z}. Here the weak axiom is satisfied. To rationalize this choice structure it would be necessary that x y, y z and z x, which does not satisfy the transitivity property. 23 Consumer Theory 1 Choice Rules (9) Microeconomics • Remark: The more subsets of X we consider, the stronger the restrictions implied by the weak axiom. • Remark: If the choice structure is defined for all subsets of X, then the approaches based on the preference relation and on the weak axiom are equivalent. To consider all budget sets is a strong requirement. Another way to get equivalence, is to replace the weak axiom by the strong axiom of revealed preference. 24 Consumer Theory 2 Walrasian Demand (1) Microeconomics • Competitive budget sets. • Walrasian/Marshallian Demand • Walras’ law, Cournot and Engel aggregation. MWG: Chapter 2 25 Consumer Theory 2 Consumption Set (1) Microeconomics • We have already defined the consumption set: The set of all alternatives (complete consumption plans). We assumed X = RL +. • Each x represents a different consumption plan. • Physical restrictions: divisibility, time constraints, survival needs, etc. might lead to a strict subset of RL + as consumption set X. 26 Consumer Theory 2 Budget Set (1) Microeconomics Definition - Budget Set: B • Due to constraints (e.g. income) we cannot afford all elements in X, problem of scarcity. • The budget set B is defined by the elements of X, which are achievable given the economic realities. • B ⊂ X. 27 Consumer Theory 2 Budget Set (2) Microeconomics • By the consumption set and the budget set we can describe a consumer’s alternatives of choice. • These sets do not tell us what x is going to be chosen by the consumer. • To describe the choice of the consumer we need a theory to model or describe the preferences of a consumer (or the choice structure). 28 Consumer Theory 2 Competitive Budgets (1) Microeconomics • Assumption: All L goods are traded in the market (principle of completeness), the prices are given by the price vector p, pl > 0 for all l = 1, . . . , L. Notation: p 0. Assumption - the prices are constant and not affected by the consumer. • Given a wealth level w ≥ 0, the set of affordable bundles is described by p · x = p1x1 + · · · + pLxL ≤ w. • Definition - Walrasian Budget Set: The set Bp,w = {x ∈ RL +|p · x ≤ w} is called Walrasian or competitive budget set. [D 2.D.1] • Definition - Consumer’s problem: Given p and w choose the optimal bundle x from Bp,w. 29 Consumer Theory 2 Competitive Budgets (2) Microeconomics • Definition - Relative Price: The ratios of prices pj/pi are called Relative Prices. • Here the price of good j is expressed in terms of good i. In other words: The price of good xi is expressed in the units of good xj. • On the market we receive for one unit of xj, pj/pi · 1 units of xi. • Example: pj = 4, pi = 2. Then pj/pi = 2 and we get two units of xi for one unit of xj. 30 Consumer Theory 2 Competitive Budgets (3) Microeconomics • The budget set B describes the goods a consumer is able to buy given wealth level w. • Definition - Numeraire Good: If all prices pj are expressed in the prices of good n, then this good is called numeraire. pj/pn, j = 1, . . . , L. The relative price of the numeraire is 1. • There are L − 1 relative prices. 31 Consumer Theory 2 Competitive Budgets (4) Microeconomics • The set {x ∈ RL +|p · x = w} is called budget hyperplane, for L = 2 it is called budget line. • Given x and x in the budget hyperplane, p · x = p · x = w holds. This results in p(x − x ) = 0, i.e. p and (x − x ) are orthogonal - see Figure 2.D.3 page 22. • The budget hyperplane is a convex set. In addition it is closed and bounded ⇒ compact. 0 ∈ Bp,w (given the assumption that p 0). 32 Consumer Theory 2 Demand Functions (1) Microeconomics • Definition - Walrasian demand correspondence: The correspondence assigning to a pair (p, w) a set of consumption bundles is called Walrasian demand correspondence x(p, w); i.e. (p, w) x(w, p). If x(p, w) is single valued for all p, w, x(., .) is called demand function. • Definition - Homogeneity of degree zero: x(., .) is homogeneous of degree zero if x(αp, αw) = x(p, w) for all p, w and α > 0. [D 2.E.1] • Definition - Walras law, budget balancedness: x(., .) satisfies Walras law if for every p 0 and w > 0, we get p · x = w for all x ∈ x(p, w). That is, the consumer spends all income w with her/his optimal consumption decision. [D 2.E.2] 33 Consumer Theory 2 Demand Functions (2) Microeconomics • Assume that x(., .) is a function: • With p fixed at ¯p, the function x(¯p, .) is called Engel function. • If the demand function is differentiable we can derive the gradient vector: Dwx(p, w) = (∂x1(p, w)/∂w, . . . , xL(p, w)/∂w). If ∂xl(p, w)/∂w ≥ 0, xl is called normal or superior, otherwise it is inferior. • See Figure 2.E.1, page 25 • Notation: Dwx(p, w) results in a 1 × L row matrix, Dwx(p, w) = ( wx(p, w)) . 34 Consumer Theory 2 Demand Functions (3) Microeconomics • With w fixed, we can derive the L × L matrix of partial derivatives with respect to the prices: Dpx(p, w). • ∂xl(p, w)/∂pk = [Dpx(p, w)]l,k is called the price effect. • A Giffen good is a good where the own price effect is positive, i.e. ∂xl(p, w)/∂pl > 0 • See Figure 2.E.2-2.E.4, page 26. 35 Consumer Theory 2 Demand Functions (4) Microeconomics • Proposition: If a Walrasian demand function x(., .) is homogeneous of degree zero and differentiable, then for all p and w: L k=1 ∂xl(p, w) ∂pk pk + ∂xl(p, w) ∂w w = 0 for l = 1, . . . , L;, or in matrix notation Dpx(p, w)p + Dwx(p, w) = 0. [P 2.E.1] • Proof: By the Euler theorem (if g(.) is homogeneous of degree r, then ∂g(x)/∂x · x = rg(x), [MWG, Theorem M.B.2, p. 929]), the result follows directly when using the stacked vector x = (p , w) . Apply this to x1(p, w), . . . , xL(p, w). 36 Consumer Theory 2 Demand Functions (5) Microeconomics • Definition - Price Elasticity of Demand: ηij = ∂xi(p,w) ∂pj pj xi(p,w) . • Definition - Income Elasticity: ηiw = ∂xi(p,w) ∂w w xi(p,w) . • Definition - Income Share: si = pixi(p, w) w , where si ≥ 0 and n i=1 si = 1. 37 Consumer Theory 3 The Axiomatic Approach (1) Microeconomics • Axioms on preferences • Preference relations, behavioral assumptions and utility (axioms, utility functions) • The consumer’s problem • Walrasian/Marshallian Demand MWG, Chapter 3.A-3.D 38 Consumer Theory 3 The Axiomatic Approach (2) Microeconomics • Axiom 1 - Completeness: For all x, y ∈ X either x y, y x or both. • Axiom 2 - Transitivity: For the elements x, y, z ∈ X: If x y and y z, then x z. • We have already defined a rational preference relation by completeness and transitivity [D 3.B.1]. • If the number of elements is finite it is easy to see that one can describe a preference relation by means of a function. 39 Consumer Theory 3 The Axiomatic Approach (3) Microeconomics Sets arising from the preference relations: • (x) := {y|y ∈ X, y x} - at least as good (sub)set • (x) := {y|y ∈ X, y x} - the no better set • (x) := {y|y ∈ X, y x} - at preferred to set • (x) := {y|y ∈ X, y x} - worse than set • ∼ (x) := {y|y ∈ X, y ∼ x} - indifference set 40 Consumer Theory 3 The Axiomatic Approach (4) Microeconomics • Axiom 3.A - Local Nonsatiation: For all x ∈ X and for all ε > 0 there exists some y ∈ X such that ||x − y|| ≤ ε and y x. [D 3.B.3] • This assumptions implies that for every small distance ε there must exist at least one y, which is preferred to x. • Indifference “zones“ are excluded by this assumption. See Figure 3.B.1 on page 43. 41 Consumer Theory 3 The Axiomatic Approach (5) Microeconomics • Axiom 3.B - Monotonicity: For all x, y ∈ RL +: If x ≥ y then x y while if x y then x y (weakly monotone). It is strongly/strict monotone if x ≥ y and x = y imply x y. [D 3.B.2] • Here ≥ means that at least one element of x is larger than the elements of y, while x y implies that all elements of x are larger than the elements of y. • Remark: Local nonsatiation vs. monotonicity: The latter implies that more is always better, while Axiom 3.A only implies that in a set described by ||x − y|| ≤ ε there has to exist a preferred alternative. 42 Consumer Theory 3 The Axiomatic Approach (6) Microeconomics • Discuss the differences of Axioms 3.A and 3.B (what are their impacts on indifference sets?), e.g. by means Figures 3.B.1 and 3.B.2, page 43. 43 Consumer Theory 3 The Axiomatic Approach (7) Microeconomics • Last assumption on taste - “mixtures are preferred to extreme realizations“ • See Figure 3.B.3, page 44. • Axiom 4.A - Convexity: For every x ∈ X, if y x and z x then νy + (1 − ν)z x for ν ∈ [0, 1]. [D 3.B.4] • Axiom 4.B - Strict Convexity: For every x ∈ X, y x, z x and y = z then νy + (1 − ν)z x for ν ∈ (0, 1). [D 3.B.5] • Given these assumptions, indifference curves become (strict) convex. 44 Consumer Theory 3 The Axiomatic Approach (8) Microeconomics • Definition - Homothetic Preferences: A monotone preference on X is homothetic if all indifference sets are related by proportional expansions along rays. I.e. x ∼ y then αx ∼ αy. [D 3.B.6] • Definition - Quasilinear Preferences: A monotone preference on X = (−∞, ∞) × RL−1 is quasilinear with respect to commodity 1 if : (i) all indifference sets are parallel displacements of each other along the axis of commodity 1. That is, if x ∼ y then x + αe1 ∼ y + αe1 and e1 = (1, 0, . . . ). (ii) Good one is desirable: x + αe1 x for all α > 0. [D 3.B.7] 45 Consumer Theory 3 The Axiomatic Approach (9) Microeconomics • With the next axiom we regularize our preference order by making it continuous: • Axiom 5 - Continuity: A preference order is continuous if it is preserved under limits. For any sequence (x(n) , y(n) ) with x(n) y(n) for all n, and limits x, y (x = limn→∞ x(n) and y = limn→∞ y(n) ) we get x y. [D 3.C.1] • Equivalently: For all x ∈ X the set “at least as good as“ ( (x)) and the set “no better than“ ( (x)) are closed in X. 46 Consumer Theory 3 The Axiomatic Approach (10) Microeconomics • Topological property of the preference relation (important assumption in the existence proof of a utility function). • By this axiom the set (x) and (x) are open sets (the complement of a closed set is open ...). (x) is the complement of X\ (x). • The intersection of (x)∩ (x) is closed (intersection of closed sets). Hence, indifference sets are closed. • Consider a sequence of bundles y(n) fulfilling y(n) x, for all n. For y(n) converging to y, Axiom 5 imposes that y x. 47 Consumer Theory 3 The Axiomatic Approach (11) Microeconomics Lexicographic order/dictionary order: • Given two partially order sets X1 and X2, an order is called lexicographical on X1 × X2 if (x1, x2) (x1, x2) if and only if x1 < x1 (or x1 = x1, x2 < x2). That is, “good 1 is infinitely more desired than good 2”. • Example in R2 + (Example of Debreu): x = (0, 1) and y(n) = (1/n, 0), y = (0, 0). For all n, y(n) x, while for n → ∞: y(n) → y= (0, 0) (0, 1) = x. • The lexicographic ordering is a rational (strict) preference relation (we have to show completeness and transitivity). 48 Consumer Theory 3 The Axiomatic Approach (12) Microeconomics • Axioms 1 and 2 guarantee that an agent is able to make consistent comparisons among all alternatives. • Axiom 5 imposes the restriction that preferences do not exhibit “discontinuous behavior“ ; mathematically important • Axioms 3 and 4 make assumptions on a consumer’s taste (satiation, mixtures). 49 Consumer Theory 3 Utility Function (1) Microeconomics • Definition: Utility Function: A real-valued function u : RL + → R is called utility function representing the preference relation if for all x, y ∈ RL + u(x) ≥ u(y) if and only if x y. • I.e. a utility function is a mathematical device to describe the preferences of a consumer. • Pair-wise comparisons are replaced by comparing real valued functions evaluated for different consumption bundles. • Function is of no economic substance (for its own). 50 Consumer Theory 3 Utility Function (2) Microeconomics • First of all we want to know if such a function exists. • Theorem: Existence of a Utility Function: If a binary relation is complete, transitive and continuous, then there exists a continuous real valued function function u(x) representing the preference ordering . • Proof: by assuming monotonicity see page 47; Debreu’s (1959) proof is more advanced.[P 3.C.1] 51 Consumer Theory 3 Utility Function (3) Microeconomics • Consider y = u(x) and the transformations v = g(u(x)); v = log y, v = y2 , v = a + by, v = −a − by (see MWG, page 49). Do these transformations fulfill the properties of a utility function? • Theorem: Invariance to Positive Monotonic Transformations: Consider the preference relation and the utility function u(x) representing this relation. Then also v(x) represents if and only if v(x) = g(u(x)) is strictly increasing on the set of values taken by u(x). 52 Consumer Theory 3 Utility Function (4) Microeconomics Proof: • ⇒ Assume that x y with u(x) ≥ u(y): A strictly monotone transformation g(.) then results in g(u(x) ≥ g(u(y)). I.e. v(x) is a utility function describing the preference ordering of a consumer. 53 Consumer Theory 3 Utility Function (5) Microeconomics Proof: • ⇐ Now assume that g(u(x)) is a utility representation, but g is not strictly positive monotonic on the range of u(.): Then g(u(x)) is not > to g(u(y)) for some pair x, y where u(x) > u(y). Hence, for the pair x, y we have x y since u(x) > u(y), but g(u(x)) ≤ g(u(y)). This contradicts the assumption that v(.) = g(u(.)) is a utility representation of . 54 Consumer Theory 3 Utility Function (6) Microeconomics • By Axioms 1,2,5 the existence of a utility function is guaranteed. By the further Axioms the utility function exhibits the following properties. • Theorem: Preferences and Properties of the Utility Function: – u(x) is strictly increasing if and only if is strictly monotonic. – u(x) is quasiconcave if and only if is convex: u(xν ) ≥ min{u(x), u(y)}, where xν = νx + (1 − ν)y. – u(x) is strictly quasiconcave if and only if is strictly convex. That is, u(xν ) > min{u(x), u(y)} for xν = νx + (1 − ν)y, x = y and ν ∈ (0, 1). 55 Consumer Theory 3 Utility Function (7) Microeconomics • Definition: Indifference Curve: Bundles where utility is constant (in R2 ). • Marginal rate of substitution and utility: Assume that u(x) is differentiable, then du(x1, x2) = ∂u(x1, x2) ∂x1 dx1 + ∂u(x1, x2) ∂x2 dx2 = 0 dx2 dx1 = − ∂u(x1, x2)/∂x1 ∂u(x1, x2)/∂x2 MRS12 = ∂u(x1, x2)/∂x1 ∂u(x1, x2)/∂x2 . • The marginal rate of substitution describes the trade-off between goods 1 and 2 that marginally keep the consumer indifferent at a given consumption bundle (x1, x2). That is, the “amount of good 2” the consumer has to obtain for giving up “one unit of good 1” while staying at the same utility level. 56 Consumer Theory 3 Utility Function (8) Microeconomics • If u(x) is differentiable and the preferences are strictly monotonic, then marginal utility is strictly positive. • With strictly convex preferences the marginal rate of substitution is a strictly decreasing function (i.e. in R2 the slope of the indifference curve becomes flatter). • For a quasiconcave utility function (i.e. u(xν ) ≥ min{u(x1 ), u(x2 )}, with xν = νx1 + (1 − ν)x2 ) and its Hessian H(u(x))D2 (u(x)) we get: yH(u(x))y ≤ 0 for all vectors y, where grad(u(x)) · y = 0. That is, when moving from x to y that is tangent to the indifference surface at x utility does not increase (decreases if the equality is strict). 57 Consumer Theory 3 Consumer’s Problem (1) Microeconomics • The consumer is looking for a bundle x∗ such that x∗ ∈ B and x∗ x for all x in the feasible set B. • Assume that the preferences are complete, transitive, continuous, strictly monotonic and strictly convex. Then can be represented by a continuous, strictly increasing and strictly quasiconcave utility function. Moreover we can assume that we can take first and second partial derivatives of u(x). These are usual assumptions, we can also solve the utility maximization problem (UMP) with less stringent assumptions. 58 Consumer Theory 3 Consumer’s Problem (1) Microeconomics • We assume prices pi > 0, p = (p1, . . . , pL) is the vector of prices. We assume that the prices are fixed from the consumer’s point of view. (Notation: p 0 means that all coordinates of p are strictly larger than zero.) • The consumer is endowed with wealth w. 59 Consumer Theory 3 Consumer’s Problem (2) Microeconomics • Budget set induced by w: Bp,w = {x|x ∈ RL + ∧ p · x ≤ w}. • With the constant w and the consumption of the other goods constant, we get: dw = p1dx1 + p2dx2 = 0 dx2 dx1 = − p1 p2 with other prices constant . • Budget line with two goods; slope −p1/p2. See Figure 2.D.1, page 21, MWG. 60 Consumer Theory 3 Consumer’s Problem (3) Microeconomics • Definition - Utility Maximization Problem [UMP]: Find the optimal solution for: max x u(x) s.t. xi ≥ 0 , p · x ≤ w. The solution x(p, w) is called Walrasian demand . • Remark: Some textbooks call the UMP also Consumer’s Problem. • Remark: Some textbooks call x(p, w) Marshallian demand. 61 Consumer Theory 3 Consumer’s Problem (4) Microeconomics • Proposition - Existence: If p 0, w > 0 and u(x) is continuous, then the utility maximization problem has a solution. [P 3.D.1] • Proof: By the assumptions Bp,w is compact. u(x) is a continuous function. By the Weierstraß theorem (Theorem M.F.2(ii), p. 945, MWG; maximum value theorem in calculus), there exists an x ∈ Bp,w maximizing u(x). 62 Consumer Theory 3 Consumer’s Problem (5) Microeconomics • By altering the price vector p and income w, the consumer’s maximization provides us with the correspondence x(p, w), which is called Walrasian demand correspondence. If preferences are strictly convex we get Walrasian demand functions x(p, w). • What happens to the function if w or pj changes? See MWG, Figure 3.D.1 - 3.D.4 63 Consumer Theory 3 Consumer’s Problem (6) Microeconomics • In a general setting demand need not be a smooth function. • Theorem - Differentiable Walrasian Demand Function: Let x∗ 0 solve the consumers maximization problem at price p0 0 and w0 > 0. If u(x) is twice continuously differentiable, ∂u(x)/∂xi > 0 for some i = 1, . . . , n and the bordered Hessian of u(x), D2 u(x) u(x) u(x) 0 , has a non-zero determinant at x∗ , then x(p, w) is differentiable at p0, w0. • More details are provided in MWG, p. 94-95. 64 Consumer Theory 3 Consumer’s Problem (7) Microeconomics • In a general setting demand need not be a smooth function. • Theorem - Differentiable Walrasian Demand Function: Let x∗ 0 solve the consumers maximization problem at price p0 0 and w0 > 0. If u(x) is twice continuously differentiable, ∂u(x)/∂xi > 0 for some i = 1, . . . , n and the bordered Hessian of u(x), D2 u(x) u(x) u(x) 0 , has a non-zero determinant at x∗ , then x(p, w) is differentiable at p0, w0. • More details are provided in MWG, p. 94-95. 65 Correspondences (1) Microeconomics • Generalized concept of a function. • Definition - Correspondence: Given a set A ∈ Rn , a correspondence f : A Rk is a rule that assigns a set f(x) ⊆ Y ⊂ Rk to every x ∈ A. • If f(x) contains exactly one element for every x ∈ A, then (up to abuse of notation) f is a function. • A ⊆ Rn and Y ⊆ Rk are the domain and the codomain. • Literture: e.g. MWG, chapter M.H, page 949. 66 Correspondences (2) Microeconomics • The set {(x, y)|x ∈ A , y ∈ Rk , y ∈ f(x)} is called graph of the correspondence. • Definition - Closed Graph: A correspondence has a closed graph if for any pair of sequences x(m) → x ∈ A, with x(m) ∈ A and y(m) → y, with y(m) ∈ f(x(m) ), we have y ∈ f(x). 67 Correspondences (3) Microeconomics • Regarding continuity there are two concepts with correspondences. • Definition - Upper Hemicontinuous: A correspondence is UHC if the graph is closed and the images of compact sets are bounded. That is, for every compact set B ⊆ A, the set f(B) = {y ∈ Rk : y ∈ f(x) for some x ∈ B} is bounded. • Definition - Lower Hemicontinuous: Given A ⊆ Rn and a compact set Y ⊆ Rk , the correspondence is LHC if for every sequence x(m) → x, x(m) , x ∈ A for all m, and every y ∈ f(x), we can find a sequence y(m) → y and an integer M such that y(m) ∈ f(x(m) ) for m > M. 68 Consumer Theory 3 Consumer’s Problem (8) Microeconomics • Theorem - Properties of x(p, w): Consider a continuous utility function u(x) representing a rational locally nonsatiated preference relation defined on the consumption set X = RL +. Then x(p, w) has the following properties: [P 3.D.2] – Homogeneity of degree zero in (p, w). – Walras’ law: p · x = w for all x ∈ x(p, w). – Convexity/uniqueness: If is convex, so that u(x) is quasiconcave, then x(p, w) is a convex set. If is strictly convex, where u(x) is strictly quasiconcave, then x(p, w) consists of a single element. 69 Consumer Theory 3 Consumer’s Problem (9) Microeconomics Proof: • Property 1 - Homogeneity in p, w: We have to show that x(µp, µw) = µ0 x(p, w). Plug in µp and µw in the optimization problem ⇒ Bp,w = Bµp,µw. The result follows immediately. 70 Consumer Theory 3 Consumer’s Problem (10) Microeconomics Proof: • Property 2- Walras’ law: If x ∈ x(p, w) and p · x < w, then there exists a y in the neighborhood of x, with y x and p · y < w by local nonsatiation. Therefore x cannot be an optimal bundle. This argument holds for all interior points of Bp,w. 71 Consumer Theory 3 Consumer’s Problem (10) Microeconomics Proof: • Property 3 - x(p, w) is a convex set: If preferences are convex then u(xν ) ≥ min{u(x), u(y)}, where xν = νx + (1 − ν)y; replace ≥ by > if is strictly convex. I.e. u(x) is quasiconcave. We have to show that if x, y ∈ x(p, w), then xν ∈ x(p, w). From the above property x, y and xν have to be elements of the budget hyperplane {x|x ∈ X and p · x = w}. Since x and y solve the UMP we get u(x) = u(y), therefore u(xν ) ≤ u(x) = u(y). By quasiconcavity of u(x) we get u(xν ) ≥ u(x) = u(y), such that u(xν ) = u(x) = u(y) holds for arbitrary x, y ∈ x(p, w). I.e. the set x(p, w) has to be convex. 72 Consumer Theory 3 Consumer’s Problem (11) Microeconomics Proof: • Property 3 - x(p, w) is single valued if preferences are strictly convex: Assume, like above, the x and y solve the UMP; x = y. Then u(x) = u(y) ≥ u(z) for all z ∈ Bp,w. By the above result x, y are elements of the budget hyperplane. • Since preferences are strictly convex, u(x) is strictly quasiconcave ⇒ u(xν ) > min{u(x), u(y)}. xν = νx + (1 − ν)y and x , y are some arbitrary elements of the budget hyperplane; (a contradiction to strict convexity). • Now u(xν ) > min{u(x), u(y)}, also for x, y. Therefore the pair x, y cannot solve the UMP. Therefore, x(p, w) has to be single valued. 73 Consumer Theory 4 Duality Microeconomics • Instead of looking at u(x), we’ll have an alternative look on utility via prices, income and the utility maximization problem ⇒ indirect utility • Expenditure function, the dual problem and Hicksian demand • Income- and substitution effects, Slutsky equation MWG, Chapter 3.D-3.H 74 Consumer Theory 4 Indirect Utility (1) Microeconomics • We have already considered the direct utility function u(x) in the former parts. • Start with the utility maximization problem max x u(x) s.t. p · x ≤ w x∗ ∈ x(p, w) solves this problem for (p, w) 0. • Definition - Indirect Utility: By the highest levels of utility attainable with p, w, we define a maximal value function. This function is called indirect utility function v(p, w). It is the maximum value function corresponding to the consumer’s optimization problem (utility maximization problem). 75 Consumer Theory 4 Indirect Utility (2) Microeconomics • v(p, w) is a function, by Berge’s theorem of the maximum x(p, w) is upper hemicontinuous and v(p, w) is continuous (see MWG, page 963, [M.K.6], more details in Micro II). • If u(x) is strictly quasiconcave such that maximum x∗ is unique, we derive the demand function x∗ = x(p, w). • In this case the indirect utility function is the composition of the direct utility function and the demand function x(p, w), i.e. v(p, w) = u(x∗ ) = u(x(p, w)). 76 Consumer Theory 4 Indirect Utility (3) Microeconomics • Theorem: Properties of the Indirect Utility Function v(p, w): [P 3.D.3] Suppose that u(x) is a continuous utility function representing a locally nonsatiated preference relation on the consumption set X = RL +. Then the indirect utility function v(p, w) is – Continuous in p and w. – Homogeneous of degree zero in p, w. – Strictly increasing in w. – Nonincreasing in pl, l = 1, . . . , L. – Quasiconvex in (p, w). 77 Consumer Theory 4 Indirect Utility (4) Microeconomics Proof: • Property 1 - Continuity: follows from Berge’s theorem of the maximum. • Property 2 - Homogeneous in (p, w): We have to show that v(µp, µw) = µ0 v(p, w) = v(p, w); µ > 0. Plug in µp and µw in the optimization problem ⇒ v(µp, µw) = {maxx u(x) s.t. µp · x ≤ µw} ⇔ {maxx u(x) s.t. p · x ≤ w} = v(p, w). 78 Consumer Theory 4 Indirect Utility (5) Microeconomics Proof: • Property 3 - increasing in w: Given the solutions of the UMP with p and w, w , where w > w: x(p, w) and x(p, w ). • The corresponding budget sets are Bp,w and Bp,w , by assumption Bp,w ⊂ Bp,w (here we have a proper subset). • Define Sp,w = {x ∈ X|p · x = w} (Walrasian budget hyperplane). Then Bp,w is still contained in Bp,w \ Sp,w . • Therefore also Sp,w ∈ (Bp,w \ Sp,w ). From the above consideration we know that for any y ∈ Sp,w, we have p · y < w . By local nonsatiation there are better bundles in Bp,w . 79 Consumer Theory 4 Indirect Utility (6) Microeconomics Proof: • Since v(p, w) is a maximal value function, it has to increase if w increases. • In other words: By local nonsatiation Walras law has to hold, i.e. x(p, w) and x(p, w ) are subsets of the budget hyperplanes {x|x ∈ X and p · x = w}, {x|x ∈ X and p · x = w }, respectively. We know where we find the optimal bundles. The hyperplane for w is a subset of Bp,w and Bp,w (while the hyperplane for w is not contained in Bp,w). Interior points cannot be an optimum under local nonsatiation. • If v(p, w) is differentiable this result can be obtained by means of the envelope theorem. 80 Consumer Theory 4 Indirect Utility (7) Microeconomics Proof: • Property 4 - non-increasing in pl: W.l.g. pl > pl, then we get Bp,w and Bp ,w, where Bp ,w ⊆ Bp,w. But Sp ,w is not fully contained in Bp,w \ Sp,w. (Observe the ”common point” in R2 .) The rest is similar to Property 3. 81 Consumer Theory 4 Indirect Utility (8) Microeconomics Proof: • Property 5 - Quasiconvex: Consider two arbitrary pairs p1 , x1 and p2 , x2 and the convex combinations pν = νp1 + (1 − ν)p2 and wν = νw1 + (1 − ν)w2; ν ∈ [0, 1]. • v(p, w) would be quasiconvex if v(pν , wν) ≤ max{v(p1 , w1), v(p2 , w2)}. • Define the consumption sets: Bj = {x|p(j) · x ≤ wj} for j = 1, 2, ν. 82 Consumer Theory 4 Indirect Utility (9) Microeconomics Proof: • First we show: If x ∈ Bν, then x ∈ B1 or x ∈ B2. This statement trivially holds for ν equal to 0 or 1. For ν ∈ (0, 1) we get: Suppose that x ∈ Bν but x ∈ B1 or x ∈ B2 is not true (then x ∈ B1 and x ∈ B2), i.e. p1 · x > w1 ∧ p2 · x > w2 Multiplying the first term with ν and the second with 1 − ν results in νp1 · x > νw1 ∧ (1 − ν)p2 · x > (1 − ν)w2 83 Consumer Theory 4 Indirect Utility (10) Microeconomics Proof: • Summing up both terms results in: (νp1 + (1 − ν)p2 ) · x = pν · x > νw1 + (1 − ν)w2 = wν which contradicts our assumption that x ∈ Bν. • From the fact that xν ∈ x(pν , wν) is either ∈ B1 or ∈ B2, it follows that v(pν , wν) ≤ max{v(p1 , w1), v(p2 , w2)}. The last expression corresponds to the definition of a quasiconvex function. 84 Consumer Theory 4 Expenditure Function (1) Microeconomics • With indirect utility we looked at maximized utility levels given prices and income. • Now we raise the question a little bit different: what expenditures e are necessary to attain an utility level u given prices p. • Expenditures e can be described by the function e = p · x. 85 Consumer Theory 4 Expenditure Function (2) Microeconomics • Definition - Expenditure Minimization Problem [EMP]: minx p · x s.t. u(x) ≥ u, x ∈ X = RL +, p 0. (We only look at u ≥ u(0). U = {u|u ≥ u(0) ∧ u ∈ Range(u(x))} ) • It is the dual problem of the utility maximization problem. The solution of the EMP h(p, u) will be called Hicksian demand correspondence. • Definition - Expenditure Function: The minimum value function e(p, u) solving the expenditure minimization problem minx p · x s.t. u(x) ≥ u, p 0, is called expenditure function. • Existence: The Weierstraß theorem guarantees the existence of an x∗ s.t. p · x∗ are the minimal expenditures necessary to attain an utility level u. 86 Consumer Theory 4 Expenditure Function (3) Microeconomics • Theorem: Properties of the Expenditure Function e(p, u): [P 3.E.2] If u(x) is continuous utility function representing a locally nonsatiated preference relation. Then the expenditure function e(p, u) is – Continuous in p, u domain Rn ++ × U. – ∀p 0 strictly increasing in u. – Non-decreasing in pl for all l = 1, . . . , L. – Concave in p. – Homogeneous of degree one in p. 87 Consumer Theory 4 Expenditure Function (5) Microeconomics Proof: • Property 1 - continuous: Apply the theorem of the maximum. 88 Consumer Theory 4 Expenditure Function (6) Microeconomics Proof: • Property 2 - increasing in u: We have to show that is u2 > u1 then e(p, u2) > e(p, u1). • Suppose that h1 ∈ h(p, u1) and h2 ∈ h(p, u2) solve the EMP for u2 and u1, but e(p, u2) ≤ e(p, u1). We show that this result in a contradiction. I.e. u2 > u1 but 0 ≤ p · h2 ≤ p · h1. • Then by continuity of u(x) and local nonsatiation we can find an α ∈ (0, 1) such that αh2 is preferred to h1 (remember u2 > u1 is assumed) with expenditures αp · h2 < p · h1. This contradicts that h1 solves the EMP for p, u1. 89 Consumer Theory 4 Expenditure Function (7) Microeconomics Proof: • Property 2 - with calculus: From minx p · x s.t. u(x) ≥ u , x ≥ 0 we derive the Lagrangian: L(x, λ) = p · x + λ(u − u(x)) . • From this Kuhn-Tucker problem we get: ∂L ∂xi = pi − λ ∂u(x) ∂xi ≥ 0 , ∂L ∂xi xi = 0 ∂L ∂λ = u − u(x) ≤ 0 , ∂L ∂λ λ = 0 90 Consumer Theory 4 Expenditure Function (8) Microeconomics Proof: • λ = 0 would imply that utility could be increased without increasing the expenditures (in an optimum) ⇒ u = u(x) and λ > 0. • Good xi is demanded if the price does not exceed λ∂u(x) ∂xi for all xi > 0. • The envelope theorem tells us that ∂e(p, u) ∂u = ∂L(x, u) ∂u = λ > 0 • Since u(x) is continuous and increasing the expenditure function has to be unbounded. 91 Consumer Theory 4 Expenditure Function (9) Microeconomics Proof: • Property 3 - non-decreasing in pl: similar to property 3. 92 Consumer Theory 4 Expenditure Function (10) Microeconomics Proof: • Property 4 - concave in p: Consider an arbitrary pair p1 and p2 and the convex combination pν = νp1 + (1 − ν)p2 . The expenditure function is concave if e(pν, u) ≥ νe(p1 , u) + (1 − ν)e(p2 , u). • For minimized expenditures it has to hold that p1 x1 ≤ p1 x and p2 x2 ≤ p2 x for all x fulfilling u(x) ≥ u. • x∗ ν minimizes expenditure at a convex combination of p1 and p2 . • Then p1 x1 ≤ p1 x∗ ν and p2 x2 ≤ p2 x∗ ν have to hold. 93 Consumer Theory 4 Expenditure Function (11) Microeconomics Proof: • Multiplying the first term with ν and the second with 1 − ν and taking the sum results in νp1 x1 + (1 − ν)p2 x2 ≤ pνx∗ ν. • Therefore the expenditure function is concave in p. 94 Consumer Theory 4 Expenditure Function (12) Microeconomics Proof: • Property 5 - homogeneous of degree one in p: We have to show that e(µp, u) = µ1 e(p, u); µ > 0. Plug in µp in the optimization problem ⇒ e(µp, u) = {minx µp · x s.t. u(x) ≥ u}. Objective function is linear in µ, the constraint is not affected by µ. With calculus we immediately see the µ cancels out in the first order conditions ⇒ xh remains the same ⇒ µ{minx p · x s.t. u(x) ≥ u} = µe(p, u). 95 Consumer Theory 4 Hicksian Demand (1) Microeconomics • Theorem: Hicksian demand: [P 3.E.3] Let u(x) be continuous utility function representing a locally nonsatiatated preference order; p 0. Then the Hicksian demand correspondence has the following properties: – Homogeneous of degree zero. – No excess utility u(x) = u. – Convexity/uniqueness: If is convex, then h(p, u) is a convex set. If is strictly convex, then h(p, u) is single valued. 96 Consumer Theory 4 Hicksian Demand (2) Microeconomics Proof: • Homogeneity follows directly from the EMP. min{p · x s.t. u(x) ≥ u} ⇔ α min{p · x s.t. u(x) ≥ u} ⇔ min{αp · x s.t. u(x) ≥ u} for α > 0. • Suppose that there is an x ∈ h(p, u) with u(x) > u. By the continuity of u we find an α ∈ (0, 1) such that x = αx and u(x ) > u. But with x we get p · x < p · x. A contradiction that x solves the EMP. • For the last property see the theorem on Walrasian demand or apply the forthcoming theorem. 97 Consumer Theory 4 Expenditure vs. Indirect Utility (1) Microeconomics • With (p, w) the indirect utility function provides us with the maximum of utility u. Suppose w = e(p, u). By this definition v(p, e(p, u)) ≥ u. • Given p, u and an the expenditure function, we must derive e(p, v(p, w)) ≤ w. • Given an x∗ solving the utility maximization problem, i.e. x∗ ∈ x(p, w). Does x∗ solve the EMP if u = v(p, w)? • Given an h∗ solving the EMP, i.e. h∗ ∈ h(p, u). Does h∗ solve the UMP if w = e(p, u)? 98 Consumer Theory 4 Expenditure vs. Indirect Utility (2) Microeconomics • Theorem: Equivalence between Indirect Utility and Expenditure Function: [P 3.E.1] Let u(x) be continuous utility function representing a locally nonsatiatated preference order; p 0. – If x∗ is optimal in the UMP with w > 0, then x∗ is optimal in the EMP when u = u(x∗ ). e(p, u(x∗ )) = w. – If h∗ is optimal in the EMP with u > u(0), then h∗ is optimal in the UMP when w = e(p, u). v(p, e(p, u)) = u. 99 Consumer Theory 4 Expenditure vs. Indirect Utility (3) Microeconomics Proof: • We prove e(p, v(p, w)) = w by means of a contradiction. p, w ∈ Rn ++ × R++. By the definition of the expenditure function we get e(p, v(p, w)) ≤ w. In addition h∗ ∈ h(p, u). To show equality assume that e(p, u) < w, where u = v(p, w) and x∗ solves the UMP: e(p, u) is continuous in u. Choose ε such that e(p, u + ε) < w and e(p, u + ε) =: wε. The properties of the indirect utility function imply v(p, wε) ≥ u + ε. Since wε < w and v(p, w) is strictly increasing in w (by local nonstatiation) we get: v(p, w) > v(p, wε) ≥ u + ε but u = v(p, w), which is a contradiction. Therefore e(p, v(p, w)) = w and x∗ also solves the EMP, such that x∗ ∈ h(p, u) when u = v(p, u). 100 Consumer Theory 4 Expenditure vs. Indirect Utility (4) Microeconomics Proof: • Next we prove v(p, e(p, u)) = u in the same way. p, u ∈ Rn ++ × U. By the definition of the indirect utility function we get v(p, e(p, u)) ≥ u. Assume that v(p, w) > u, where w = e(p, u) and h∗ solves the EMP: v(p, w) is continuous in w. Choose ε such that v(p, w − ε) > u and v(p, w − ε) =: uε. The properties of the expenditure function imply e(p, uε) ≤ w − ε. Since uε > u and e(p, u) is strictly increasing in u we get: e(p, u) < e(p, uε) ≤ w − ε but w = e(p, u), which is a contradiction. Therefore v(p, e(p, u)) = u. In addition h∗ also solves the UMP. 101 Consumer Theory 4 Hicksian Demand (3) Microeconomics • Theorem: Hicksian/ Compensated law of demand: [P 3.E.4] Let u(x) be continuous utility function representing a locally nonsatiatated preference order and h(p, u) consists of a single element for all p 0. Then the Hicksian demand function satisfies the compensated law of demand: For all p and p : (p − p )[h(p , u) − h(p , u)] ≤ 0. 102 Consumer Theory 4 Hicksian Demand (4) Microeconomics Proof: • By the EMP: p · h(p , u) − p · h(p , u) ≤ 0 and p · h(p , u) − p · h(p , u) ≤ 0 have to hold. • Adding up the inequalities yields the result. 103 Consumer Theory 4 Shephard’s Lemma (1) Microeconomics • Investigate the relationship between a Hicksian demand function and the expenditure function. • Theorem - Shephard’s Lemma: [P 3.G.1] Let u(x) be continuous utility function representing a locally nonsatiatated preference order and h(p, u) consists of a single element. Then for all p and u, the gradient vector of the expenditure function with respect to p gives Hicksian demand, i.e. pe(p, u) = h(p, u). 104 Consumer Theory 4 Shephard’s Lemma (2) Microeconomics Proof by means of calculus: • Suppose that the envelope theorem can be applied (see e.g. MWG [M.L.1], page 965): • Then the Lagrangian is given by: L(x, λ) = p · x + λ(u − u(x)). • λ > 0 follows from u = u(x). 105 Consumer Theory 4 Shephard’s Lemma (3) Microeconomics Proof with calculus: • The Kuhn-Tucker conditions are: ∂L ∂xi = pi − λ ∂u(x) ∂xi ≥ 0 ∂L ∂xi xi = 0 ∂L ∂λ = u − u(x) ≥ 0 ∂L ∂λ λ = 0 106 Consumer Theory 4 Shephard’s Lemma (4) Microeconomics Proof with calculus: • Good xi is demanded if the price does not exceed λ∂u(x) ∂xi for all xi > 0. • The envelope theorem tells us that ∂e(p, u) ∂pl = ∂L(x, u) ∂pl = hl(p, u) for l = 1, . . . , L. 107 Consumer Theory 4 Shephard’s Lemma (5) Microeconomics Proof: • The expenditure function is the support function µk of the non-empty and closed set K = {x|u(x) ≥ u}. Since the solution is unique by assumption, µK(p) = pe(p, u) = h(p, u) has to hold by the Duality theorem. • Alternatively: Assume differentiability and apply the envelope theorem. 108 Consumer Theory 4 Expenditure F. and Hicksian Demand (1) Microeconomics • Furthermore, investigate the relationship between a Hicksian demand function and the expenditure function. • Theorem:: [P 3.E.5] Let u(x) be continuous utility function representing a locally nonsatiatated and strictly convex preference relation on X = RL +. Suppose that h(p, u) is continuously differentiable, then – Dph(p, u) = D2 pe(p, u) – Dph(p, u) is negative semidefinite – Dph(p, u) is symmetric. – Dph(p, u)p = 0. 109 Consumer Theory 4 Expenditure F. and Hicksian Demand (2) Microeconomics Proof: • To show Dph(p, u)p = 0, we can use the fact that h(p, u) is homogeneous of degree zero in prices (r = 0). • By the Euler theorem [MWG, Theorem M.B.2, p. 929] we get L l=1 ∂h(p, u) ∂pl pl = rh(p, u). 110 Consumer Theory 4 Walrasian vs. Hicksian Demand (1) Microeconomics • Here we want to analyze what happens if income w changes: normal vs. inferior good. • How is demand effected by prices changes: change in relative prices - substitution effect, change in real income - income effect • Properties of the demand and the law of demand. • How does a price change of good i affect demand of good j. • Although utility is continuous and strictly increasing, there might be goods where demand declines while the price falls. 111 Consumer Theory 4 Walrasian vs. Hicksian Demand (2) Microeconomics • Definition - Substitution Effect, Income Effect: We split up the total effect of a price change into – an effect accounting for the change in the relative prices pi/pj (with constant utility or real income) ⇒ substitution effect. Here the consumer will substitute the relatively more expensive good by the cheaper one. – an effect induced by a change in real income (with constant relative prices) ⇒ income/wealth effect. 112 Consumer Theory 4 Walrasian vs. Hicksian Demand (3) Microeconomics • Hicksian decomposition - keeps utility level constant to identify the substitution effect. • The residual between the total effect and the substitution effect is the income effect. • See Figures in MWG, Chapter 2 and 113 Consumer Theory 4 Walrasian vs. Hicksian Demand (4) Microeconomics • Here we observe that the Hicksian demand function exactly accounts for the substitution effect. • The difference between the change in Walrasian (total effect) demand induced by a price change and the change in Hicksian demand (substitution effect) results in the income effect. • Note that the income effect need not be positive. 114 Consumer Theory 4 Walrasian vs. Hicksian Demand (5) Microeconomics • Formal description of these effects is given by the Slutsky equation. • Theorem - Slutsky Equation: [P 3.G.3] Assume that the consumer’s preference relation is complete, transitive, continuous, locally nonsatiated and strictly convex defined on X = RL +. Then for all (p, w) and u = v(p, w) we have ∂xl(p, w) ∂pj T E = ∂hl(p, u) ∂pj SE −xj(p, w) ∂xl(p, w) ∂w IE l, j = 1, . . . , L. 115 Consumer Theory 4 Walrasian vs. Hicksian Demand (6) Microeconomics • Equivalently: Dph(p, u) = Dpx(p, w) + Dwx(p, w)x(p, w) • Remark: In the following proof we shall assume that h(p, u) and x(p, w) are differentiable. (Differentiability of h(p, u) follows from duality theory presented in Section 3.F. This is a topic of the Micro II course. ) 116 Consumer Theory 4 Walrasian vs. Hicksian Demand (7) Microeconomics Proof: • First, we use the Duality result on demand: hl(p, u) = xl(p, e(p, u)) and take partial derivatives with respect to pj: ∂hl(p, u) ∂pj = ∂xl(p, e(p, u)) ∂pj + ∂xl(p, e(p, u)) ∂w ∂e(p, u) ∂pj . • Second: By the relationship between the expenditure function and the indirect utility it follows that u = v(p, w) and e(p, u) = e(p, v(p, w)) = w. 117 Consumer Theory 4 Walrasian vs. Hicksian Demand (8) Microeconomics Proof: • Third: Shephard’s Lemma tells us that ∂e(p,u) ∂pj = hj(p, u), this gives ∂hl(p, u) ∂pj = ∂xl(p, w) ∂pj + ∂xl(p, w) ∂w hj(p, u) 118 Consumer Theory 4 Walrasian vs. Hicksian Demand (9) Microeconomics Proof: • Forth: Duality between Hicksian and Walrasian demand implies that h(p, v(p, w)) = x(p, w) with v(p, w) = u. Thus ∂e(p,u) ∂pj = xj(p, w). • Arranging terms yields: ∂xl(p, w) ∂pj = ∂hl(p, u) ∂pj − xj(p, w) ∂xl(p, w) ∂w . 119 Consumer Theory 4 Walrasian vs. Hicksian Demand (10) Microeconomics • From the Sultsky equation we can construct the following matrix: Definition - Slutsky Matrix: S(p, w) :=      ∂x1(p,w) ∂p1 + x1(p, w) ∂x1(p,w) ∂w · · · ∂x1(p,w) ∂pL + xL(p, w) ∂x1(p,w) ∂w . . . ... . . . ∂xL(p,w) ∂p1 + x1(p, w) ∂xL(p,w) ∂w · · · ∂xL(p,w) ∂pL + xL(p, w) ∂xL(p,w) ∂w      120 Consumer Theory 4 Walrasian vs. Hicksian Demand (11) Microeconomics • Theorem Suppose that e(p, u) is twice continuously differentiable. Then the Slutsky Matrix S(p, w) is negative semidefinite, symmetric and satisfies S(p, w)p = 0. 121 Consumer Theory 4 Walrasian vs. Hicksian Demand (12) Microeconomics Proof: • Negative semidefiniteness follows from the negative semidefiniteness of Dph(p, u) which followed from the concavity of the expenditure function. • Symmetry follows from the existence of the expenditure function and Young’s theorem. • S(p, w) · p = 0 follows from an Euler theorem (see [MWG, Theorem M.B.2, p. 929]) argument already used in [P 3.G.2] 122 Consumer Theory 4 Roy’s Identity (1) Microeconomics • Goal is to connect Walrasian demand with the indirect utility function. • Theorem - Roy’s Identity: [P 3.G.4] Let u(x) be continuous utility function representing a locally nonsatiatated and strictly convex preference relation defined on X = RL +. Suppose that the indirect utility function v(p, w) is differentiable for any p, w 0, then x(p, w) = − 1 wv(p, w) pv(p, w) i.e. xl(p, w) = − ∂v(p, w)/∂pl ∂v(p, w)/∂w . 123 Consumer Theory 4 Roy’s Identity (1) Microeconomics Proof: • Roy’s Identity: Assume that the envelope theorem can be applied to v(p, w). • Let (x∗ , λ∗ ) maximize {maxx u(x) s.t. p · x ≤ w} then the partial derivatives of the Lagrangian L(x, λ) with respect to pl and w provide us with: ∂v(p, w) ∂pl = ∂L(x∗ , λ∗ ) ∂pl = −λ∗ x∗ l , l = 1, . . . , L. ∂v(p, w) ∂w = ∂L(x∗ , λ∗ ) ∂w = λ∗ . 124 Consumer Theory 2 Indirect Utility (11) Microeconomics Proof: • Plug in −λ from the second equation results in ∂v(p, w) ∂pl = − ∂v(p, w) ∂w x∗ l such that − ∂v(p, w)/∂pl ∂v(p, w)/∂w = xl(p, w). • Note that ∂v(p, w)/∂w by our properties on the indirect utility function. 125 Theorem of the Maximum (1) Microeconomics • Consider a constrained optimization problem: max f(x) s.t. g(x, q) = 0 where q ∈ Q is a vector of parameters. Q ∈ RS and x ∈ RN . f(x) is assumed to be continuous. C(q) is the constraint set implied by g. 126 Theorem of the Maximum (1) Microeconomics • Definition: x(q) is the set of solutions of the problem, such that x(q) ⊂ C(q) and v(q) is the maximum value function, i.e. f(x) evaluated at an optimal x ∈ x(q). • Theorem of the Maximum: Suppose that the constraint correspondence is continuous and f is continuous. Then the maximizer correspondence x : Q → RN is upper hemicontinuous and the value function v : Q → R is continuous. [T M.K.6], page 963. 127 Duality Theorem (1) Microeconomics • Until now we have not shown that c(w, y) or e(p, u) is differentiable when u(x) is strictly quasiconcave. • This property follows from the Duality Theorem. • MWG, Chapter 3.F, page 63. 128 Duality Theorem (2) Microeconomics • A set is K ∈ Rn is convex if αx + (1 − α)y ∈ K for all x, y ∈ K and α ∈ [0, 1]. • A half space is a set of the form {x ∈ Rn |p · x ≥ c}. • p = 0 is called the normal vector: if x and x fulfill p · x = p · x = c, then p · (x − x ) = 0. • The boundary set {x ∈ Rn |p · x = c} is called hyperplane. The half-space and the hyperplane are convex. 129 Duality Theorem (3) Microeconomics • Assume that K is convex and closed. Consider ¯x /∈ K. Then there exists a half-space containing K and excluding ¯x. There is a p and a c such that p · ¯x < c ≤ p · x for all x ∈ K (separating hyperplane theorem). • Basic idea of duality theory: A closed convex set can be equivalently (dually) described by the intersection of half-spaces containing this set. • MWG, figure 3.F.1 and 3.F.2 page 64. 130 Duality Theorem (4) Microeconomics • If K is not convex the intersection of the half-spaces that contain K is the smallest, convex set containing K. (closed convex hull of K, abbreviated by ¯K). • For any closed (but not necessarily convex) set K we can define the support function of K: µK(p) = inf{p · x|x ∈ K} • When K is convex the support function provides us with the dual description of K. • µK(p) is homogeneous of degree one and concave in p. 131 Duality Theorem (5) Microeconomics • Theorem - Duality Theorem: Let K be a nonempty closed set and let µK(p) be its support function. Then there is a unique ¯x ∈ K such that ¯p · ¯x = µK(¯p) if and only if µK(p) is differentiable at ¯p. In this case pµK(¯p) = ¯x. • Proof see literature. E.g. see section 25 in R.T. Rockafellar, Convex Analysis, Princeton University Press, New York 1970. 132 Consumer Theory 4 Revealed Preference Theory (1) Microeconomics • Weak Axiom of Revealed Preference. • Strong Axiom of Revealed Preference. • Revealed preferences and utility maximization. MWG, Chapter 1.C and 2.F., 3.J. 133 Consumer Theory 4 Revealed Preference Theory (2) Microeconomics • Samuelson’s idea: Cannot we start with observed behavior instead of assumptions on preferences. • Idea: if a consumer buys a bundle x0 instead of an other affordable bundle x1 , then the first bundle is called revealed preferred to x1 (see Consumer Theory 1). • Definition - Weak Axiom on Revealed Preference: [D 2.F.1] A Walrasian demand function x(p, w) satisfies the weak axiom of revealed preference if for any two wealth price situations (p,w) and (p’,w’) the following relationship holds: If p · x(p , w ) ≤ w and x(p , w ) = x(p, w) then p · x(p, w) > w . 134 Consumer Theory 4 Revealed Preference Theory (3) Microeconomics • Interpret the weak axiom by means for Figure 2.F.1, page 30. • We assume that x(p, w) is a function, which is homogeneous of degree zero and Walras’ law holds. 135 Consumer Theory 4 Revealed Preference Theory (4) Microeconomics • From the former parts we already know: Theorem - Weak Revealed Preference and Utility maximization: If x(p, w) solves the utility maximization problem with strictly increasing and strictly quasiconcave utility function, then the weak axiom of revealed preference has to hold. • See also P 1.D.1 136 Consumer Theory 4 Revealed Preference Theory (5) Microeconomics Proof: • Consider a pair x0 and x1 where x0 = x(p0 , w) solves the utility maximization problem for p0 , x1 for p1 . • Assume u(x0 ) > u(x1 ): w = p0 · x0 ≥ p0 · x1 . Then p1 · x0 > p1 · x1 = w. Otherwise a consumer would have chosen x0 if it were affordable in the second maximization problem. • I.e. p1 · x0 > p1 · x1 has to be fulfilled. Since x0 and x1 are arbitrary pairs, the weak axiom of revealed preference has to hold. 137 Consumer Theory 4 Slutsky Compensation (1) Microeconomics • Definition - Slutsky compensation: Given a bundle x0 = x(p, w) and income is compensated such that the consumer can always buy the bundle x0 , i.e. w = p · x(p, w). Then demand is called Slutsky compensated demand xS (p, w(x0 )). • Discuss this concept by means of Figure 2.F.2, page 31. 138 Consumer Theory 4 Slutsky Compensation (2) Microeconomics • Proposition: Suppose that the Walrasian demand function x(p, w) is homogeneous of degree zero and satisfies Walras’ law. Then x(p, w) satisfies the weak axiom if and only if the following property holds: For any compensated price change form the initial situation (p, w) to a new pair (p , w ), where w = p · x(p, w), we have (p − p) · [x(p , w ) − x(p, w)] ≤ 0 with strict inequality whenever x(p, w) = x(p , w ). [P 2.F.1] • Remark: x(p , w ) = xS (p , w(x0 )). 139 Consumer Theory 4 Slutsky Compensation (3) Microeconomics Proof: • (i) The weak axiom implies (p − p) · [x(p , w ) − x(p, w)] ≤ 0 with strict inequality for different demands: If x(p , w ) = x(p, w) then [x(p , w ) − x(p, w)] = 0. • Suppose x(p , w ) = x(p, w) and expand (p − p) · [x(p , w ) − x(p, w)] to p · [x(p , w ) − x(p, w)] − p · [x(p , w ) − x(p, w)]. By Walras’ law and the construction of compensated demand the first term is 0. • By compensated demand we get p · x(p, w) = w . I.e. x0 = x(p, w) can be bought with p , w . By the weak axiom x(p , w ) /∈ Bp,w, such that p · x(p , w ) > w. Walras’ law implies p · x(p, w) = w. This yields p · [x(p , w ) − x(p, w)] > 0, such that ... holds. 140 Consumer Theory 4 Slutsky Compensation (4) Microeconomics Proof: • (ii) (p − p) · [x(p , w ) − x(p, w)] ≤ 0 implies the weak law if x(p , w ) = x(p, w): • If we consider compensated demand, then the weak axiom has to hold (replace Walrasian demand by compensated demand in the Theorem - Weak Revealed Preference and Utility maximization. • It is necessary that the weak axiom holds for all compensated demand changes: Assume u(x0 ) > u(x1 ): w = p0 · x0 ≥ p0 · x1 . Suppose that p1 · x0 ≤ p1 · x1 = w. Then x0 cannot be an optimum by local non-satiation. • By these arguments the weak law holds if p · x(p , w ) > w whenever p · x(p, w) = w and x(p , w ) = x(p, w). 141 Consumer Theory 4 Slutsky Compensation (5) Microeconomics Proof: • By this argument we can test for the weak axiom by looking at compensated price changes. (We show that ¬H ⇒ ¬C.) If the weak law does not hold, there is a compensated price change such that p x(p, w ) ≤ w ⇒ p x(p, w) ≤ w , < for different x (p · x(p, w) = w). By Walras law we get p · [x(p , w ) − x(p, w)] ≤ 0 and p · [x(p , w ) − x(p, w)] ≥ 0. • This results in (p − p) · [x(p , w ) − x(p, w)] ≥ 0; > for x(p , w ) = x(p, w). This contradicts that (p − p) · [x(p , w ) − x(p, w)] ≤ 0 holds. 142 Consumer Theory 4 Revealed Preference Theory (6) Microeconomics • Definition - Strong Axiom of Revealed Preference: [3.J.1] The market demand satisfies the strong axiom of revealed preference if for any list (p1 , w1 ), . . . , (pN , wN ) with x(pn+1 , wn+1 ) = x(pn , wn ) for all n ≤ N − 1, we have pN x(p1 , w1 ) > wN whenever pn · x(pn+1 , wn+1 ) ≤ wn for all n ≤ N − 1. • I.e. if x(p1 , w1 ) is directly or indirectly revealed preferred to x(pN , wN ), then x(pN , wN ) cannot be directly or indirectly be revealed preferred to x(p1 , w1 ). Or for different bundles x1 , x2 , . . . , xk : If xq is revealed preferred to x2 and x2 is preferred to x3 , then x1 is revealed preferred to x3 . 143 Consumer Theory 4 Revealed Preference Theory (7) Microeconomics • Theorem - Revealed Preference and Demand (II): If the Walrasian demand function x(p, w) satisfies the strong axiom of revealed preference then there is a rational preference relation that rationalizes x(p, w). I.e. for all (p, w), x(p, w) y for every y = x(p, w) with y ∈ Bp,w. [P 3.J.1]. • Proof - see page 92. 144 Consumer Theory 5 Welfare Analysis (1) Microeconomics • Measurement of Welfare • Concept of the Equivalent Variation, the Compensating Variation and the Consumer Surplus. • Pareto improvement and Pareto efficient Literature: MWG, Chapter 3.I, page 80-90. 145 Consumer Theory 5 Welfare Analysis (2) Microeconomics • From a social point of view - can we judge that some market outcomes are better or worse? • Positive question: How will a proposed policy affect the welfare of an individual? • Normative question: How should we weight different effects on different individuals? 146 Consumer Theory 5 Welfare Analysis (3) Microeconomics • Definition - Pareto Improvement: When we can make someone better off and no one worse off, then a Pareto improvement can be made. • Definition - Pareto Efficient: A situation where there is no way to make somebody better off without making someone else worth off is called Pareto efficient. I.e. there is no way for Pareto improvements. • Strong criterion. 147 Consumer Theory 5 Consumer Welfare Analysis (1) Microeconomics • Preference based consumer theory investigates demand from a descriptive perspective. • Welfare Analysis can be used to perform a normative analysis. • E.g. how do changes of prices or income affect the well being of a consumer. 148 Consumer Theory 5 Consumer Welfare Analysis (2) Microeconomics • Given a preference relation and Walrasian demand x(p, w), a price change from p0 to p1 increases the well-being of a consumer if indirect utility increases. I.e. v(p1 , w) > v(p0 , w). • Here we are interested in so called money metric indirect utility functions. E.g. expressing indirect utility in terms of monetary units. 149 Consumer Theory 5 Consumer Welfare Analysis (3) Microeconomics • Suppose u1 > u0, u1 = v(p1, w) arises from p1 , w and u0 = v(p0, w) from p0 , w. • With p fixed at ¯p, the property of the expenditure function that e(p, u) is increasing in u yields: e(¯p, u1)) = e(¯p, v(¯p, ˜w1 )) = ˜w1 > e(¯p, v(¯p, ˜w0 )) = e(¯p, u0) = ˜w0 - i.e. it is an indirect utility function which measures the degree of well-being in money terms. • See Figure 3.I.1, page 81. 150 Consumer Theory 5 Consumer Welfare Analysis (4) Microeconomics • Based on these considerations we set ¯p = p0 or p1 and w = e(p0 , u0 ) = e(p1 , u1 ); we define: – Definition - Equivalent Variation: “old prices” EV (p0 , p1 , w) = e(p0 , u1 )−e(p0 , u0 ) = e(p0 , u1 )−e(p1 , u1 ) = e(p0 , u1 )−w – Definition - Compensating Variation: “new prices” CV (p0 , p1 , w) = e(p1 , u1 )−e(p1 , u0 ) = e(p0 , u0 )−e(p1 , u0 ) = w−e(p1 , u0 ) 151 Consumer Theory 5 Consumer Welfare Analysis (5) Microeconomics • EV measures the money amount that a consumer is indifferent between accepting this amount and the status after the price change (i.e. to attain a utility level u1 ). • CV measures the money amount a consumer is willing to pay to induce the price change from p0 to p1 (i.e. to obtain utility level u0 at the new price p1 ). This money amount can be negative as well. • Discuss Figure 3.1.2, page 82; if p1 falls then the consumer is prepared to pay the amount CV , i.e. CV > 0. 152 Consumer Theory 5 Consumer Welfare Analysis (6) Microeconomics • Both measures are associated with Hicksian demand. • Suppose the only p1 changes, then p0 1 = p1 1 and p0 l = p1 l for l ≥ 2. With w = e(p0 , u0 ) = e(p1 , u1 ) and h1(p, u) = ∂e(p, u)/∂p1 we get EV (p0 , p1 , w) = p0 1 p1 1 h1((p1, p−), u1 )dp1 CV (p0 , p1 , w) = p0 1 p1 1 h1((p1, p−), u0 )dp1 153 Consumer Theory 5 Consumer Welfare Analysis (7) Microeconomics • Discuss these integrals - MWG, Figure 3.1.3, page 83. Here the following case is considered. p0 and p1 are L dimensional price vectors. Only the first component p1 is changed. The other prices p− := (p2, . . . , pL) are kept constant. w is constant as well. • EV, CV increase if utility increases and vice versa. • If x1 is a normal good, then the slope of the Walrasian demand function x1(p, w) is smaller than the slopes of h1(p, .) (in absolute terms). • We get EV (p0 , p1 , w) > CV (p0 , p1 , w) if the good is normal (in absolute value), the converse is true for inferior goods. • EV (p0 , p1 , w) = CV (p0 , p1 , w) with zero income effect for good 2. This is the case with quasilinear preferences for good two (see [D 3.B.7]). 154 Consumer Theory 5 Consumer Welfare Analysis (8) Microeconomics • EV (p0 , p1 , w) = CV (p0 , p1 , w) with zero income effect for good 1. In this case EV (p0 , p1 , w) = CV (p0 , p1 , w) is also equal to the change in Marshallian Consumer Surplus. • Definition - Marshallian Consumer Surplus: MCSl(p, w) = ∞ p xl((pl, p−), w)dpl • Definition - Area Variation: AV (p0 , p1 , w) = p0 l p1 l x(pl, p−, w)dpl. 155 Consumer Theory 5 Area Variation Measure (1) Microeconomics • Definition - Area Variation: AV (p0 , p1 , w) = p0 1 p1 1 x(p1, p−, w)dp1. • It measures the change in Marshallian consumer surplus. • If the income effect is zero this measure corresponds to EV and CV . (see Marshallian Consumer Surplus) • The argument that AV provides are good approximation of EV or CV can but need not hold. See MWG, Figure 3.1.8, page 90. Jehle/Reny, 1st edition, Theorem 6.3.2, page 278: Willing’s upper and lower bounds on the difference between CS and CV. 156 Consumer Theory 5 Partial Information (1) Microeconomics • Consider a bundle x0 , price vectors p0 , p1 and wealth w. Often a complete Walrasian demand function cannot be observed, however: • Theorem - Welfare and Partial Information I: Consider a consumer with complete, transitive, continuous, and locally non-satiated preferences. If (p1 − p0 ) · x0 < 0, then the consumer is strictly better of with (p1 , w) compared to (p0 , w). [P 3.I.1] 157 Consumer Theory 5 Partial Information (2) Microeconomics Proof: • With non-satiation the consumer chooses a set on the boundary of the budget set, such that p0 · x = w. Then p1 · x < w. • ⇒ x is affordable within the budget set under p1 . By the assumption of local non-satiation, there exists a closed set with distance ≤ ε including a better bundle which remains within the budget set. Then the consumer is strictly better off with p1 . 158 Consumer Theory 5 Partial Information (3) Microeconomics • What happens if (p1 − p0 ) · x0 > 0 ? This implies (αp1 + (1 − α)p0 − p0 ) · x0 > 0 for α > 0. • Theorem - Welfare and Partial Information II: Consider a consumer with a twice differentiable expenditure function. If (p1 − p0 ) · x0 > 0, then there exists an ¯α ∈ (0, 1) such that for all 0 < α ≤ ¯α, we have e((1 − α)p0 + αp1 ), u0 ) > w the consumer is strictly better off under p0 , w than under (1 − α)p0 + αp1 , w. [P 3.I.2] 159 Consumer Theory 5 Partial Information (4) Microeconomics Proof: • We want to show that CV is negative, if we move from p0 to p1 . Let pα = (1 − α)p0 + αp1 . We want to show that CV = e(p0 , u0 ) − e(pα , u0 ) < 0 for some ¯α ≥ α > 0. In other words e(pα , u0 ) − e(p0 , u0 ) > 0. • Taylor expand e(p, u) at p0 , u0 : e(pα , u0 ) = e(p0 , u0 ) + (pα − p0 ) pe(p0 , u0 ) + R(p0 , pα ) where R(p0 , pα )/||pα − p0 || → 0 if pα → p0 . e(., .) has to be at least C1 . (fulfilled since second derivatives are assumed to exist). 160 Consumer Theory 5 Partial Information (5) Microeconomics Proof: • By the properties of this approximation, there has to exist an ¯α, where the Lagrange residual can be neglected. Then sgn(e(pα , u0 ) − e(p0 , u0 )) = sgn (pα − p0 ) pe(p0 , u0 ) for all α ∈ [0, ¯α]. • This results in e(pα , u0 ) − e(p0 , u0 ) > 0 by the assumption that (pα − p0 ) pe(p0 , u0 ) > 0 and the fact that pe(p0 , u0 ) = h(p0 , u0 ) = x(p0 , e(p0 , u0 )). 161 Consumer Theory 5 Partial Information (6) Microeconomics • Remark: Note that with a differentiable expenditure function the second order term is non-positive, since the expenditure function is concave. • Remark: We can show the former theorem also in this way (differentiability assumptions have to hold in addition). There the non-positive second order term does not cause a problem, since there we wanted to show that e(p1 , u0 ) − e(p0 , u0 ) < 0 if (p1 − p0 ) · x0 < 0. 162 Consumer Theory 6 Aggregate Demand Microeconomics • Aggregate Demand • Aggregate Welfare • Aggregate Demand and the Weak Axiom • Existence of a Representative Consumer Literature: MWG, Chapter 4 163 Consumer Theory 6 Motivation (1) Microeconomics • We already know that individual demand can be expressed as a function of prices and the individual wealth level. • Can aggregate demand be expressed as a function of prices and the aggregate wealth level? 164 Consumer Theory 6 Motivation (2) Microeconomics • Individual demand derived from a rational preference relation satisfies the weak axiom of revealed preference. • Does aggregate demand satisfy the weak axiom? 165 Consumer Theory 6 Motivation (3) Microeconomics • Consider the welfare measures (CV,EV,AV). • When does aggregate demand have welfare significance. What is the meaning of welfare measures calculated from aggregate demand. 166 Consumer Theory 6 Aggregate Demand and Wealth (1) Microeconomics • Consider individual Walrasian demands xi(p, wi), i = 1, . . . , I. • Definition - Aggregate Demand: x(p, (w1, . . . , wI)) = I i=1 xi(p, wi). • Definition - Aggregate Wealth: w = I i=1 wi. • When is it possible to write aggregate demand in the simpler form x(p, i wi)? • Consider (w1, . . . , wI) and (w1, . . . , wI) with I i=1 wi = I i=1 wi. Are the demands x(p, i wi) and x(p, i wi) equal for arbitrary pairs of wealth levels with equal aggregate wealth? 167 Consumer Theory 6 Aggregate Demand and Wealth (2) Microeconomics • Start with (w1, . . . , wI) and the differential changes (dw1, . . . , dwI) such that i dwi = 0. • With a differentiable aggregate demand function, the requirement that aggregate demand does not change requires I i=1 ∂xli(p, wi) ∂wi dwi = 0. 168 Consumer Theory 6 Aggregate Demand and Wealth (3) Microeconomics • The assumption that (w1, . . . , wI) and (dw1, . . . , dwI) are arbitrary (with dwi = 0) implies: I i=1 ∂xli(p,wi) ∂wi dwi = 0 if and only if ∂xli(p, wi) ∂wi = ∂xlj(p, wj) ∂wj for every l, l = 1, . . . , L and i, j ∈ I, and every (w1, . . . , wI). • This condition implies that the wealth effect is the same for each consumer and each wealth level. 169 Consumer Theory 6 Aggregate Demand and Wealth (4) Microeconomics • This last issue can be satisfied if the wealth expansion paths are parallel straight lines. See MWG, Figure 4.B.1, page 107. • This is the case with homothetic & identical preferences or quasilinear preferences with respect to the same good (see MWG, D 3.B.6,7 at page 45). • More general result - next theorem. 170 Consumer Theory 6 Aggregate Demand and Wealth (5) Microeconomics • Proposition [P 4.B.1] A necessary and sufficient condition for the set of consumers to exhibit parallel straight wealth expansion paths at the price vector p is that preferences admit indirect utility functions of the Gorman form, with the coefficients on wi the same for all consumers i. That is vi(p, wi) = ai(p) + b(p)wi . • Proof: see exercise and reference in MWG, page 107, 123. 171 Consumer Theory 6 Aggregate Demand and Wealth (6) Microeconomics • Assumption on preferences is quite restrictive in MWG [P 4.B.1]. The requirements were quite general - we have considered arbitrary wealth distributions. • Simpler approach works via wealth distribution rules. • Definition - Wealth Distribution Rule: A family of functions (w1(p, w), . . . , wI(p, w)) assigning to each individual i a wealth level wi(p, w), fulfilling I i=1 wi = w, is called wealth distribution rule. • When we plug in wi(p, w) into xi(p, wi) we get Walrasian demands xi(p, wi(p, w)) which are functions of p and w. In this case aggregate demand must be a function of p and w. 172 Consumer Theory 6 Aggregate Demand & Weak Axiom (1) Microeconomics • Consider x(p, w1, . . . , wI) = I i=1 xi(p, wi). • Individual demand: Continuity, homogeneity of degree zero and Walras law. These properties directly carry over to x(p, w1, . . . , wI). • Individual demand derived from a rational preference relation satisfies the weak axiom (see Chapter 3). MWG ”arguably the most central positive property of the individual Walrasian demand” • Does aggregate demand satisfy the weak axiom? 173 Consumer Theory 6 Aggregate Demand & Weak Axiom (2) Microeconomics • Consider x(p, w1, . . . , wI) = I i=1 xi(p, wi) and a wealth distribution rule, such that aggregate demand can be written as x(p, w) = I i=1 xi(p, wi(p, w)) where w = I i=1 wi. • Assume that wi(p, w) = αiw in the following. I.e. the distribution rule is independent from prices. • Definition - Weak Axiom and Aggregate Demand:[D 4.C.1] The aggregate demand function x(p, w) satisfies the weak axiom if p · x(p , w ) ≤ w and x(p, w) = x(p , w ) imply p · x(p, w) > w for any (p, w) and (p , w ). 174 Consumer Theory 6 Aggregate Demand & Weak Axiom (3) Microeconomics • Counterexample MWG, [4.C.1], page 110: Individual demands fulfill weak axiom but aggregate does not. • Consider 2 consumers and 2 goods with w1 = w2 = w/2. For p and p , we get the demands x1(p, w/2), etc. fulfilling the weak axiom. Here x1(p, w) is weakly preferred to x1(p , w) and x2(p , w) is weakly preferred to x2(p, w), see Figure 4.C.1. • Aggregate demands are x1(p, w/2) + x2(p, w/2) and x1(p , w/2) + x2(p , w/2). A convex combination of the individual demands is within the budget hyperplains Bp,w and Bp ,w. Assume a mixture-weight of 1/2. 175 Consumer Theory 6 Aggregate Demand & Weak Axiom (4) Microeconomics • This implies that p · x1(p, w/2) + p · x2(p, w/2) = w and p · x1(p , w/2) + p · x2(p , w/2) = w. • But 1/2p · x(p , w) < w/2 and 1/2p · x(p, w) < w/2. Multiply both sides with 2, and we observe that x(p, w) does not satisfy the weak axiom of revealed preference. (p · x(p , w ) ≤ w requires p · x(p, w) > w for any (p, w) and (p , w ).) 176 Consumer Theory 6 Aggregate Demand & Weak Axiom (5) Microeconomics • For individual demands we know that the weak axiom holds if demand satisfies the compensated law of demand (see P 2.F.1): If for any (p, wi) and price change p the wealth wi = p · x(p, wi) and (p − p) [xi(p , wi) − xi(p, wi)] ≤ 0 with strict inequality for x(p, wi) = x(p , wi). • In the notation of this section this results in: αiw = p x(p, αiw) then (p − p) [xi(p , αiw ) − xi(p, αiwi)] ≤ 0 with strict inequality for xi(p, αiw) = xi(p , αiw ). 177 Consumer Theory 6 Aggregate Demand & Weak Axiom (6) Microeconomics • Problem: price-wealth change compensated in the aggregate need not be compensated individually. I.e. if w = p · x(p, w) this does not imply that αiw = p xi(p, αiw). Problems arises due to differences in the wealth effects. That is, although w is sufficient to buy the aggregate bundle x(p, w) with the new price vector p , this property need not hold on an individual level. • To get a condition that the weak axiom holds in the aggregate we consider non-compensated demands. 178 Consumer Theory 6 Aggregate Demand & Weak Axiom (7) Microeconomics • Definition - Uncompensated Law of Demand (ULD)[D 4.C.2] The individual demand function xi(p, wi) satisfies the uncompensated law of demand if (p − p) [xi(p , wi) − xi(p, wi)] ≤ 0 for any p, p and wi with strict inequality for x(p, wi) = x(p , wi). The analogous definition applies to the aggregate demand function. 179 Consumer Theory 6 Aggregate Demand & Weak Axiom (8) Microeconomics • Similar to the weak axiom we get the following results: If xi(p, wi) satisfies the ULD property then Dpxi(p, w) is negative semidefinite and vice versa. The same argument holds for aggregate demand. • The ULD property aggregates. • Proposition[P 4.C.1] If every consumers’ Walrasian demand function xi(p, wi) satisfies the uncompensated law of demand, so does aggregate demand x(p, w) = I i=1 xi(p, αiw). As a consequence the aggregate demand function satisfies the weak axiom. 180 Consumer Theory 6 Aggregate Demand & Weak Axiom (9) Microeconomics Proof: • Step 1: Show that aggregate demand satisfies ULD if individual demands do. • We start with (p, w), (p , w ) and xi(p, αiw) = xi(p , αiw). • By ULD we get (p − p) [xi(p , αiw) − xi(p, αiw)] < 0. Summing up over i, results in (p − p) [x(p , w) − x(p, w)] < 0. This holds for all p, p and w. 181 Consumer Theory 6 Aggregate Demand & Weak Axiom (10) Microeconomics Proof: • Step 2: Show that the WA holds, i.e. if p · x(p , w ) ≤ w then p · x(p, w) > w . • To show that ULD for aggregate demand implies the weak axiom we take (p, w), (p , w ), x(p, w) = x(p , w ) and p · x(p , w ) ≤ w. Define p = (w/w )p . Since demand is homogeneous of degree zero in p, w we get x(p , w) = x(p , w ). • From (p − p) [x(p , w) − x(p, w)] < 0, p · x(p , w) ≤ w and Walras’ law if follows that p · x(p, w) > w. By the definition of p this implies p · x(p, w) > w such the weak axiom is satisfied. 182 Consumer Theory 6 Aggregate Demand & Weak Axiom (11) Microeconomics • How restrictive is Proposition 4.C.1? • Textbook provides sufficient conditions where ULD is satisfied. • Proposition[P 4.C.2] If i is homothetic, then xi(p, wi) satisfies the uncompensated law of demand property. • In the homothetic case the income effects are well behaved. This is the only case that does so. (Here we can write demands as xi(p, wi) = f(p)wi. By the Euler theorem ∂xi(p,wi) ∂wi wi = xil(p, wi), etc.) 183 Consumer Theory 6 Aggregate Demand & Weak Axiom (12) Microeconomics • ULD requires that the substitution effect is strong compared to the income effect. • Proposition[P 4.C.3] Suppose that i is defined on the consumption set X = RL + and is representable by a twice continuously differentiable concave function ui(.). If − xi · D2 ui(xi) · xi xi xui(xi) < 4 for all xi, then xi(p, wi) satisfies the uncompensated law of demand property. 184 Consumer Theory 6 Representative Consumer (1) Microeconomics • When does the computation of an aggregate welfare measure make sense? ⇒ normative representative consumer. • When can we treat the aggregate demand function as if it were generated by a representative consumer whose preferences can be used to measure social welfare? • When can we construct a fictional individual whose utility maximization problem given the society’s budget set results in aggregate demand? ⇒ positive representative consumer. • We shall observe that a normative representative consumer need not exist even if a positive representative consumer exists. 185 Consumer Theory 6 Representative Consumer (2) Microeconomics • When does the computation of an aggregate welfare measure make sense? ⇒ normative representative consumer. • We start with individual demands xi(p, wi) and a wealth distribution rule (w1(p, w), . . . , wI(p, w)) with wi(p, w) homogeneous of degree one in w and continuous. • Definition - Positive Representative Consumer[D 4.D.1] A positive representative consumer exists if there is a rational preference relation on RL + such that the aggregate demand function x(p, w) is the Walrasian demand function generated by this preference relation (max u(x) s.t. p · x(p, w) = w). That is x(p, w) x whenever x = x(p, w) and p · x ≤ w. 186 Consumer Theory 6 Representative Consumer (3) Microeconomics • Existence: The aggregate demand function satisfies Walras’ law, homogeneity of degree zero and symmetric and negative definite Slutsky matrix. By the integrability theorem there exists a utility function u(x) generating x(p, w), where u(x) goes back to a rational, continuous and monotone preference relation. • ⇒ a positive representative consumer exists under fairly mild conditions. 187 Consumer Theory 6 Representative Consumer (4) Microeconomics • For welfare comparisons this question becomes more difficult. • Definition - Bergson-Samuelson Social Welfare Function [D 4.D.2] A Bergson-Samuelson social welfare function W : RI → R assigns an utility value to each possible vector of utility levels (u1, . . . , uI) for the I consumers in the economy. • Expresses the society’s judgment how individual utilities have to be compared to produce an ordering of possible social outcomes. • Assume that W is strictly increasing, strictly concave and differentiable. ui(.) is strictly concave. 188 Consumer Theory 6 Representative Consumer (5) Microeconomics • Definition - Welfare Maximization Problem (WMP): Suppose that there is a benevolent social planner choosing a wealth distribution (w1(p, w), . . . , wI(p, w)) for each (p, w), such that W is maximized. I.e. maxw1,...,wI W(v1(p, w1), . . . , vI(p, wI)) s.t. I i=1 wi ≤ w • Under the above assumptions a solution of this problem exists. The maximized value is denoted by v(p, w). • The next proposition shows that this indirect utility function provides us with a positive representative consumer. 189 Consumer Theory 6 Representative Consumer (6) Microeconomics • Proposition[P 4.D.1] Suppose that for each p, w the wealth distribution rule (w1(p, w), . . . , wI(p, w)) solves the maximization problem WLP. The optimum value function v(p, w) is an indirect utility function of a positive representative consumer for the aggregate demand function x(p, w) = i xi(p, wi(p, w)) 190 Consumer Theory 6 Representative Consumer (7) Microeconomics Proof: • By Berge’s theorem of the maximum a continuous value function exists. • Since the wealth distribution rule, ui(.) and W(u1, . . . , uI) are differentiable, and v(p, w) is a composition of these functions, v(p, w) has to be differentiable. • Idea of the proof: We use Roy’s identity to derive xR(p, w) from the indirect utility function v(p, w). Then we establish that xR(p, u) is equal to aggregate demand x(p, w). 191 Consumer Theory 6 Representative Consumer (8) Microeconomics Proof: • Consider the optimization problem: maxw1,...,wI W(v1(p, w1), . . . , vI(p, wI)) s.t. I i=1 wi ≤ w with Lagrangeian L(w1, . . . , wI) = W(v1(p, w1), . . . , vI(p, wI)) + λ(w − i wi) 192 Consumer Theory 6 Representative Consumer (9) Microeconomics Proof: • From the first order conditions we get: λ = ∂W ∂v1 ∂v1 ∂w1 = · · · = ∂W ∂vI ∂vI ∂wI . • Since v(p, w) is a value function, the envelope theorem applied to the social indirect utility function v(p, w) results in λ = ∂v(p, w) ∂w . 193 Consumer Theory 6 Representative Consumer (10) Microeconomics Proof: • We already know that Roy’s identity takes us back from indirect utility to demand: xR(p, w) = − pv(p, w) wv(p, w) • In addition the envelope theorem applied to the social indirect utility function v(p, w) results in: ∂v(p, w) ∂pl = I i=1 ∂W (.) ∂vi ∂vi(p, w) ∂pl + λ ∂wi ∂pl = I i=1 ∂W (.) ∂vi ∂vi(p, w) ∂pl . for l = 1, . . . , L. 194 Consumer Theory 6 Representative Consumer (11) Microeconomics Proof: • The second term is zero because i wi(p, w) = w and Walras’ law holds. • In matrix notation this expression results in: pv(p, w) = I i=1 ∂W(.) ∂vi pvi(p, w). • The above FOC implies that pv(p, w) = I i=1 ∂W (.) ∂vi pvi(p, w) = I i=1 λ (∂vi/∂wi) pvi(p, w). 195 Consumer Theory 6 Representative Consumer (12) Microeconomics Proof: • The first order condition, xR, the fact that ∂v(p,w) ∂w = λ, the wealth distribution rule wi(p, w) and Roy’s identity (for individual demands xi(p, wi) = − 1 ∂vi/∂wi pvi(p, wi)) yields: xR(p, w) = − 1 λ i λ ∂vi/∂wi pvi(p, wi(p, w)) = i xi(p, wi(p, w)) = x(p, w) . 196 Consumer Theory 6 Representative Consumer (13) Microeconomics • Definition - Normative Representative Consumer[D 4.D.3] The positive representative consumer for the aggregate demand x(p, w) is a normative representative consumer relative to the social welfare function W(.) if for every (p, w) the distribution of wealth (w1(p, w), . . . , wI(p, w)) solves the welfare maximization problem (WMP) and therefore the maximum value function of this problem is an indirect utility function for . 197 Consumer Theory 6 Representative Consumer (14) Microeconomics • If a normative representative consumer exists then welfare comparisons based on aggregate demand make sense. • Example where positive and normative representative consumer exist - [E 4.D.1], page 118, where individual preferences are homothetic and W(u1, . . . , u2) = αi log ui, αi > 0 and i αi = 1. • If aggregate demand can be represented by a representative consumer this does not imply the existence of a normative representative consumer. This depends on W(.) and whether this function is maximized by the corresponding wealth distribution rule. Note that we have assumed this in proposition P 4.D.1. For a counterexample see Example page 120 (in small print). 198 Production 1 Motivation Microeconomics • Production • Production possibility sets and the production function • Marginal product, marginal rate of substitution and returns to scale. MasColell, Chapter 5 199 Production 1 Firms (1) Microeconomics • In this section we treat the firm as a black box. We abstract from ownership, management, organization, etc. • Assumption: A firm maximizes its profit. • How can we justify this assumption? 200 Production 1 Production Possibility Set (1) Microeconomics • Definition - Production: The process of transforming inputs to outputs is called production. • The state of technology restricts what is possible in combining inputs to produce output (technological feasibility). • Definition - Production Possibility Set: A set Y ∈ RL describing possible production plans is called production possibility set, Y = {y ∈ RL | y is a feasible production plan}. yi < 0 are called inputs, yi > 0 outputs. 201 Production 1 Production Possibility Set (2) Microeconomics • Often the production possibility set is described by a function F(.) called transformation function. This function has the property Y = {y ∈ RL |F(y) ≤ 0} and F(y) = 0 if and only if we are on the boundary of the set Y . {y ∈ RL |F(y) = 0} is called transformation frontier. • Definition - Marginal Rate of Transformation: If F(.) is differentiable and F(¯y) = 0, then for commodities k and l the ration MRTlk(¯y) = ∂F(¯y)/∂yl ∂F(¯y)/∂yk is called marginal rate of transformation of good l for good k. 202 Production 1 Production Possibility Set (3) Microeconomics • If l and k are outputs we observe how output of l increases if k is decreases. • With inputs .... In this case the marginal rate of transformation is called marginal rate of technical substitution. • With a single output q, production is often described by means of a production function q = f(z1, . . . , zm), where the inputs zi ≥ 0, i = 1, . . . , m. In this case Y = {(−z1, . . . , −zm, q) |q − f(z1, . . . , zm) ≤ 0 and z1, . . . , zm ≥ 0}. 203 Production 1 Production Possibility Set (4) Microeconomics • Assumption and Properties of production possibility sets P1 Y is non-empty. P2 Y is closed. I.e. Y includes its boundary, if yn ∈ Y converges to y then y ∈ Y . P3 No free lunch. If yl ≥ 0 for l = 1, . . . , L, then y = 0. It is not possible to produce something from nothing. Therefore Y ∩ RL + = 0 ∈ Y (note that 0 ∈ Y has to be assumed here). See Figure 5.B.2, page 131. 204 Production 1 Production Possibility Set (5) Microeconomics P4 Possibility of inaction: 0 ∈ Y . This assumption hold at least ex-ante, before the setup of the firm. If we have entered into some irrevocable contracts, then a sunk cost might arise. P5 Free Disposal: New inputs can be acquired without any reduction of output. If y ∈ Y and y ≤ y then y ∈ Y . For any y ∈ Y and x ∈ RL +, we get y − x ∈ Y . See Figure 5.B.4, page 132. P6 Irreversibility: If y ∈ Y and y = 0, then −y /∈ Y . It is impossible to reverse a possible production vector. We do not come from output to input. 205 Production 1 Production Possibility Set (6) Microeconomics P7 Nonincreasing returns to scale: If y ∈ Y , then αy ∈ Y for all α ∈ [0, 1]. I.e. any feasible input-output vector y can be scaled down. See Figure 5.B.5. P8 Nondecreasing returns to scale: If y ∈ Y , then αy ∈ Y for any scale α ≥ 1. I.e. any feasible input-output vector y can be scaled up. See Figure 5.B.6. P9 Constant returns to scale: If y ∈ Y , then αy ∈ Y for any scale α ≥ 0. I.e. any feasible input-output vector y can be scaled up and down. 206 Production 1 Production Possibility Set (7) Microeconomics P10 Additivity - free entry: If y ∈ Y and y ∈ Y , then y + y ∈ Y . This implies that ky ∈ Y for any positive integer k. • Example: Output is an integer. If y and y are possible, additivity means that y + y is still possible and the production of y has no impact on y and vice versa. E.g. we have two independent plants. • As regards free-entry: If the aggregate production set Y is additive, then unrestricted entry is possible. To see this, if y ∈ Y is produced by firm A and y ∈ Y by firm B, then y + y ∈ Y if additivity holds. That is, the production plans of firm A do not interfere with the production plans of firm B (and vice versa). In other words, the aggregate production set has to satisfy additivity whenever unrestricted entry is possible. 207 Production 1 Production Possibility Set (8) Microeconomics P11 Convexity: Y is a convex set. I.e. if y ∈ Y and y ∈ Y , then αy + (1 − α)y ∈ Y . • Convexity implies nonincreasing returns to scale. • We do not increase productivity by using unbalanced input combinations. If y and y produce the same output, then a convex combination of the correspond inputs must at least produce an output larger or equal to the output with y and y . 208 Production 1 Production Possibility Set (9) Microeconomics P12 Y is convex cone: Y is a convex cone if for any y, y ∈ Y and constants α, β ≥ 0, αy + βy ∈ Y . Conjunction between convexity and constant returns to scale property. 209 Production 1 Production Possibility Set (10) Microeconomics • Proposition[P 5.B.1]: The production set Y is additive and satisfies the nonincreasing returns to scale property if and only if it is is convex cone. 210 Production 1 Production Possibility Set (11) Microeconomics Proof: • If Y is a convex cone then Y is additive and satisfies the nonincreasing returns to scale by the definition of a convex cone. • We have to show that with additivity and nonincreasing returns to scale we get αy + βy ∈ Y for any y, y and α, β > 0 (note that with α ≥ 0, β = 0, α = 0, β ≥ 0 the relation αy + βy ∈ Y follows from RTS and additivity): Let γ = max{α, β} > 0. By additivity γy ∈ Y and γy ∈ Y . • α/γ and β/γ are ≤ 1. Due to nonincreasing returns to scale αy = (α/γ)γy and βy = (β/γ)γy ∈ Y . By additivity αy + βy ∈ Y . 211 Production 1 Production Possibility Set (12) Microeconomics • Proposition[P 5.B.2]: For any convex production set Y ⊂ RL with 0 ∈ Y , there is a constant returns to scale convex production set Y ∈ RL+1 such that Y = {y ∈ RL |(y, −1) ∈ Y }. Y is called extended production set. • Proof: Let Y = {y ∈ RL+1 |y = α(y, −1), y ∈ Y, α ≥ 0}. If y ∈ Y , then the first L components are in Y by construction. Since βy1 + (1 − β)y2 ∈ Y we get β(y1 , −1) + (1 − β)(y2 , −1) ∈ Y . αy ∈ Y by construction. 212 Production 1 Production Function (1) Microeconomics • Often it is sufficient to work with an output q ≥ 0 and inputs z = (z1, . . . , zm) where zi ≥ 0. • Definition - Production Function: A function describing the the relationship between q and z is called production function f. • Remark: The production functions assigns the maximum of output q that can be attained to an input vector z. f(z) = max{q ≥ 0|z ∈ Rm + }; (output efficient production). 213 Production 1 Production Function (2) Microeconomics • Assumption PF on Production Function: The production function f : Rm + → R+ is continuous, strictly increasing and strictly quasiconcave on Rm + ; f(0) = 0. • Assumption PF’ - Production Function: The production function f : Rm + → R+ is continuous, increasing and quasiconcave on Rm + ; f(0) = 0. • How can we motivate these assumptions? 214 Production 1 Production Function (3) Microeconomics • Considering production functions two approaches are common: (i) variation one factor, (ii) variation all factors in the same proportion. • Definition - Marginal Product: If f is differentiable then ∂f(z) ∂zi = MPi(z) is called marginal product of the input factor zi. • By Assumption P5 all marginal products are strictly larger than zero, with P5’ MPi(z) ≥ 0. • Definition - Average Product: The fraction f(z)/zi = APi(z) is called average product of the input factor zi. 215 Production 1 Production Function (4) Microeconomics • Definition - Isoquant: The set Q(q) where output is constant is called q-level isoquant. I.e. Q(q) = {z ≥ 0|f(z) = q}. • In addition to Q(q) we can define the the contour set ¯S(q) = {z ≥ 0|f(z) ≥ q}. Since f is quasiconcave, this set is convex ⇒ isoquants are convex curves. 216 Production 1 Production Function (5) Microeconomics • In addition, by means of the isoquant we can observe how input factors can be substituted to remain on the same level of output. • Definition - Marginal Rate of Technical Substitution: MRTSij(z) = MPi MPj • The slope of the isoquant is given by − dzj dzi = MPi MPj • Discuss: MPi MPj > 0 (≥ 0) and the concept of technical efficiency: To remain on the same level of output at least one input has to be increased if one input factor has been decreased. 217 Production 1 Production Function (6) Microeconomics • In general the MRTS of two input depends on all other inputs (note that the MPi depends on z). • In applied work it is often assumed that inputs can be classified, such that the MRTS within a class is not affected by inputs outside this class. 218 Production 1 Production Function (7) Microeconomics • Definition - Separable Production Function: Suppose that the inputs can be partitioned into S > 1 classes N1, . . . , NS; N = {1, . . . , n} is an index set. The production function is called weakly separable if the MRTS between inputs within the same group is independent of the inputs used in the other groups: ∂(MPi/MPj) ∂zk = 0 for all i, j ∈ Ns and k /∈ Ns. For S > 2 it is strongly separable if the MRTS between two inputs from different groups is independent of all inputs outside those groups: ∂(MPi/MPj) ∂zk = 0 for all i ∈ Ns, j ∈ Nt and k /∈ Ns ∪ Nt. 219 Production 1 Production Function (8) Microeconomics • Since MRTSij is sensitive to the dimension of the measurements of zi and zj an elasticity can be used. • Definition - Elasticity of Substitution: For a differentiable production function the elasticity of substitution between inputs zi and zj is defined by σij := d(zj/zi) d(MPi/MPj) (MPi/MPj) (zj/zi) = d log(zj/zi) d log(MPi/MPj) . • With a quasiconcave production function σij ≥ 0 220 Production 1 Production Function (9) Microeconomics • Theorem - Linear Homogeneous Production Functions are Concave: Let f satisfy Assumption P5’. If f is homogenous of degree one, then f(z) is concave in z. 221 Production 1 Production Function (10) Microeconomics Proof: • We have to show f(zν ) ≥ νf(z1 ) + (1 − ν)f(z2 ), where zν = νz1 + (1 − ν)z2 . • Step 1: By assumption f(µz) = µf(z) = µy. Then 1 = f(z/y). I.e. f(z1 /y1 ) = f(z2 /y2 ) = 1. (Set µ = 1/y.) • Since f(z) is quasiconcave: f(zν ) ≥ min{f(z1 ), f(z2 )}. • Therefore f ν(z1 /y1 ) + (1 − ν)(z2 /y2 ) ≥ 1. 222 Production 1 Production Function (11) Microeconomics Proof: • Choose ν∗ = y1 /(y1 + y2 ). Then f (z1 + z2 )/(y1 + y2 ) ≥ 1. • By the homogeneity of f we derive: f z1 + z2 ≥ y1 + y2 = f(z1 ) + f(z2 ) . 223 Production 1 Production Function (12) Microeconomics Proof: • Step 2: Now we show that f(zν ) ≥ νf(z1 ) + (1 − ν)f(z2 ) holds. • By homogeneity f(νz1 ) = νf(z1 ) and f((1 − ν)z2 ) = (1 − ν)f(z2 ) • Insert into the above expressions: f νz1 + (1 − ν)z2 ≥ f(νz1 ) + f((1 − ν)z2 ) f(νz1 ) + f((1 − ν)z2 ) = νf(z1 ) + (1 − ν)f(z2 ) 224 Production 1 Production Function (13) Microeconomics • Another way to look at the properties of production is to alter inputs proportionally. I.e. zi/zj remains constant. • Discuss: This analysis is of interest especially for the long run behavior of a firm. 225 Production 1 Production Function (14) Microeconomics • Definition - Returns to Scale. A production function f(z) exhibits – Constant returns to scale if f(µz) = µf(z) for µ > 0 and all z. – Increasing returns to scale if f(µz) > µf(z) for µ > 1 and all z. – Decreasing returns to scale if f(µz) < µf(z) for µ > 1 and all z. 226 Production 1 Production Function (15) Microeconomics • With constant returns the scale the production function has to be homogeneous of degree one. • Homogeneity larger than one is sufficient for increasing returns to scale but not necessary. • Most production function/technologies often exhibit regions with constant, increasing and decreasing returns to scale. 227 Production 1 Production Function (16) Microeconomics • Definition - Local Returns to Scale. The elasticity of scale at z is defined by LRTS(z) := lim µ→1 d log(f(µz)) d log µ = n i=1 MPizi f(z) . A production function f(z) exhibits – local constant returns to scale if LRTS(z) is equal to one. – local increasing returns to scale if LRTS(z) is larger than one. – local decreasing returns to scale if LRTS(z) is smaller than one. 228 Production 2 Profits and Cost (1) Microeconomics • Profit Maximization • Cost minimization • Price taking • Cost, profit and supply function MasColell, Chapter 5.C 229 Production 2 Profits (1) Microeconomics • Assume that p = (p1, . . . , pL) are larger than zero and fixed (price taking assumption). • We assume that firms maximize profits. • Given an Input-Output vector y, the profit generated by a firm is p · y. • We assume that Y is non-empty, closed and free disposal holds. 230 Production 2 Profits (2) Microeconomics • Definition: Given the production possibility set Y , we get the profit maximization problem max y p · y s.t. y ∈ Y. • If Y can be described by a transformation function F, this problem reads as follows: max y p · y s.t. F(y) ≤ 0. • Define π(p) = supy p · y s.t. y ∈ Y . 231 Production 2 Profits (3) Microeconomics • Definition - Profit function π(p): The maximum value function associated with the profit maximization problem is called profit function. The firm’s supply correspondence y(p) is the set of profit maximizing vectors {y ∈ Y |p · y = π(p)}. • The value function π(p) is defined on extended real numbers (¯R = R ∪ {−∞, +∞}). The set Sp = {p · y|y ∈ Y } is a subset of R. {p · y|y ∈ Y } has an upper bound in ¯R. For p where Sp is unbounded (from above) in R we set π(p) = ∞. • If Y is compact a solution (and also the max) for the profit maximization problem exits. If this is not the case π(p) = ∞ is still possible. The profit function exists by Bergs theorem of the maximum if the constraint correspondence is continuous. • We follow MWG and write maxy p · y s.t. y ∈ Y , although ....; Jehle/Reny call π(p, w) well defined if π(p, w) < ∞. 232 Production 2 Profits (4) Microeconomics • Suppose that F(.) is differentiable, then we can formulate the profit maximization problem as a Kuhn-Tucker problem: • The Lagrangian is given by: L(y, λ) = p · y − λF(y) • Then the Kuhn-Tucker conditions are given by: ∂L ∂yl = pl − λ ∂F(y) ∂yl ≤ 0 , ∂L ∂yl yl = 0 ∂L ∂λ = −F(y) ≥ 0 ∂L ∂λ λ = 0 233 Production 2 Profits (5) Microeconomics • For those inputs and output different from zero we get: p = λ yF(y) This implies that pl pk = ∂F/∂yl ∂F/∂yk = MRTlk. • Since the left hand side is positive by assumption, the fraction of the right hand side and λ have to be positive. 234 Production 2 Profits (6) Microeconomics • If yl, yk > 0, i.e. both goods are outputs, then yl, yk have to be chosen such that the fraction of marginal rates of transformation is equal to the ratio of prices. • If yl, yk < 0, i.e. both goods are inputs, then yl, yk have to be chosen such that the fraction of marginal rates of transformation (= marginal rate of technical substitution) is equal to the ratio of prices. • If yl > 0, yk < 0, i.e. yl is an output and yk is an input, then pl = ∂F/∂yl ∂F/∂yk pk. Later on we shall observe that ∂F/∂yl ∂F/∂yk pk is the marginal cost of good l. See Figure 5.C.1. page 136. 235 Production 2 Profits - Single Output Case (1) Microeconomics • Suppose the there is only one output q ≥ 0 and input z ≥ 0. The relationship between q and z is described by a differentiable production function. The price of q is p > 0. Input factor prices are w 0. We assume that the second order conditions are met. • The profit maximization problem now reads as follows: π(p, w) := {max z,q≥0 pf(z) − w · z s.t. f(z) ≥ q} • The input factor demand arising from this problem z = z(w, q) is called input factor demand. 236 Production 2 Profits - Single Output Case (2) Microeconomics • Is the profit function well defined? • Suppose that pf(z) − w · z ≥ 0. What happens if f(z) exhibits increasing returns to scale? • Here pf(µz) − w · µz > pµf(z) − w · µz for all µ > 1. • That is, the profit can always be increased when increasing µ. • With constant returns to scale no problem arises when π(w, p) = 0. Then pf(µz) − w · µz = pµf(z) − w · µz = 0 for all µ. 237 Production 2 Profits - Single Output Case (3) Microeconomics • From these remarks we get the (long run) problem: max{pq − w · z} s.t f(z) ≥ q • The Lagrangian is now given by: L(q, z, λ) = pq − w · z + λ(f(z) − q) • The marginal product will be abbreviated by MPi = ∂f(z) ∂zi . 238 Production 2 Profits - Single Output Case (4) Microeconomics • Then the Kuhn-Tucker conditions are given by: ∂L ∂y = p + λ ≤ 0 , ∂L ∂q q = 0 ∂L ∂zi = −wi − λMPi ≤ 0 , ∂L ∂zi zi = 0 ∂L ∂λ = f(z) − q ≥ 0 , ∂L ∂λ λ = 0 239 Production 2 Profits - Single Output Case (5) Microeconomics • This yields: wi = p ∂f(z) ∂zi , ∀zi > 0 • Definition - Marginal Revenue Product: p∂f(z) ∂zi . • For inputs i and j we derive: ∂f(z)/∂zi ∂f(z)/∂zj = wi wj 240 Production 2 Profit Function (1) Microeconomics • By means of π(p) we can reconstruct −Y , if −Y is a convex set. • That is to say: π(p) follows from {maxy p · y s.t. y ∈ Y }, which is equivalent to {miny − p · y s.t. y ∈ Y } and {min−y p · (−y) s.t. (−y) ∈ −Y }. • Remember the concept of a support function: By means of the support function µX(p) we get by means of {x|p · x ≥ µX(p)} a dual representation of the closed and convex set X. • Here −π(p) = µ−Y (p) where µ−Y (p) = miny{p · (−y)|y ∈ Y } such that −π(p) is a support function of −Y . 241 Production 2 Profit Function (2) Microeconomics • Proposition: [5.C.1] Suppose that π(p) is the profit function of the production set Y and y(p) is the associated supply correspondence. Assume that Y is closed and satisfies the the free disposal property. Then 1. π(p) is homogeneous of degree one. 2. π(p) is convex. 3. If Y is convex, then Y = {y ∈ RL |p · y ≤ π(p) , ∀p 0} 4. y(p) is homogeneous of degree zero. 5. If Y is convex, then y(p) is convex for all p. If Y is strictly convex, then y(p) is single valued. 6. Hotelling’s Lemma: If y(¯p) consists of a single point, then π(p) is differentiable at ¯p and π(¯p) = y(¯p). 7. If y is differentiable at ¯p, then Dy(¯p) = D2 π(¯p) is a symmetric and positive semidefinite matrix with Dy(¯p)¯p = 0. 242 Production 2 Profit Function (3) Microeconomics Proof: • π(p) is homogeneous of degree one and y(p) is homogeneous of degree zero follow from the structure of the optimization problem. If y ∈ y(p) solves {max p · y s.t. F(y) ≤ 0} then it also solves α{max p · y s.t. F(y) ≤ 0} and {max αp · y s.t. F(y) ≤ 0}, such that y ∈ y(αp) for any α > 0. • This hold for every y ∈ y(p) ⇒ y(p) is homogeneous of degree zero and π(p) is homogeneous of degree one by the structure of the profit equation. 243 Production 2 Profit Function (4) Microeconomics Proof: • π(p) is convex: Consider p1 and p2 and the convex combination pν . y1 , y2 and yν are arbitrary elements of the optimal supply correspondences. • We get p1 y1 ≥ p1 yν and p2 y2 ≥ p2 yν • Multiplying the first term with ν and the second with 1 − ν, where ν ∈ [0, 1] results in νp1 y1 + (1 − ν)p2 y2 ≥ νp1 yν + (1 − ν)p2 yν = pν yν which implies νπ(p1 ) + (1 − ν)π(p2 ) ≥ π(pν ) 244 Production 2 Profit Function (5) Microeconomics Proof: • If Y is convex then Y = {y ∈ RL |p · y ≤ π(p) , for all p 0}: If Y is convex, closed and free disposal holds, then π(p) provides a dual description of the production possibility set. 245 Production 2 Profit Function (6) Microeconomics Proof: • If Y is convex then y(p) is a convex set, with strict convexity y(p) is a function: If Y is convex then yν = νy1 + (1 − ν)y2 ∈ Y . • If y1 and y2 solve the PMP for p, then π(p) = p · y1 = p · y2 . A rescaling of the production vectors has to result in yν = νy1 + (1 − ν)y2 where p · yν = π(p) has to hold. This follows from p · y1 = p · y2 = π(p)= νπ(p) + (1 − ν)π(p)= νp·y1 +(1−ν)p·y2 = pν ·y1 +p(1−ν)·y2 = p(ν ·y1 +(1−ν)·y2 ). 246 Production 2 Profit Function (7) Microeconomics Proof: • Suppose that yα solves the PMP and Y is strictly convex (every point on the boundary is an extreme point, i.e. this point is not a convex combination of other points in Y ). yα is an element of Y ∩ H(p, π(p)). H(p, π(p)) stands for an isoprofit hyperplane. Suppose that there is another solution y solving the profit maximization problem (PMP). So y, y are elements of this hyperplane. Since y, y ∈ Y this implies that Y cannot be strictly convex. • Remark by Proposition P 5.F.1, page 150, y(p) cannot be an interior point of y. Suppose that an interior point y solves the PMP, then π(p) = p · y . For any interior point, there is an y such that y ≥ y and y = y . Since p 0 this implies p · y > p · y such that an interior point cannot be optimal. 247 Production 2 Profit Function (8) Microeconomics Proof: • Hotellings lemma: Follows directly from the duality theorem: pπ(¯p) = y(¯p); (see [P 3.F.1], page 66). 248 Production 2 Profit Function (9) Microeconomics Proof: • Property 7: If y(p) and π are differentiable, then Dpy(¯p) = D2 pπ(p). By Young’s theorem this matrix is symmetric, since π(p) is convex in p the matrix has to be positive semidefinite (see Theorem M.C.2, p.. 932). • Dpy(p)p = 0L×1 follows from the Euler theorem (see MWG, Theorem M.B.2, p. 929). 249 Production 2 Profit Function (10) Microeconomics • By Hotellings lemma inputs and outputs react in the same direction as the price change: Output increases if output prices increase, while inputs decrease if its prices increase (law of supply), i.e.: (p − p )[y(p) − y(p )] ≥ 0 • This law holds for any price change (there is no budget constraint, therefore any form of compensation is not necessary. We have no wealth effect but only substitution effects). • We can also show that the law of supply holds also for the non-differentiable case. (We know that p1 y1 ≥ p1 y for any y1 ∈ y(p1 ) and p2 y2 ≥ p2 y for any y2 ∈ y(p1 ), sum up ....) 250 Production 2 Cost Function (1) Microeconomics • Profit maximization implies cost minimization! • Production does not tell us anything about the minimal cost to get output. • On the other hand side - if the firm is not a price taker in the output market, we cannot use the profit function, however the results on the cost function are still valid. • With increasing returns to scale where the profit function can only take the values 0 or +∞, the cost function is better behaved since the output is kept fixed there. 251 Production 2 Cost Function (2) Microeconomics • Assume that the input factor prices w 0 are constant. In addition we assume that the production function is at least continuous. • Definition - Cost: Expenditures to acquire input factors z to produce output q; i.e. w · z. • Definition - Cost Minimization Problem (CMP): minz w · z s.t. f(z) ≥ q. The minimal value function C(w, q) is called cost function. The optimal input factor choices are called conditional factor demand correspondence z(w, q). 252 Production 2 Cost Function (3) Microeconomics • Existence: Construct the set {z|f(z) ≥ q}. Under the usual assumptions on the production function the set is closed. By compactifying this set by means of {z|f(z) ≥ q, zi ≤ w · ¯z/wi} for some ¯z with f(¯z) = q we can apply the Weierstraß theorem. • By Berge’s theorem of the maximum we get a continuous cost function C(w, q), if the constraint correspondence is continuous. 253 Production 2 Cost Function (4) Microeconomics • Definition - Marginal Cost: LMC(q) = ∂c(w,q) ∂q is called marginal cost. • Definition - Average Cost: LAC(q) = c(w,q) q is called average cost. 254 Production 2 Cost Function (5) Microeconomics • Theorem: Properties of the Cost Function c(w, q): [P 5.C.2] Suppose that c(w, q) is a cost function of a single output technology Y with production function f(z) and z(w, q) is the associated conditional factor demand correspondence. Assume that Y is closed and satisfies the free disposal property. Then (i) c(w, q) is homogeneous of degree one in w and nondecreasing in q. (ii) Concave in w. (iii) If the set {z ≥ 0|f(z) ≥ q} is convex for every q, then Y = {(−z, q)|w · z ≥ c(w, q) for all w 0}. (iv) z(w, q) is homogeneous of degree zero in w. (v) If the set {z ≥ 0|f(z) ≥ q} is convex then z(w, q) is a convex set, with strict convexity z(w, q) is a function. 255 Production 2 Cost Function (6) Microeconomics • Theorem: Properties of the Cost Function c(w, q): [P 5.C.2] Suppose that c(w, q) is a cost function of a single output technology Y with production function f(z) and z(w, q) is the associated conditional factor demand correspondence. Assume that Y is closed and satisfies the free disposal property. Then (vi) Shepard’s lemma: If z( ¯w, q) consists of a single point, then c(.) is differentiable with respect to w at ¯w and wc( ¯w, q) = z( ¯w, q). (vii) If z(.) is differentiable at ¯w then Dwz( ¯w, q) = D2 c( ¯w, q) is symmetric and negative semidefinite with Dwz( ¯w, q) ¯w = 0. (viii) If f(.) is homogeneous of degree one, then c(.) and z(.) are homogeneous of degree one in q. (ix) If f(.) is concave, then c(.) is a convex function of q (marginal costs are nondecreasing in q). 256 Production 2 Cost Function (7) Microeconomics • By means of the cost function we can restate the profit maximization problem (PMP): max q≥0 pq − C(w, q) • The first order condition becomes: p − ∂C(w, q) ∂q ≤ 0 with (p − ∂C(w,q) ∂q ) = 0 if q > 0. 257 Production 3 Aggregate Supply and Efficiency Microeconomics • Aggregate Supply • Joint profit maximization is a result of individual profit maximization • Efficient Production Mas-Colell Chapters 5.D, 5.E 258 Production 3 Aggregate Supply (1) Microeconomics • Consider J units (firms, plants) with production sets Y1, . . . , YJ equipped with profit functions πj(p) and supply correspondences yj(p), j = 1, . . . , J. • Definition - Aggregate Supply Correspondence: The sum of the yj(p) is called aggregate supply correspondence: y(p) := J j yj(p) = {y ∈ R L |y = J j yj for some yj ∈ yj(p)}, j = 1, . . . , J} • Definition - Aggregate Production Set: The sum of the individual Yj is called aggregate production set: Y = J j Yj = {y ∈ R L |y = J j yj for some yj ∈ Yj, j = 1, . . . , J} 259 Production 3 Aggregate Supply (2) Microeconomics • Proposition The law of supply also holds for the aggregate supply function. • Proof: Since (p − p )[yj(p) − yj(p )] ≥ 0 for all j = 1, . . . , J it has also to hold for the sum. • Definition: π∗ (p) and y∗ (p) are the profit function and the supply correspondence of the aggregate production set Y . 260 Production 3 Aggregate Supply (3) Microeconomics • Proposition[5.E.1] For all p 0 we have – π∗ (p) = J j πj(p) – y∗ (p) = J j yj(p) (= { J j yj|yj ∈ yj(p)}) • Suppose that prices are fixed, this proposition implies that the aggregate profit obtained by production of each unit separately is the same as if we maximize the joint profit. 261 Production 3 Aggregate Supply (4) Microeconomics Proof: • π∗ (p) = J j πj(p): Since π∗ is the maximum value function obtained from the aggregate maximization problem, we have π∗ (p) ≥ p · ( j yj) = j p · yj which implies π∗ (p) ≥ j πj(p). • To show equality, note that there are yj in Yj such that y = j yj. Then p · y = j p · yj ≤ j πj(p) for all y ∈ Y . 262 Production 3 Aggregate Supply (5) Microeconomics Proof: • y∗ (p) = J j yj(p): Here we have to show that j yj(p) ⊂ y∗ (p) and y∗ (p) ⊂ j yj(p). Consider yj ∈ yj(p), then p · ( j yj) = j πj(p) = j pyj = j πj(p) = π∗ (p) (the last step by the first part of [5.E.1]). • From this argument is follows that yj(p) ⊂ y∗ (p). • To get the second direction we start with y ∈ y∗ (p). Then y = j yj with yj ∈ Yj. Since p · y = p · ( j ·yj) = π∗ (p) and π∗ (p) = J j πj(p), it must be that p · yj = πj(p) (because yj ∈ Yj implies p · yj ≤ πj(p)). Hence, we get y∗ (p) ⊂ j yj(p). 263 Production 3 Aggregate Supply (6) Microeconomics • The same aggregation procedure can also be applied to derive aggregate cost. 264 Production 3 Efficiency (1) Microeconomics • We want to check whether or what production plans are wasteful. • Definition:[D 5.F.1] A production vector is efficient, if there is no y ∈ Y such that y ≥ y and y = y. • There is no way to increase output with given inputs or to decrease input with given output (sometimes called technical efficiency). • Discuss MWG, Figure 5.F.1, p. 150. 265 Production 3 Efficiency (2) Microeconomics • Proposition[P 5.F.1] If y ∈ Y is profit maximizing for some p 0, then y is efficient. • Version of the first fundamental theorem of welfare economics. See MWG, Chapter 16 C, p. 549. • It also tells us that a profit maximizing firm does not choose interior points in the production set. 266 Production 3 Efficiency (3) Microeconomics Proof: • We show this by means of a contradiction: Suppose that there is a y ∈ Y such that y = y and y ≥ y. Because p 0 we get p · y > p · y, contradicting the assumption that y solves the PMP. • For interior points suppose that y is the interior. By the same argument we see that this is neither efficient nor optimal. 267 Production 3 Efficiency (4) Microeconomics • This result implies that a firm chooses y in the convex part of Y (with a differentiable transfer function F(.) this follows immediately from the first order conditions; otherwise we choose 0 or ∞). • The result also holds for nonconvex production sets - see Figure 5.F.2, page 150. • Generally it is not true that every efficient production vector is profit maximizing for some p ≥ 0, this only works with convex Y . 268 Production 3 Efficiency (6) Microeconomics • Proposition[P 5.F.2] Suppose that Y is convex. Then every efficient production y ∈ Y is profit maximizing for some p ≥ 0 and p = 0. 269 Production 3 Efficiency (7) Microeconomics Proof: • Suppose that y is efficient. Construct the set Py = {y ∈ RL |y y}. This set has to be convex. Since y is efficient the intersection of Y and Py has to be empty. • This implies that we can use the separating hyperplane theorem [T M.G.2], page 948: There is some p = 0 such that p · y ≥ p · y for every y ∈ Py and y ∈ Y . This implies p · y ≥ p · y for every y y. Therefore, we also must have p ≥ 0. If some pl < 0 then we could have p · y < p · y for some y y with yl − yl sufficiently large. This procedure works for each arbitrary y. p = 0. 270 Production 3 Efficiency (8) Microeconomics Proof: • It remains to show that y maximizes the profit: Take an arbitrary y ∈ Y , y was fixed, p has been derived by the separating hyperplane theorem. Then p · y ≥ p · y for every y ∈ Py. y ∈ Py can be chosen arbitrary close to y, such that p · y ≥ p · y still has to hold. I.e. y maximizes the profit given p. • Regarding this proof see also MWG, Figure 5.F.3, p. 151. 271 Production 4 Objectives of the Firm (1) Microeconomics • Until now we have assumed that the firm maximizes its profit. • The price vector p was assumed to be fixed. • We shall see that although preference maximization makes sense when we consider consumers, this need not hold with profit maximization with firms. • Only if p is fixed we can rationalize profit maximization. 272 Production 4 Objectives of the Firm (2) Microeconomics • The objectives of a firm should be a result of the objectives of the owners controlling the firm. That is to say, firm owners are also consumers who look at their preferences. So profit maximization need not be clear even if the firm is owned by one individual. • MWG argue (”optimistically”) that the problem of profit maximization is resolved, when the prices are fixed. This arises with firms with no market power. 273 Production 4 Objectives of the Firm (3) Microeconomics • Consider a production possibility set Y owned by consumers i = 1, . . . , I. The consumers own the shares θi, with I i=1 θi = 1. y ∈ Y is a production decision. wi is non-profit wealth. • Consumer i maximizes utility maxxi≥0 u(xi), s.t. p · xi ≤ wi + θip · y. • With fixed prices the budget set described by p · xi ≤ wi + θip · y increases if p · y increases. • With higher p · y each consumer i is better off. Here maximizing profits p · y makes sense. 274 Production 4 Objectives of the Firm (4) Microeconomics • Problems arise (e.g.) if – Prices depend on the action taken by the firm. – Profits are uncertain (risk attitude plays a role). – Firms are not controlled by its owners. 275 Production 4 Objectives of the Firm (5) Microeconomics • Suppose that the output of a firm is uncertain. It is important to know whether output is sold before or after uncertainty is resolved. • If the goods are sold on a spot market (i.e. after uncertainty is resolved), then also the owner’s attitude towards risk will play a role in the output decision. Maybe less risky production plans are preferred (although the expected profit is lower). • If there is a futures market the firm can sell the good before uncertainty is resolved and the consumers bear the risk. Profit maximization can still be optimal. 276 Production 4 Objectives of the Firm (6) Microeconomics • Consider a two good economy with goods x1 and x2; L = 2, non-profit wealth wi = 0. Suppose that the firm can influence the price of good 1, p1 = p1(x1). We normalize the price of good 2, such that p2 = 1. z units of x2 are used to produce x1 with production function x1 = f(z). The cost is given by p2z = z. • We consider the maximization problem maxxi≥0 u(xi1, xi2), s.t. p · xi ≤ wi + θip · y. Given the above notation p = (p1(x1), 1), y = (f(z), −z). wi = 0 by assumption. The profit is p · y = p1(x1)x1 − p2z = p1(f(z))f(z) − z. 277 Production 4 Objectives of the Firm (7) Microeconomics • Assume that the preferences of the owners are such that they are only interested in good 2. • The aggregate amount of x2 the consumers can buy is 1 p2 (p1(f(z))f(z) − p2z) = p1(f(z))f(z) − z (since p2 = 1). • Hence, maxxi≥0 u(xi2), s.t. p · xi ≤ wi + θip · y results in max p1(f(z))f(z) − z. 278 Production 4 Objectives of the Firm (8) Microeconomics • Assume that the preferences of the owners are such that they only look at good 1. • The aggregate amount of x1 the consumers can buy is 1 p1(f(z)) (p1(f(z))f(z) − z) = f(z) − z/p1(f(z)). • Then maxxi≥0 u(xi1), s.t. p · xi ≤ wi + θip · y results in max f(z) − z/p1(f(z)). • We have two different optimization problems - solutions are different. 279 Production 4 Objectives of the Firm (9) Microeconomics • Example: Let p1(f(z)) = √ z, then the first order conditions are different, i.e. 1 2 √ z f(z) + √ zf (z) − 1 = 0 and f (z) − 1 2 √ z = 0. • We have considered two extreme cases: all owners prefer (i) good 2, (ii) good 1. There is no unique output decision based on max p · y. • If the preferences become heterogeneous things do not become better. 280 Expected Utility Uncertainty (1) Microeconomics • Preferences and Lotteries. • Von Neumann-Morgenstern Expected Utility Theorem. • Attitudes towards risk. • State Dependent Utility, Subjective Utility MWG, Chapter 6. 281 Expected Utility Lotteries (1) Microeconomics • A risky alternative results in one of a number of different states of the world, ωi. • The states are associated with consequences or outcomes, zn. Each zn involves no uncertainty. • Outcomes can be money prices, wealth levels, consumption bundles, etc. • Assume that the set of outcomes is finite. Then Z = {z1, . . . , zN}. • E.g. flip a coin: States {H, T} and outcomes Z = {−1, 1}, with head H or tail T. 282 Expected Utility Lotteries (2) Microeconomics • Definition - Simple Gamble/Simple Lottery: [D 6.B.1] With the consequences {z1, . . . , zN} ⊆ Z and N finite. A simple gamble assigns a probability pn to each outcome zn. pn ≥ 0 and N n=1 pn = 1. • Notation: L = (p1, . . . , pN). pi ≥ 0 is the probability of consequence zi, for i = 1, . . . , N. • Let us fix the set of outcomes Z: Different lotteries correspond to a different set of probabilities. • Definition - Set of Simple Gambles: The set of simple gambles on Z is given by LS = {(p1, . . . , pN)|pn ≥ 0 , N n=1 pn = 1} = {L|pn ≥ 0 , N N=1 pn = 1} 283 Expected Utility Lotteries (3) Microeconomics • Definition - Degenerated Lottery: ˜Ln = (0, . . . , 1, . . . , 0) = en. • ’Z ⊆ LS’, since one can identify zn with ˜Ln . 284 Expected Utility Lotteries (4) Microeconomics • With N consequences, every simple lottery can be represented by a point in a N − 1 dimensional simplex ∆(N−1) = {p ∈ RN + | pn = 1} . • At each corner n we have the degenerated case that pn = 1. • With interior points pn > 0 for all i. • See Ritzberger, p. 36,37, Figures 2.1 and 2.2 or MWG, Figure 6.B.1, page 169. • Equivalent to Machina’s triangle; with N = 3; {(p1, p3) ∈ [0, 1]2 |0 ≤ 1 − p1 − p3 ≤ 1}. 285 Expected Utility Lotteries (5) Microeconomics • The consequences of a lottery need not be a z ∈ Z but can also be a further lottery. • Definition - Compound Lottery:[D 6.B.2] Given K simple lotteries Lk and probabilities αk ≥ 0 and αk = 1, the compound lottery LC = (L1, . . . , Lk, . . . , LK; α1, . . . , αk, . . . , αK). It is the risky alternative that yields the simple lottery Lk with probability αk. • The support of the compound lottery (the set of consequences with positive probability) is the union of the supports generating this lotteries. 286 Expected Utility Lotteries (6) Microeconomics • Definition - Reduced Lottery: For any compound lottery LC we can construct a reduced lottery/simple gamble L ∈ LS. With the probabilities pk for each Lk we get p = αkpk , such that probabilities for each zn ∈ Z are pn = K k=1 αkpk n. • Examples: Example 2.5, Ritzberger p. 37 • A reduced lottery can be expressed by a convex combination of elements of compound lotteries (see Ritzberger, Figure 2.3, page 38). I.e. αpl1 + (1 − α)pl2 = plreduced . 287 Expected Utility von Neumann-Morgenstern Utility (1) Microeconomics • Here we assume that any decision problem can be expressed by means of a lottery (simple gamble). • Only the outcomes matter. • Consumers are able to perform calculations like in probability theory, gambles with the same probability distribution on Z are equivalent. 288 Expected Utility von Neumann-Morgenstern Utility (2) Microeconomics • Axiom vNM1 - Completeness: For two gambles L1 and L2 in LS either L1 L2, L2 L1 or both. • Here we assume that a consumer is able to rank lotteries (risky alternatives). I.e. Axiom vNM1 is stronger than Axiom 1 under certainty. • Axiom vNM2 - Transitivity: For three gambles L1, L2 and L3: L1 L2 and L2 L3 implies L1 L3. 289 Expected Utility von Neumann-Morgenstern Utility (3) Microeconomics • Axiom vNM3 - Continuity: [D 6.B.3] The preference relation on the space of simple lotteries is continuous if for any L1, L2, L3 the sets {α ∈ [0, 1]|αL1 + (1 − α)L2 L3} ⊂ [0, 1] and {α ∈ [0, 1]|L3 αL1 + (1 − α)L2} ⊂ [0, 1] are closed. • Later we show: for any gambles L ∈ LS, there exists some probability α such that L ∼ α¯L + (1 − α)L, where ¯L is the most preferred and L the least preferred lottery. • This assumption rules out a lexicographical ordering of preferences (safety first preferences). 290 Expected Utility von Neumann-Morgenstern Utility (4) Microeconomics • Consider the outcomes Z = {1000, 10, death}, where 1000 10 death. L1 gives 10 with certainty. • If vNM3 holds then L1 can be expressed by means of a linear combination of 1000 and death. If there is no α ∈ [0, 1] fulfilling this requirement vNM3 does not hold. 291 Expected Utility von Neumann-Morgenstern Utility (5) Microeconomics • Monotonicity: For all probabilities α, β ∈ [0, 1], α¯L + (1 − α)L β ¯L + (1 − β)L if and only if α ≥ β. • Monotonicity is implied by the axioms vNM1-vNM4. 292 Expected Utility von Neumann-Morgenstern Utility (6) Microeconomics • Axiom vNM4 - Independence, Substitution: For all probabilities L1, L2 and L3 in LS and α ∈ (0, 1): L1 L2 ⇔ αL1 + (1 − α)L3 αL2 + (1 − α)L3 . • This axiom implies that the preference orderings of the mixtures are independent of the third lottery. • This axiom has no parallel in consumer theory under certainty. 293 Expected Utility von Neumann-Morgenstern Utility (7) Microeconomics • Example: consider a bundle x1 consisting of 1 cake and 1 bottle of wine x1 = (1, 1), x2 = (3, 0); x3 = (3, 3). Assume that x1 x2 . Axiom vNM4 requires that αx1 + (1 − α)x3 αx2 + (1 − α)x3 ; here α > 0. 294 Expected Utility von Neumann-Morgenstern Utility (8) Microeconomics • Lemma - vNM1-4 imply monotonicity: Moreover, if L1 L2 then αL1 + (1 − α)L2 βL1 + (1 − β)L2 for arbitrary α, β ∈ [0, 1] where α ≥ β. For every L1 L L2, there is unique γ ∈ [0, 1] such that γL1 + (1 − γ)L2 ∼ L. • See steps 2-3 of the vNM existence proof. 295 Expected Utility von Neumann-Morgenstern Utility (9) Microeconomics • Definition - von Neumann Morgenstern Expected Utility Function: [D 6.B.5] A real valued function U : LS → R has expected utility form if there is an assignment of numbers (u1, . . . , uN) (with un = u(zn)) such that for every lottery L ∈ LS we have U(L) = zn∈Z p(zn)u(zn). A function of this structure is said to satisfy the expected utility property - it is called von Neumann-Morgenstern (expected) utility function. • Note that this function is linear in the probabilities pn. • u(zn) is called Bernoulli utility function. 296 Expected Utility von Neumann-Morgenstern Utility (10) Microeconomics • Proposition - Linearity of the von Neumann Morgenstern Expect Utility Function: [P 6.B.1] A utility function has expected utility form if and only if it is linear. That is to say: U K k=1 αkLk = K k=1 αkU(Lk) 297 Expected Utility von Neumann-Morgenstern Utility (11) Microeconomics Proof: • Suppose that U( K k=1 αkLk) = K k=1 αkU(Lk) holds. We have to show that U has expected utility form, i.e. if U( k αkLk) = k αkU(Lk) then U(L) = pnu(zn). • If U is linear then we can express any lottery L by means of a compound lottery with probabilities αn = pn and degenerated lotteries ˜Ln . I.e. L = pn ˜Ln . By linearity we get U(L) = U( pn ˜Ln ) = pnU(˜Ln ). • Define u(zn) = U(˜Ln ). Then U(L) = U( pn ˜Ln ) = pnU(˜Ln ) = pnu(zn). Therefore U(.) has expected utility form. 298 Expected Utility von Neumann-Morgenstern Utility (12) Microeconomics Proof: • Suppose that U(L) = N n=1 pnu(zn) holds. We have to show that utility is linear, i.e. if U(L) = pnu(zn) then U( k αkLk) = k αkU(Lk) • Consider a compound lottery (L1, . . . , LK, α1, . . . , αK). Its reduced lottery is L = k αkLk. • Then U( k αkLk) = n k αkpk n u(zn) = k αk n pk nu(zn) = k αkU(Lk). 299 Expected Utility von Neumann-Morgenstern Utility (13) Microeconomics • Proposition - Existence of a von Neumann Morgenstern Expect Utility Function: [P 6.B.3] If the Axioms vNM 1-4 are satisfied for a preference ordering on LS. Then admits an expected utility representation. I.e. there exists a real valued function u(.) on Z which assigns a real number to each outcome zn, n = 1, . . . , N, such that for any pair of lotteries L1 = (p1, . . . , pN) and L2 = (p1, . . . , pN) we get L1 L2 if and only if U(L1) := N n=1 pnu(zn) ≥ U(L2) := N n=1 pnu(zn) . 300 Expected Utility von Neumann-Morgenstern Utility (14) Microeconomics Proof: • Suppose that there is a best and a worst lottery. With a finite set of outcomes this can be easily shown by means of the independence axiom. In addition ¯L L. • By the definition of ¯L and L we get: ¯L Lc L, ¯L L1 L and ¯L L2 L. • We have to show that (i) u(zn) exists and (ii) that for any compound lottery Lc = βL1 + (1 − β)L2 we have U(βL1 + (1 − β)L2) = βU(L1) + (1 − β)U(L2) (expected utility structure). 301 Expected Utility von Neumann-Morgenstern Utility (15) Microeconomics Proof: • Step 1: By the independence Axiom vNM4 we get if L1 L2 and α ∈ (0, 1) then L1 αL1 + (1 − α)L2 L2. • This follows directly from the independence axiom. L1 ∼ αL1+(1−α)L1 αL1+(1−α)L2 αL2+(1−α)L2 = L2 302 Expected Utility von Neumann-Morgenstern Utility (16) Microeconomics Proof: • Step 2: Want to show that β > α , if and only if β ¯L + (1 − β)L α¯L + (1 − α)L (monotonicity): • Define γ = (β − α)/(1 − α); the assumptions imply γ ∈ [0, 1]. 303 Expected Utility von Neumann-Morgenstern Utility (17) Microeconomics Proof: • Then β ¯L + (1 − β)L = γ ¯L + (1 − γ)(α¯L + (1 − α)L) γ(α¯L + (1 − α)L) + (1 − γ)(α¯L + (1 − α)L) ∼ α¯L + (1 − α)L 304 Expected Utility von Neumann-Morgenstern Utility (18) Microeconomics Proof: • Step 2: For the converse we have to show that β ¯L + (1 − β)L α¯L + (1 − α)L results in β > α. We show this by means of the contrapositive: If β > α then β ¯L + (1 − β)L α¯L + (1 − α)L. • Thus assume β ≤ α, then α¯L + (1 − α)L β ¯L + (1 − β)L follows in the same way as above. If α = β indifference follows. 305 Expected Utility von Neumann-Morgenstern Utility (19) Microeconomics Proof: • Step 3: There is a unique αL such that L ∼ αL ¯L + (1 − αL)L. • Existence follows from ¯L L and the continuity axiom: • Ad existence: define the sets {α ∈ [0, 1]|α¯L + (1 − α)L L} and {α ∈ [0, 1]|L α¯L + (1 − α)L}. Both sets are closed. Any α belongs to at least one of these two sets. Both sets are nonempty. Their complements are open and disjoint. The set [0, 1] is connected ⇒ there is at least one α belonging to both sets. • Uniqueness follows directly from step 2. 306 Expected Utility Excursion: Connected Sets Microeconomics • Definition: Let Ω = ∅ be an arbitrary set. A class τ ⊂ 2Ω of subsets of Ω is called a topology on Ω if it has the three properties: – ∅, Ω ∈ τ – A ∩ B ∈ τ for any two sets A, B ∈ τ. – A∈F A ∈ τ for any F ⊂ τ. • The pair (Ω, τ) is called a topological space. The sets A ∈ τ are called open sets, and the sets A ⊂ Ω with Ac ∈ τ are called closed sets; Ac stands for complementary set. 307 Expected Utility Excursion: Connected Sets Microeconomics • Consider the family τR of subsets of R: O ∈ τR if and only if for each x ∈ O, there is an ε > 0 such that (x − ε, x + ε) ⊂ O. That is, elements of O are arbitrary unions of open intervals. • Fact from Math: τR forms a topology on R. It is called Euclidean topology. • We consider the closed interval [0, 1] with the following topology: A ⊂ [0, 1] is open if and only if there is an O ∈ τR such that A = O ∩ [0, 1]. This topology is induced by τR. 308 Expected Utility Excursion: Connected Sets Microeconomics • Definition: Let (X, τ) be a topological space. The space is said to be connected, if for any two non-empty closed subsets A, B ⊂ X, A ∪ B = X implies A ∩ B = ∅. • Fact from Math: [0, 1] with by τR the induced topology is connected. 309 Expected Utility von Neumann-Morgenstern Utility (20) Microeconomics Proof: • Step 4: The function U(L) = αL represents the preference relations . • Consider L1, L2 ∈ LS: If L1 L2 then α1 ≥ α2. If α1 ≥ α2 then L1 L2 by steps 2-3. • It remains to show that this utility function has expected utility form. 310 Expected Utility von Neumann-Morgenstern Utility (21) Microeconomics Proof: • Step 5: U(L) is has expected utility form. • We show that the linear structure also holds for the compound lottery Lc = βL1 + (1 − β)L2. • By using the independence we get: βL1 + (1 − β)L2 ∼ β(α1 ¯L + (1 − α1)L) + (1 − β)L2 ∼ β(α1 ¯L + (1 − α1)L) + (1 − β)(α2 ¯L + (1 − α2)L) ∼ (βα1 + (1 − β)α2)¯L + (β(1 − α1) + (1 − β)(1 − α2))L • By the rule developed in step 4, this shows that U(Lc) = U(βL1 + (1 − β)L2) = βU(L1) + (1 − β)U(L2). 311 Expected Utility von Neumann-Morgenstern Utility (22) Microeconomics • Proposition - von Neumann Morgenstern Expect Utility Function are unique up to Positive Affine Transformations: [P 6.B.2] If U(.) represents the preference ordering , then V represents the same preference ordering if and only if V = α + βU, where β > 0. 312 Expected Utility von Neumann-Morgenstern Utility (23) Microeconomics Proof: • Note that if V (L) = α + βU(L), V (L) fulfills the expected utility property (see also MWG p. 174). • We have to show that if U and V represent preferences, then V has to be an affine linear transformation of U. • If U is constant on LS, then V has to be constant. Both functions can only differ by a constant α. 313 Expected Utility von Neumann-Morgenstern Utility (24) Microeconomics Proof: • Alternatively, for any L ∈ LS and ¯L L, we get f1 := U(L) − U(L) U(¯L) − U(L) and f2 := V (L) − V (L) V (¯L) − V (L) . • f1 and f2 are linear transformations of U and V that satisfy the expected utility property. • fi(L) = 0 and fi(¯L) = 1, for i = 1, 2. 314 Expected Utility von Neumann-Morgenstern Utility (25) Microeconomics Proof: • L ∼ L then f1 = f2 = 0; if L ∼ ¯L then f1 = f2 = 1. • By expected utility U(L) = γU(¯L) + (1 − γ)U(L) and V (L) = γV (¯L) + (1 − γ)V (L). • If ¯L L L then there has to exist a unique γ, such that L L ∼ γ ¯L + (1 − γ)L ¯L. Therefore γ = U(L) − U(L) U(¯L) − U(L) = V (L) − V (L) V (¯L) − V (L) 315 Expected Utility von Neumann-Morgenstern Utility (26) Microeconomics Proof: • Then V (L) = α + βU(L) where α = V (L) − U(L) V (¯L) − V (L) U(¯L) − U(L) and β = V (¯L) − V (L) U(¯L) − U(L) . 316 Expected Utility von Neumann-Morgenstern Utility (27) Microeconomics • The concept of expected utility can be extended to a set of distributions F(x) where the expectation of u(x) exists, i.e. A u(x)dF(x) < ∞, z ∈ R and A ⊂ R. • For technical details see e.g. Robert (1994), The Bayesian Choice and DeGroot, Optimal Statistical Decisions. • Note that expected utility is a probability weighted combination of Bernoulli utility functions. I.e. the properties of the random variable z, described by the lottery l(z), are separated from the attitudes towards risk. 317 Expected Utility VNM Indifference Curves (1) Microeconomics • Indifferences curves are straight lines; see Ritzberger, Figure 2.4, page 41. • Consider a VNM utility function and two indifferent lotteries L1 and L2. It has to hold that U(L1) = U(L2). • By the expected utility theorem U(αL1 + (1 − α)L2) = αU(L1) + (1 − α)U(L2). • If U(L1) = U(L2) then U(αL1 + (1 − α)L2) = U(L1) = U(L2) has to hold and the indifferent lotteries is linear combinations of L1 and L2. 318 Expected Utility VNM Indifference Curves (2) Microeconomics • Indifference curves are parallel; see Ritzberger, Figure 2.5, 2.6, page 42. • Consider L1 ∼ L2 and a further lottery L3 L1 (w.l.g.). • From βL1 + (1 − β)L3 and βL2 + (1 − β)L3 we have received two compound lotteries. • By construction these lotteries are on a line parallel to the line connecting L1 and L2. 319 Expected Utility VNM Indifference Curves (3) Microeconomics • The independence axiom vNM4 implies that βL1 + (1 − β)L3 ∼ βL2 + (1 − β)L3 for β ∈ [0, 1]. • Therefore the line connecting the points βL1 + (1 − β)L3 and βL2 + (1 − β)L3 is an indifference curve. • The new indifference curve is a parallel shift of the old curve; by the linear structure of the expected utility function no other indifference curves are possible. 320 Expected Utility Allais Paradoxon (1) Microeconomics Lottery 0 1-10 11-99 pz 1/100 10/100 89/100 La 500,000 500,000 500,000 Lb 0 2,500,000 500,000 Ma 500,000 500,000 0 Mb 0 2,500,000 0 321 Expected Utility Allais Paradoxon (2) Microeconomics • Most people prefer La to Lb and Mb to Ma. • This is a contradiction to the independence axiom G5. • Allais paradoxon in the Machina triangle, Gollier, Figure 1.2, page 8. 322 Expected Utility Allais Paradoxon (3) Microeconomics • Expected utility theory avoids problems of time inconsistency. • Agents violating the independence axiom are subject to Dutch book outcomes (violate no money pump assumption). 323 Expected Utility Allais Paradoxon (4) Microeconomics • Three lotteries: La Lb and La Lc. • But Ld = 0.5Lb + 0.5Lc La. • Gambler is willing to pay some fee to replace La by Ld. 324 Expected Utility Allais Paradoxon (5) Microeconomics • After nature moves: Lb or Lc with Ld. • Now the agents is once again willing to pay a positive amount for receiving La • Gambler starting with La and holding at the end La has paid two fees! • Dynamically inconsistent/Time inconsistent. • Discuss Figure 1.3, Gollier, page 12. 325 Expected Utility Risk Attitude (1) Microeconomics • For the proof of the vNM-utility function we did not place any assumptions on the Bernoulli utility function u(z). • For applications often a Bernoulli utility function has to be specified. • In the following we consider z ∈ R and u (z) > 0; abbreviate lotteries with money amounts l ∈ LS. • There are interesting interdependences between the Bernoulli utility function and an agent’s attitude towards risk. 326 Expected Utility Risk Attitude (2) Microeconomics • Consider a nondegenerated lottery l ∈ LS and a degenerated lottery ˜l. Assume that El(z) = ˜zl holds. I.e. the degenerated lottery ˜l pays the expectation ˜zl of l for sure. • Definition - Risk Aversion: A consumer is risk averse if for any lottery l, ˜zl is at least as good as l. A consumer is strictly risk averse if for any lottery l, ˜z is strictly preferred to l, whenever l is non-degenerate. • Definition - Risk Neutrality: A consumer is risk neutral if ˜zl ∼ l for all l. • Definition - Risk Loving: A consumer is risk loving if for any lottery l, ˜zl is at most as good as l. A consumer is strictly risk loving if for any lottery l, l is strictly preferred to ˜zl, whenever l is non-degenerate. 327 Expected Utility Risk Attitude (3) Microeconomics • By the definition of risk aversion, we see that the utility function u(.) has to satisfy for any non-degenerate distribution F , u(E(z)) = u( zdF (z)) ≥ E(u(z)) = u(z)dF (z). • If u(z) is a concave function and z is distributed according to F (z) (such that the expectations exist), then u(z)dF (z) ≤ u( zdF (z)) Jensen’s inequality. In addition, if u(z)dF (z) ≤ u( zdF (z)) holds for any distribution F , then u(z) is concave. • For sums this implies: pzu(z) ≤ u( pzz) . For strictly concave function, < has to hold whenever F is nondegenerate, for convex functions we get ≥; for strictly convex functions > whenever F is non-degenerate. 328 Expected Utility Risk Attitude (4) Microeconomics • For a lottery l where E(u(z)) < ∞ and E(z) < ∞ we can calculate the amount C where a consumer is indifferent between receiving C for sure and the lottery l. I.e. l ∼ C and E(u(z)) = u(C) hold. • In addition we are able to calculate the maximum amount π an agent is willing to pay for receiving the fixed amount E(z) for sure instead of the lottery l. I.e. l ∼ E(z) − π or E(u(z)) = u(E(z) − π). 329 Expected Utility Risk Attitude (5) Microeconomics • Definition - Certainty Equivalent [D 6.C.2]: The fixed amount C where a consumer is indifferent between C an a gamble l is called certainty equivalent. • Definition - Risk Premium: The maximum amount π a consumer is willing to pay to exchange the gamble l for a sure state with outcome E(z) is called risk premium. • Note that C and π depend on the properties of the random variable (described by l) and the attitude towards risk (described by u). 330 Expected Utility Risk Attitude (6) Microeconomics • Remark: the same analysis can also be performed with risk neutral and risk loving agents. • Remark: MWG defines a probability premium, which is abbreviated by π in the textbook. Given a degenerated lottery and some ε > 0. The probability-premium πR is defined as u(˜lz) = (1 2 + πR )u(z + ε) + (1 2 − πR )u(z − ε). I.e. mean-preserving spreads are considered here. 331 Expected Utility Risk Attitude (7) Microeconomics • Proposition - Risk Aversion and Bernoulli Utility: Consider an expected utility maximizer with Bernoulli utility function u(.). The following statements are equivalent: – The agent is risk averse. – u(.) is a (strictly) concave function. – C ≤ E(z). (< with strict version) – π ≥ 0. (> with strict version) 332 Expected Utility Risk Attitude (8) Microeconomics Proof: (sketch) • By the definition of risk aversion: for a lottery l where E(z) = z˜l, a risk avers agent ˜l l. • I.e. E(u(z)) ≤ u(z˜l) = u(E(z)) for a VNM utility maximizer. • (ii) follows from Jensen’s inequality. • (iii) If u(.) is (strictly) concave then E(u(z)) = u(C) ≤ u(E(z)) can only be matched with C ≤ E(z). • (iv) With a strictly concave u(.), E(u(z)) = u(E(z) − π) ≤ u(E(z)) can only be matched with π ≥ 0. 333 Expected Utility Arrow Pratt Coefficients (1) Microeconomics • Using simply the second derivative u (z) of the Bernoulli utility function, causes problems with affine linear transformations. • Definition - Arrow-Pratt Coefficient of Absolute Risk Aversion: [D 6.C.3] Given a twice differentiable Bernoulli utility function u(.), the coefficient of absolute risk aversion is defined by A(z) = −u (z)/u (z). • Definition - Arrow-Pratt Coefficient of Relative Risk Aversion: [D 6.C.5] Given a twice differentiable Bernoulli utility function u(.), the coefficient of relative risk aversion is defined by R(z) = −zu (z)/u (z). 334 Expected Utility Comparative Analysis (1) Microeconomics • Consider two agents with Bernoulli utility functions u1 and u2. We want to compare their attitudes towards risk. Definition - More Risk Averse: Agent 1 is more risk averse than agent 2: Whenever agent 1 finds a lottery F at least good as a riskless outcome ˜x, then agent 2 finds F at least good as ˜x. I.e. if F 1 ˜L˜x then F 2 ˜L˜x. In terms of a VNM-ultility maximizer: If EF (u1(z)) = u1(z)dF(z) ≥ u1(˜x) then EF (u2(z)) = u2(z)dF(z) ≥ u2(˜x) for any F(.) and ˜x. 335 Expected Utility Comparative Analysis (2) Microeconomics • Define a function φ(x) = u1(u−1 2 (x)). Since u2(.) is an increasing function this expression is well defined. We, in addition, assume that the first and the second derivatives exist. • By construction with x = u2(z) we get: φ(x) = u1(u−1 2 (x)) = u1(u−1 2 (u2(z))) = u1(z). I.e. φ(x) transforms u2 into u1, such that u1(z) = φ(u2(z)). • In the following we assume that ui and φ are differentiable. In the following theorem we shall observe that φ > 0 for u1 and u2 > 0. 336 Expected Utility Comparative Analysis (3) Microeconomics • Proposition - More Risk Averse Agents [P 6.C.2]: Assume that the first and second derivatives of the Bernoulli utility functions u1 and u2 exist (u > 0 and u < 0). Then the following statements are equivalent: – Agent 1 is (strictly) more risk averse than agent 2. – u1 is a (strictly) concave transformation of u2 (that is, there exists a (strictly) concave φ such that u1(.) = φ(u2(.))) – A1(z) ≥ A2(z) (> for strict) for all z. – C1 ≤ C2 and π1 ≥ π2; (<> for strict). 337 Expected Utility Comparative Analysis (4) Microeconomics Proof: • Step 1: (i) follows from (ii): We have to show that if φ is concave, then if EF (u1(z)) = u1(z)dF (z) ≥ u1(˜x) ⇒ EF (u2(z)) = u2(z)dF (z) ≥ u2(˜x) has to follow. • Suppose that for some lottery F the inequality EF (u1(z)) = u1(z)dF (z) ≥ u1(˜x) holds. This implies EF (u1(z)) = u1(z)dF (z) ≥ u1(˜x) = φ(u2(˜x)). • By means of Jensen’s inequality we get for a concave φ(.); (with strict concave we get <) E(u1(z)) = E(φ(u2(z)) ≤ φ(E(u2(z))). • Then φ(E(u2(z))) ≥ E(u1(z)) and E(u1(z)) ≥ u1(˜x) = φ(u2(˜x)) implies φ(E(u2(z))) ≥ φ(u2(˜x)). • Since φ is increasing this implies E(u2(z)) ≥ u2(˜x). 338 Expected Utility Comparative Analysis (5) Microeconomics Proof: • (ii) follows from (i): Suppose that EF (u1(z)) = u1(z)dF(z) ≥ u1(˜x) ⇒ EF (u2(z)) = u2(z)dF(z) ≥ u2(˜x) for any F(.) and ˜x holds and φ is not concave. • Then EF (u1(z)) = u1(CF 1) has to hold as well with ˜x = CF 1. This implies EF (u1(z)) = EF (φ(u2(z))) = φ(u2(CF 1)) for lottery F. • Since φ is not concave, there exits a lottery where φ(EF (u2(z))) < EF (φ(u2(z))) = φ(u2(CF 1)). This yields EF (u2(z)) < u2(CF 1). Contradiction! 339 Expected Utility Comparative Analysis (6) Microeconomics Proof: • Step 2 (iii)∼ (ii): By the definition of φ and our assumptions we get u1(z) = dφ((u2(z))) dz = φ (u2(z))u2(z) . (since u1, u2 > 0 ⇒ φ > 0) and u1(z) = φ (u2(z))u2(z) + φ (u2(z))(u2(z))2 . 340 Expected Utility Comparative Analysis (7) Microeconomics Proof: • Divide both sides by −u1(z) < 0 and using u1(z) = ... yields: − u1(z) u1(z) = A1(z) = A2(z) − φ (u2(z)) φ (u2(z)) u2(z) . • Since A1, A2 > 0 due to risk aversion, φ > 0 and φ ≤ 0 (<) due to its concave shape we get A1(z) ≥ A2(z) (>) for all z. 341 Expected Utility Comparative Analysis (8) Microeconomics Proof: • Step 3 (iv)∼ (ii): Jensen’s inequality yields (with strictly concave φ) u1(C1) = E(u1(z)) = E(φ(u2(z)) < φ(E(u2(z))) = φ(u2(C2)) = u1(C2) • Since u1 > 0 we get C1 < C2. • π1 > π2 works in the same way. • The above considerations also work in both directions, therefore (ii) and (iv) are equivalent. 342 Expected Utility Comparative Analysis (9) Microeconomics Proof: • Step 4 (vi)∼ (ii): Jensen’s inequality yields (with strictly concave φ) u1(E(z)−π1) = E(u1(z)) = E(φ(u2(z)) < φ(E(u2(z))) = φ(u2(E(z)−π2)) = u1(E(z)−π2) • Since u1 > 0 we get π1 > π2. 343 Expected Utility Stochastic Dominance (1) Microeconomics • In an application, do we have to specify the Bernoulli utility function? • Are there some lotteries (distributions) such that F(z) is (strictly) preferred to G(z)? • E.g. if X(ω) > Y (ω) a.s.? • YES ⇒ Concept of stochastic dominance. • MWG, Figure 6.D.1., page 196. 344 Expected Utility Stochastic Dominance (2) Microeconomics • Definition - First Order Stochastic Dominance: [D 6.D.1] A distribution F(z) first order dominates the distribution G(z) if for every nondecreasing function u : R → R we have ∞ −∞ u(z)dF(z) ≥ ∞ −∞ u(z)dG(z). • Definition - Second Order Stochastic Dominance: [D 6.D.2] A distribution F(z) second order dominates the distribution G(z) if EF (z) = EG(z) and for every nondecreasing concave function u : R+ → R the inequality ∞ 0 u(z)dF(z) ≥ ∞ 0 u(z)dG(z) holds. 345 Expected Utility Stochastic Dominance (3) Microeconomics • Proposition - First Order Stochastic Dominance: [P 6.D.1] F(z) first order dominates the distribution G(z) if and only if F(z) ≤ G(z). • Proposition - Second Order Stochastic Dominance: [D 6.D.2] F(z) second order dominates the distribution G(z) if and only if ¯z 0 F(z)dz ≤ ¯z 0 G(z)dz for all ¯z in R+ . • Remark: I.e. if we can show stochastic dominance we do not have to specify any Bernoulli utility function! 346 Expected Utility Stochastic Dominance (4) Microeconomics Proof: • Assume that u is differentiable and u ≥ 0 • Step 1: First order, if part: If F(z) ≤ G(z) integration by parts yields: ∞ −∞ u(z)dF (z) − ∞ −∞ u(z)dG(z) = ∞ −∞ u(z)F (z)dz − ∞ −∞ u(z)G (z)dz = u(z)(F (z) − G(z))| ∞ −∞ − ∞ −∞ u (z)(F (z) − G(z))dz = − ∞ −∞ u (z)(F (z) − G(z))dz ≥ 0 . • The above inequality holds since the terms inside the integral (F(z) − G(z)) ≤ 0. In addition, limt→∞ F(t) = 1 and limt→−∞ F(t) = 0 and likewise for G(.). 347 Expected Utility Stochastic Dominance (5) Microeconomics Proof: • Step 2: First order, only if part: If FOSD then F (z) ≤ G(z) holds. Proof by means of contradiction. • Assume there is a ¯z ∈ R such that F (¯z) > G(¯z). ¯z > −∞ by construction. Set u(z) = 0 for z ≤ ¯z and u(z) = 1 for z > ¯z. Here we get ∞ −∞ u(z)dF (z) − ∞ −∞ u(z)dG(z) = ∞ −∞ u(z)F (z)dz − ∞ −∞ u(z)G (z)dz = ∞ ¯z F (z)dz − ∞ ¯z G (z)dz = (1 − F (¯z)) − (1 − G(¯z)) = −F (¯z) + G(¯z) < 0 348 Expected Utility Stochastic Dominance (6) Microeconomics Proof: • Second Order SD: Assume that u is twice continuously differentiable, such that u (z) ≤ 0, w.l.g. u(0) = 0. • Remark: The equality of means implies: 0 = ∞ 0 zdF (z) − ∞ 0 zdG(z) = ∞ 0 zF (z)dz − ∞ 0 zG (z)dz = z(F (z) − G(z))| ∞ 0 − ∞ 0 (F (z) − G(z))dz = − ∞ 0 (F (z) − G(z))dz . 349 Expected Utility Stochastic Dominance (7) Microeconomics Proof: • Step 3: Second order, if part: Integration by parts yields: ∞ 0 u(z)dF (z) − ∞ 0 u(z)dG(z) = u(z)(F (z) − G(z))| ∞ 0 − ∞ 0 u (z)(F (z) − G(z))dz = − ∞ 0 u (z)(F (z) − G(z))dz = −u (z) z 0 (F (x) − G(x))dx| ∞ 0 − ∞ 0 −u (z) z 0 (F (x) − G(x))dx dz = ∞ 0 u (z) z 0 (F (x) − G(x))dx dz ≥ 0 • Note that u ≤ 0 by assumption. 350 Expected Utility Stochastic Dominance (8) Microeconomics Proof: • Step 4: Second order, only if part: Consider a ¯z such that u(z) = ¯z for all z > ¯z and u(z) = z for all z ≤ ¯z. This yields: ∞ 0 u(z)dF(z) − ∞ 0 u(z)dG(z) = ¯z 0 zdF(z) − ¯z 0 zdG(z) + ¯z ((1 − F(¯z)) − (1 − G(¯z))) = z (F(z) − G(z)) |¯z 0 − ¯z 0 (F(z) − G(z)) dz − ¯z (F(¯z) − G(¯z)) = − ¯z 0 (F(z) − G(z)) dz ≥ 0 . 351 Expected Utility Stochastic Dominance (9) Microeconomics • Definiton - Monotone Likelihood Ratio Property: The distributions F(z) and G(z) fulfill, the monotone likelihood rate property if G(z)/F(z) is non-increasing in z. • For x → ∞ G(z)/F(z) = 1 has to hold. This and the fact that G(z)/F(z) is non-increasing implies G(z)/F(z) ≥ 1 for all z. • Proposition - First Order Stochastic Dominance follows from MLP: MLP results in F(z) ≤ G(z). 352 Expected Utility Arrow-Pratt Approximation (1) Microeconomics • By means of the Arrow-Pratt approximation we can express the risk premium π in terms of the Arrow-Pratt measures of risk. • Assume that z = w + kx, where w is a fixed constant (e.g. wealth), x is a mean zero random variable and k ≥ 0. By this assumption the variance of z is given by V(z) = k2 V(x) = k2 E(x2 ). • Proposition - Arrow-Pratt Risk Premium with respect to Additive risk: If risk is additive, i.e. z = w + kx, then the risk premium π is approximately equal to 0.5A(w)V(z). 353 Expected Utility Arrow-Pratt Approximation (2) Microeconomics Proof: • By the definition of the risk premium we have E(u(z)) = E(u(w + kx)) = u(w − π(k)). • For k = 0 we get π(k) = 0. For risk averse agents dπ(k)/dk ≥ 0. • Use the definition of the risk premium and take the first derivate with respect to k on both sides: E(xu (w + kx)) = −π (k)u (w − π(k)) . 354 Expected Utility Arrow-Pratt Approximation (3) Microeconomics Proof: • For the left hand side we get at k = 0: E(xu (w + 0x)) = u (w)E(x) = 0 since E(x) = 0 by assumption. • Matching LHS with RHS results in π (k) = 0 at k = 0, while u (.) > 0 by assumption. 355 Expected Utility Arrow-Pratt Approximation (4) Microeconomics Proof: • Taking the second derivative with respect to k yields: E(x2 u (w + kx)) = (π (k))2 u (w − π(k)) − π (k)u (w − π(k)) • At k = 0 this results in (note that π (0) = 0): π (0) = − u (w) u (w) E(x2 ) = A(w)E(x2 ) 356 Expected Utility Arrow-Pratt Approximation (5) Microeconomics • A second order Taylor expansion of π(k) around k = 0 results in π(k) ≈ π(0) + π (0)k + π (0) 2 k2 • Thus π(k) ≈ 0.5A(w)E(x2 )k2 • Since E(x) = 0 by assumption, the risk premium is proportional to the variance of x, that is V(z) = k2 E(x2 ). 357 Expected Utility Arrow-Pratt Approximation (6) Microeconomics • For multiplicative risk we can proceed as follows: z = w(1 + kx) where E(x) = 0. • Proceeding the same way results in: π(k) w ≈ − wu (w) u (w) k2 E(x2 ) = 0.5R(w)E(x2 )k2 • Proposition - Arrow-Pratt Relative Risk Premium with respect to Multiplicative risk: If risk is multiplicative, i.e. z = w(1 + kx), then the relative risk premium π/w is approximately equal to 0.5R(w)k2 V(x). • Interpretation: Risk premium per monetary unit of wealth. 358 Expected Utility Decreasing Absolute Risk Aversion (1) Microeconomics • It is widely believed that the more wealthy an agent, the smaller his/her willingness to pay to escape a given additive risk. • Definition - Decreasing Absolute Risk Aversion[D 6.C.4]: The Bernoulli utility function for money exhibits decreasing absolute risk aversion if the Arrow-Pratt coefficient of absolute risk aversion −u (.) u (.) is a decreasing function of wealth w. 359 Expected Utility Decreasing Absolute Risk Aversion (2) Microeconomics • Proposition - Decreasing Absolute Risk Aversion: [P 6.C.3] The following statements are equivalent – The risk premium is a decreasing function in wealth w. – Absolute risk aversion A(w) is decreasing in wealth. – −u (z) is a concave transformation of u. I.e. u is sufficiently convex. 360 Expected Utility Decreasing Absolute Risk Aversion (3) Microeconomics Proof: (sketch) • Step 1, (i) ∼ (iii): Consider additive risk and the definition of the risk premium. Treat π as a function of wealth: E(u(w + kx)) = u(w − π(w)) . • Taking the first derivative yields: E(1u (w + kx)) = (1 − π (w))u (w − π(w)) . 361 Expected Utility Decreasing Absolute Risk Aversion (4) Microeconomics Proof: (sketch) • This yields: π (w) = − E(u (w + kx)) − u (w − π(w)) u (w − π(w)) . • π (w) decreases if E(u (w + kx)) − u (w − π(w)) ≥ 0. • This is equivalent to E(−u (w + kx)) ≤ −u (w − π(w)). • Note that we have proven that if E(u2(z)) = u2(z − π2) then E(u1(z)) ≤ u1(z − π2) if agent 1 were more risk averse. 362 Expected Utility Decreasing Absolute Risk Aversion (5) Microeconomics Proof: (sketch) • Here we have the same mathematical structure (see slides on Comparative Analysis): set z = w + kx, u1 = −u and u2 = u. • ⇒ −u is more concave than u such that −u is a concave transformation of u. 363 Expected Utility Decreasing Absolute Risk Aversion (6) Microeconomics Proof: (sketch) • Step 2, (iii) ∼ (ii): Next define P(w) := −u u which is often called degree of absolute prudence. • From our former theorems we get: P(w) ≥ A(w) has to be fulfilled (see A1 and A2). • Take the first derivative of the Arrow-Pratt measure yields: A (w) = − 1 (u (w))2 (u (w)u (w) − (u (w)) 2 ) = − u (w) (u (w)) (u (w)/u (w) − u (w)/u (w)) = u (w) (u (w)) (P (w) − A(w)) 364 Expected Utility Decreasing Absolute Risk Aversion (7) Microeconomics Proof: (sketch) • A decreases in wealth if A (w) ≤ 0. • We get A (w) ≤ 0 if P(w) ≥ A(w). 365 Expected Utility HARA Utility (1) Microeconomics • Definition - Harmonic Absolute Risk Aversion: A Bernoulli utility function exhibits HARA if its absolute risk tolerance (= inverse of absolute risk aversion) T(z) := 1/A(z) is linear in wealth z. • I.e. T(z) = −u (z)/u (z) is linear in z • These functions have the form u(z) = ζ (η + z/γ) 1−γ . • Given the domain of z, η + z/γ > 0 has to hold. 366 Expected Utility HARA Utility (2) Microeconomics • Taking derivatives results in: u (z) = ζ 1 − γ γ (η + z/γ) −γ u (z) = −ζ 1 − γ γ (η + z/γ) −γ−1 u (z) = ζ (1 − γ)(γ + 1) γ2 (η + z/γ) −γ−2 367 Expected Utility HARA Utility (3) Microeconomics • Risk aversion: A(z) = (η + z/γ) −1 • Risk Tolerance is linear in z: T(z) = η + z/γ • Absolute Prudence: P(z) = γ+1 γ (η + z/γ) −1 • Relative Risk Aversion: R(z) = z (η + z/γ) −1 368 Expected Utility HARA Utility (4) Microeconomics • With η = 0, R(z) = γ: Constant Relative Risk Aversion Utility Function: u(z) = log(z) for γ = 1 and u(z) = z1−γ 1−γ for γ = 1. • This function exhibits DARA; A (z) = −γ2 /z2 < 0. 369 Expected Utility HARA Utility (5) Microeconomics • With γ → ∞: Constant Absolute Risk Aversion Utility Function: A(z) = 1/η. • Since u (z) = Au (z) we get u(z) = − exp(−Az)/A. • This function exhibits increasing relative risk aversion. 370 Expected Utility HARA Utility (6) Microeconomics • With γ = −1: Quadratic Utility Function: • This functions requires z < η, since it is decreasing over η. • Increasing absolute risk aversion. 371 Expected Utility State Dependent Utility (1) Microeconomics • With von Neumann Morgenstern utility theory only the consequences and their corresponding probabilities matter. • I.e. the underlying cause of the consequence does not play any role. • If the cause is one’s state of health this assumption is unlikely to be fulfilled. • Example car insurance: Consider fair full cover insurance. Under VNM utility U(l) = pu(w − P) + (1 − p)u(w − P), etc. If however it plays a role whether we have a wealth of w − P in the case of no accident or getting compensated by the insurance company such the wealth is w − P, the agent’s preferences depend on the states accident and no accident. 372 Expected Utility State Dependent Utility (2) Microeconomics • With VNM utility theory we have considered the set of simple lotteries LS over the set of consequences Z. Each lottery li corresponds to a probability distribution on Z. • Assume that Ω has finite states. Define a random variable f mapping from Ω into LS. Then f(ω) = lω for all ω of Ω. I.e. f assigns a simple lottery to each state ω. • If the probabilities of the states are given by π(ω), we arrive at the compound lotteries lSDU = π(ω)lω. • I.e. we have calculated probabilities of compound lotteries. 373 Expected Utility State Dependent Utility (3) Microeconomics • The set of lSDU will be called LSDU. Such lotteries are also called horse lotteries. • Note that also convex combinations of lSDU are ∈ LSDU. • Definition - Extended Independence Axiom: The preference relation satisfies extended independence if for all l1 SDU, l2 SDU, lSDU ∈ LSDU and α ∈ (0, 1) we have l1 SDU lSDU if and only if αl1 SDU + (1 − α)l2 SDU αlSDU + (1 − α)l2 SDU. 374 Expected Utility State Dependent Utility (4) Microeconomics • Proposition - Extended Expected Utility/State Dependent Utility: Suppose that Ω is finite and the preference relation satisfies continuity and in independence on LSDU. Then there exists a real valued function u : Z × Ω → R such that l1 SDU l2 SDU if and only if ω∈Ω π(ω) z∈supp(l1 SDU (ω)) pl1(z|ω)u(z, ω) ≥ ω∈Ω π(ω) z∈supp(l2 SDU (ω)) pl2(z|ω)u(z, ω) . 375 Expected Utility State Dependent Utility (4) Microeconomics • u is unique up to positive linear transformations. • Proof: see Ritzberger, page 73. • If only consequences matter such that u(z, ω) = u(z) then state dependent utility is equal to VNM utility. 376 Competitive Markets Outline Microeconomics • Partial equilibrium analysis • Perfect Competition • Entry and perfect competition MWG: Chapter 10 A, B, C, F 377 Competitive Markets Pareto Optimality (1) Microeconomics • Consider I consumers, indexed i = 1, . . . , I. Xi ⊂ RL + are the consumption sets. Each consumer chooses a consumption bundle xi, the utility is given by ui(xi). • J firms, indexed j = 1, . . . , J. The production possibility sets are Yj ∈ RL . The production vectors are yj. • L goods, indexed l = 1, . . . , L. • Total endowments of good l is wl ≥ 0. The total net amount of goods available is wl + j ylj, l = 1, . . . , L. • We assume that the initial endowments and technological possibilities (i.e. the firms) are owned by consumers. Shares θij, where I i=1 θij = 1 for all j = 1, . . . , J. 378 Competitive Markets Pareto Optimality (2) Microeconomics • Remark: Often the endowments are abbreviated by el. Here we stick to MWG and use wl. • Definition - Economic Allocation [D 10.B.1]: An economic allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is a specification of a consumption vector xi ∈ Xi for each consumer i = 1, . . . , I and a production vector yj ∈ Yj for each firm j = 1, . . . , J. The allocation is feasible if I i=1 xli ≤ wl + J j=1 ylj for l = 1, . . . , L. 379 Competitive Markets Pareto Optimality (3) Microeconomics • Definition - Pareto Optimality [D 10.B.2]: A feasible allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is Pareto optimal (efficient) if there is no other feasible allocation (x1, . . . , xI, y1, . . . , yJ) such that ui(xi) ≥ ui(xi) for all i = 1, . . . , I and ui(xi) > ui(xi) for some i. • Definition - Utility Possibility Set: ”The set of attainable utility levels”. U = {(u1, . . . , uI) ∈ R I |∃ feasible allocation (x, y): ui ≤ ui(xi) for i = 1, . . . , I} • Pareto efficient allocations are on the north-east boundary of this set. See Figure 10.B.1. 380 Competitive Markets Competitive Equilibria (1) Microeconomics • Definition - Competitive Economy – Suppose that consumer i initially owns wli, where wl = I i=1 wli for l = 1, . . . , L, wi = (wi1, . . . , wiL). – Consumers i owns the shares θi = (θi1, . . . , θij, . . . , θiJ), where I i=1 θij = 1 for j = 1, . . . , J. – Markets exist for all L goods and all firms are price takers; the prices are p = (p1, . . . , pL). 381 Competitive Markets Competitive Equilibria (2) Microeconomics • Definition - Walrasian Equilibrium [D 10.B.3] The allocation (x, y) and the price vector p ∈ RL constitute a competitive (Walrasian) equilibirium if the following conditions are met: 382 Competitive Markets Competitive Equilibria (3) Microeconomics • Profit maximization: each firm j solves maxyj∈Yj p · yj where yj ∈ Yj. • Utility maximization: each consumer i solves max xi∈Xi u(xi) s.t. p · xi ≤ p · wi + J j=1 θij(p · yj). • Market clearing: For each good l = 1, . . . , L: I i=1 xli = wl + J j=1 yli. 383 Competitive Markets Competitive Equilibria (3) Microeconomics • Proposition [P 10.B.1]: If the allocation (x, y) and the price vector p 0 satisfy the market clearing condition for all goods l = k and if every consumer’s budget constraint is satisfied with equality (p · xi = p · wi + J j=1 θij(p · yj)), then the market for good k also clears. 384 Competitive Markets Competitive Equilibria (4) Microeconomics Proof: • p · xi = p · wi + J j=1 θij(p · yj) holds for all i = 1, . . . , I. • This yields i l plxli = i(p · wi + J j=1 θijyj). • Rearranging terms results in (i): l pl i xli = l plxl = l pl i (wli+ J j=1 θijylj) = l pl(wl+yl) for all l. 385 Competitive Markets Competitive Equilibria (5) Microeconomics Proof: • Markets clear for markets l = k by assumption such that (ii): plxl = pl(wl + yl) for l = k. • Combining (i) and (ii) results in pkxk = pk(wk + yk). 386 Competitive Markets Partial Equilibrium (1) Microeconomics • Marshallian partial equilibrium analysis investigates the market for one good (or several goods). • Argument I: when the expenditure for the good is small, only a small fraction of wealth will be spent on this good, such that the wealth effect is small. • Argument II: Due to small size of the market (and similarly dispersed substitution effects), a change in the price of the good considered has a neglectable impact on the other prices. • Consider the other goods as a single composite commodity, which we call numeraire (see Hicksian composite commodity). 387 Competitive Markets Partial Equilibrium (2) Microeconomics • Two good quasilinear economy: good l and the numeraire, xi is the consumption of consumer i of the second good, mi is the consumption of the numeraire. • Each consumer has quasilinear utility: ui(mi, xi) = mi + φi(xi), i = 1, . . . , I. To avoid any boundary problems (mi, xi) ∈ R × R+. φi is bounded and twice differentiable, with φi > 0, φi < 0 for xi ≥ 0 and φi(0) = 0. • Good l is the good of the market under study, mi stands for the rest. We already know that there are no wealth effects for non-numeraire goods, i.e. there are no wealth effects for xi. p is the price of the good considered, the price of the numeraire good should be one. 388 Competitive Markets Partial Equilibrium (3) Microeconomics • Each firm j = 1, . . . , J uses m as an input to produce good l; zj is the amount imputed into the production process. qj ≥ 0 is the amount produced by firm j. zj = cj(qj) is the cost to produce qj units of good l. • The production possibility set is given by Yj = {(−zj, qj)|qj, zj ≥ cj(qj)}, c(.) is twice differentiable with c > 0 and c ≥ 0 for all qj ≥ 0. • The initial endowments of good l are zero, while the consumers own wmi of the numeraire, wm = I i=1 wmi. 389 Competitive Markets Partial Equilibrium (4) Microeconomics • Goal: try to find competitive equilibrium/equilibria. • Firms maximize profits: maxqj≥0 pqj − cj(qj) • First order condition: p ≤ cj(qj) with equality if qj > 0. • Second order condition is assumed to be satisfied. 390 Competitive Markets Partial Equilibrium (5) Microeconomics • Consumers maximize utility: maxmi,xi mi + φ(xi) s.t. mi + pxi ≤ wmi + J j=1 θij(pqj − cj(qj)). • Given our assumptions on ui(.) the budget constraint has to hold with equality. 391 Competitive Markets Partial Equilibrium (6) Microeconomics • Either by a plug in of mi from the budget constraint in the utility function or by applying Kuhn/Tucker we get φ (xi) ≤ p with equality for xi > 0. • Then mi = wmi + J j=1 θij(pqj − cj(qj)) − pxi, input zj = cj(qj). 392 Competitive Markets Partial Equilibrium (7) Microeconomics • For the two good economy it is sufficient to check whether the market for good l clears. • Therefore the allocation (x, q) and the price p are a competitive equilibrium if p ≤ c (qj) with equality for qj > 0 φ (xi) ≤ p with equality for xi > 0 I i=1 xi = J j=1 qj . • J + I + 1 equations for the same number of unknowns x, y and p. 393 Competitive Markets Partial Equilibrium (8) Microeconomics • As long as maxi φi(0) > minj cj(0) aggregate consumption and production have to be strictly positive. • In this setting the equilibrium outcome is independent of the distribution of the endowments wmi and the shares θij (ownership structure). This is a result of the quasilinear economy. • Aggregate demand x(p) = xi(p). By the assumptions of φ, the Walrasian demand correspondence is single valued, continuous and downward sloping (at any p < max φi(0)). Since preferences are quasilinear x(p) does not depend on wealth. 394 Competitive Markets Partial Equilibrium (9) Microeconomics • Aggregate supply q(p) = J j=1 qj(p) is continuous and non-decreasing for all p > 0. It is strictly increasing for all p > min cj(0). Then we get the equilibrium price p by intersecting market demand and supply. See Figures page 320. • An equilibrium price need not exist. • If cj(.) is linear, then the firms’ production levels are not uniquely determined. • In the above model we have assumed that the preferences and the technologies are convex. If we deviate from this assumption a competitive equilibrium need not exist (see e.g. Figure 10.C.8, page 324). 395 Competitive Markets Free Entry (1) Microeconomics • Until now J has been fixed. In the following we shall consider ”long run behavior”. • Assume that an infinite number of firms has potential access to a production technology to produce good l. The cost function is c(q) where c(0) = 0 is assumed. • x(p) denotes the aggregate demand function, P(.) is its inverse. • Assumption: the identical active firms produce the same level of output q. With J active firms (qj > 0) the total output is Q = qJ. 396 Competitive Markets Free Entry (2) Microeconomics • Remark: J is a non-negative integer. We assume that firms are sufficiently small such that the fact that J is an integer can be neglected. • J is endogenously determined by market entry and exit. Exit if the profit is smaller zero, entry with positive profit. • In equilibrium the profits have to be zero. 397 Competitive Markets Free Entry (3) Microeconomics • Definition - Long-run Competitive Equilibrium [D 10.F.1]: Given the aggregate demand function x(p) and a cost function c(q) for each potentially active firm having c(0) = 0, a triple (p, q, J) is a long-run competitive equilibrium if – Profit maximization: q solves maxq≥0 pq − c(q). – Market clearing: x(p) = Jq. – Free entry condition: pq − c(q) = 0. 398 Competitive Markets Free Entry (4) Microeconomics • First order condition: p ≤ c (q) with equality if q > 0. • Supply correspondence q(p): For any firm q(p) solves the profit maximization problem - see Chapter 5. • Long-run Aggregate Supply Correspondence Q(p) = Jq(p): (i) Q(p) → ∞ if π(p) > 0, (ii) Q(p) = {Q ≥ 0|Q = Jq for some integer J ≥ 0 and q ∈ q(p)} if π(p) = 0. J is the number of active firms. 399 Competitive Markets Free Entry (5) Microeconomics • Market clearing demands for Q = x(p), Q ∈ Q(p). • Free entry condition and the aggregate supply correspondence result in pq − c(q) = 0. • ⇒ p is a competitive equilibrium if x(p) ∈ Q(p). 400 Competitive Markets Free Entry (6) Microeconomics • Consider constant returns to scale such that c(q) = cq, c > 0. • Assume that x(c) > 0. • The first order condition for the firms result in p ≤ c. Since x(c) > 0, we must have (p − c)q = 0 in an equilibrium. Then p = c. • J and q are not determined, but Jq = x(c). • Q(p) = ∞ if p > c, Q(p) ∈ [0, ∞) if p = c and 0 if p < c. 401 Competitive Markets Free Entry (7) Microeconomics • Suppose that c(q) is increasing and strictly convex. • Assume that x(c (0)) > 0. • If p > c (0), then π(p) > 0 such that Q(p) becomes infinite. If p ≤ c (0) the long run supply is zero, while x(p) > 0. In words, in the first case p = MC(q) > AC(q), such that J → ∞. In the second case the first order condition yields q = 0. • Here no equilibrium exists. See Figure 10.F.2, page 337. 402 Competitive Markets Free Entry (8) Microeconomics • Suppose that there exists a unique q > 0 where average cost is minimized (the output level is also called efficient scale). • Assume that c = c(q)/q is minimized at ¯q. Here we get c (¯q) = c(¯q)/¯q = ¯c. Assume that x(¯c) > 0. • If p > c (¯q) = c(¯q)/¯q this cannot be an equilibrium since π(p) > 0 and Q(p) becomes infinite. • For p < c (¯q) = c(¯q)/¯q, π(p) < 0 and Q(p) = 0 which is not compatible with x(p) ≥ x(¯c) > 0. 403 Competitive Markets Free Entry (9) Microeconomics • If p = c (¯q) = c(¯q)/¯q, then π(p) = 0 and Q(p) = J ¯q = x(¯c). • Thus Q(p) = ∞ if p > ¯c, Q(p) = 0 if p < ¯c and {Q ≥ 0|Q = J ¯q for some integer J ≥ 0}. • With x(p) ∈ Q(p) we get an equilibrium. • See Figure 10.F.4 404 Competitive Markets Some further results (1) Microeconomics • With quasilinear preferences the boundary of the utility possibility set is a hyperplane. • Unit to unit transfer of utility is possible. With these preferences {(u1, . . . , uI)| I i=1 ui ≤ I i=1 φi(xi) + wm − J j=1 cj(qj)} • The boundary of this set is a hyperplane with normal vector (1, . . . , 1). 405 Competitive Markets Some further results (2) Microeconomics • Marshallian surplus: S = I i=1 φi(xi) − J j=1 cj(qj) Maximizing S given the constraint i xi − j qj = 0 yields the same outcome as the allocation attained with the competitive economy. The Lagrange multiplier in this constraint optimization problem is equal to p. 406 Competitive Markets Some further results (3) Microeconomics • Proposition - First Fundamental Theorem of Welfare Economics [10.D.1] If the price p and the allocation (x, y) constitute a competitive equilibrium, the this allocation is Pareto optimal. • Proposition - Second Fundamental Theorem of Welfare Economics [10.D.2] For any Pareto optimal levels of utility (u1, . . . , uI) there are transfers of the numeraire commodity (T1, . . . , TI) satisfying I i=1 Ti = 0, such that a competitive equilibrium reached from the endowments (wm1 + T1, . . . , wmI + TI) yields precisely the utilities (u1, . . . , uI). 407 General Equilibrium Outline Microeconomics • Motivation and main questions to be investigated: – Does a competitive economy result in a Pareto efficient allocation? – Can any Pareto efficient allocation be obtained by means of a price system in a competitive economy? • Edgeworth Box • Robinson Crusoe economies • Small open economies and trade • General vs. partial equilibrium MWG, Chapter 15 408 General Equilibrium Motivation (1) Microeconomics • Consider the economy as a closed and interrelated system. • With the partial equilibrium approach these interrelations are mainly ignored. • The exogenous variables in general equilibrium theory are reduced to a small number of physical realities (number of agents, technologies available, preferences of the agents, endowments of various agents). 409 General Equilibrium Motivation (2) Microeconomics • First we consider: – A pure exchange economy: no production is possible, commodities are ultimately consumed, the individuals are permitted to trade the commodities among themselves. With two consumers and two goods this can be represented in the Edgeworth box. – One consumer - one firm economy, to get a first impression on the impacts of production. 410 General Equilibrium Motivation (3) Microeconomics • Consider I consumers, indexed i = 1, . . . , I. Xi ⊂ RL are the consumption sets. Each consumer chooses a consumption bundle xi, the utility is given by ui(xi). The preferences are i. • J firms, indexed j = 1, . . . , J. The production possibility sets are Yj ∈ RL . The production vectors are yj. • L goods, indexed = 1, . . . , L. 411 General Equilibrium Motivation (4) Microeconomics • Total endowments of good is ω ≥ 0. The total net amount of good available is ω + j y j, = 1, . . . , L. • We assume that the initial endowments and technological possibilities (i.e. the firms) are owned by consumers. Shares θij, where I i=1 θij = 1 for all j = 1, . . . , J. • The wealth of consumer i is wi(p) = p · ωi. • Remark: often the endowments are abbreviated by e . Here we stick to MWG and use ω . 412 General Equilibrium Motivation (5) Microeconomics • Definition - Economic Allocation [D 10.B.1]: An economic allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is a specification of a consumption vector xi ∈ Xi for each consumer i = 1, . . . , I and a production vector yj ∈ Yj for each firm j = 1, . . . , J. The allocation is feasible if I i=1 x i ≤ ω + J j=1 y j for = 1, . . . , L. 413 General Equilibrium Motivation (6) Microeconomics • Definition - Competitive Economy – Suppose that consumer i initially owns ω i, where ω = I i=1 ω i for = 1, . . . , L, ωi = (ωi1, . . . , ωiL). – Consumer i owns the shares θi = (θi1, . . . , θij, . . . , θiJ), where I i=1 θij = 1 for j = 1, . . . , J. – Markets exist for all L goods and all firms are price takers; the prices are p = (p1, . . . , pL). 414 General Equilibrium Motivation (7) Microeconomics • Definition - Walrasian/Competitive Equilibrium [D 10.B.3] The allocation (x, y) and the price vector p ∈ RL constitute a competitive (Walrasian) equilibirium if the following conditions are met: – Profit maximization: each firm j solves maxyj∈Yj p · yj where yj ∈ Yj. – Utility maximization: each consumer i solves max xi∈Xi u(xi) s.t. p · xi ≤ p · ωi + J j=1 θij(p · yj). – Market clearing: For each good = 1, . . . , L: I i=1 x i = ω + J j=1 y j. 415 General Equilibrium Motivation (8) Microeconomics • Definition - Pareto Optimality [D 10.B.2]: A feasible allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is Pareto optimal (efficient) if there is no other feasible allocation (x1, . . . , xI, y1, . . . , yJ) such that ui(xi) ≥ ui(xi) for all i = 1, . . . , I and ui(xi) > ui(xi) for some i. • Definition - Utility Possibility Set: ”The set of attainable utility levels”. U = {(u1, . . . , uI) ∈ R I |∃ feasible allocation (x, y): ui ≤ ui(xi) for i = 1, . . . , I} • Pareto efficient allocations are on the north-east boundary of this set. See MWG, Figure 10.B.1. 416 Edgeworth Box (1) Microeconomics • We consider a pure exchange economy. • Consumers posses initial endowments of commodities. Economic activity consists of trading and consumption. • Now we restrict to a two good - two consumer exchange economy. Then, L = 2, X1 = X2 = R2 +, Y1 = Y2 = −R2 + (the free disposal technology). i is the index of the consumer, the index of our goods. • xi = (x1i, x2i) ∈ Xi. i are the preferences of consumer i. • The initial endowments are ω i ≥ 0. The endowment vector of consumer i is ωi = (ω1i, ω2i). The total endowments of good are ¯ω = ω 1 + ω 2. We assume that ¯ω > 0 for = 1, 2. 417 Edgeworth Box (2) Microeconomics • From the above Definition [D 10.B.1] it follows that an economic allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is a specification of a consumption vector xi ∈ Xi for each consumer i = 1, . . . , I and a production vector yj ∈ Yj for each firm j = 1, . . . , J. It is feasible if I i=1 x i ≤ ¯ω + J j=1 y j for = 1, . . . , L. • For the Edgeworth Box an allocation is some consumption vector x = (x11, x21, x21, x22) ∈ R4 +. • An allocation is feasible if x 1 + x 2 ≤ ¯ω for = 1, 2. 418 Edgeworth Box (3) Microeconomics • Definition - Nonwasteful allocation: If x 1 + x 2 = ¯ω for = 1, 2, then the allocation is called nonwasteful. • Nonwasteful allocations can be described by means of an Edgeworth box. • See MWG, Figure 15.B.1. • For a given price vector p = (p1, p2) the budget line intersects the initial endowment point wi = (w1i, w2i). The slope is −p1 p2 . Note that only the relative price −p1 p2 matters, with −λp1 λp2 , λ ∈ R++, we get the same Edgeworth box with the same budget sets. • See MWG, Figure 15.B.2. 419 Edgeworth Box (4) Microeconomics • Next we assume that the preferences of both consumers are strongly monotone and strictly convex. • For each price p consumer i obtains the budget set Bi(p). By solving the utility maximization problem maxx1i,x2i u(xi) s.t. p · xi ≤ wi(p) we obtain the optimal quantities x1i(p), x2i(p). By collecting x1i(p), x2i(p) for different p, we obtain the mapping OCi : R2 + → R2 +, p → (x1i(p), x2i(p)). This mapping is called offer curve. • By the assumptions on the preferences the solution of the UMP is unique, hence here we obtain a function. 420 Edgeworth Box (5) Microeconomics • The consumer’s offer curve lies within the upper contour set of ωi. • See MWG, Figures 15.B.3.-15.B.5. 421 Edgeworth Box (6) Microeconomics • Definition [D 15.B.1] A Walrasian/Competivie Equilibrium for an Edgeworth box economy is a price vector p∗ and a feasible allocation x∗ = (x∗ 1, x∗ 2) in the Edgeworth box such that for i = 1, 2, x∗ i i xi for all xi ∈ Bi(p∗ ). • At any equilibrium the offer curves intersect. • Consumer’s demand is homogeneous of degree zero in p, i.e. only the relative price matters. • See MWG, Figures 15.B.7 and 15.B.8. 422 Edgeworth Box (7) Microeconomics • A Walrasian equilibrium need not be unique. • See MWG, Figure 15.B.9. • This could already happen with quasilinear preferences, where the preferences are such that different numeraire goods are used. • MWG, Chapter 10 constructs a model where all agents have quasilinear preferences with respect to the same numeraire good. 423 Edgeworth Box (8) Microeconomics • Recall: Definition - Quasilinear Preferences: A monotone preference relation on X = (−∞, ∞) × RL−1 is quasilinear with respect to commodity one (the numeraire good) if : (i) all indifference sets are parallel displacements of each other along the axis of commodity one. I.e. x ∼ y then x + αe1 ∼ y + αe1 and e1 = (1, 0, . . . ). (ii) Good one is desirable: x + αe1 x for all α > 0. [D 3.B.7] • A Walrasian equilibrium need not exist. • This happens e.g. if (i) one consumer only desires only one good or (ii) preferences are non-convex. • See MWG, Figure 15.B.10. 424 Edgeworth Box (9) Microeconomics • Definition - Pareto Optimality [D 15.B.2]: A feasible allocation x in the Edgeworth box is Pareto optimal (or Pareto efficient) if there is no other allocation x in the Edgeworth box with xi i xi for i = 1, 2 and xi i xi for some i. The set of all Pareto optimal allocations is called Pareto set. The contract curve is the part of the Pareto set where both consumers do at least as well as at their initial endowments. • See MWG, Figures 15.B.11 and 15.B.12. 425 Edgeworth Box (10) Microeconomics • We observe in the Edgeworth box that ”every Walrasian equilibrium allocation x∗ belongs to the Pareto set”. This corresponds to the first theorem of welfare economics. Figure 15.B.7(a) and 15.B.8, MWG, page 520. • Regarding the second theorem: a planner can (under convexity assumptions, see MWG, Chapter 16) achieve any desired Pareto efficient allocation. • Hence we define: 426 Edgeworth Box (11) Microeconomics • Definition - Equilibrium with Transfers [D 15.B.3]: An allocation x in the Edgeworth box is supportable as an equilibrium with transfers, if there is a price system p∗ and wealth transfers T1 and T2 satisfying T1 + T2 = 0, such that for each consumer i we have x∗ i xi for all xi ∈ R2 + such that p∗ · xi ≤ p∗ · ωi + Ti. • In the Edworth box we observe that with continuous, strongly monotone and strictly convex preferences any Pareto optimal allocation is supportable. • See MWG, Figure 15.B.13. • See MWG, Figure 15.B.14 - to observe how the second theorem fails with non-convex preferences. 427 One-Consumer, One-Producer (1) Microeconomics • We introduce production in the most simple way. • There are two price taking agents, a single consumer and a single firm. • There are two goods, labor (or leisure) of the consumer and the consumption good produced by the firm. • The preferences defined over leisure x1 and the consumption good x2 are continuous, strongly monotone and strictly convex. The initial endowment consists of ¯L units of leisure and no endowment of the consumption good. 428 One-Consumer, One-Producer (2) Microeconomics • The firm uses labor to produce the consumption good under the increasing and strictly concave production function q = f(z), where z is labor input and q the amount of x2 produced. • The firm maximizes its profit: max z≥0 pf(z) − wz given the prices (p, w). This optimization problem results in the optimal labor demand z(p, w) and output q(p, w). The profit is π(p, w). 429 One-Consumer, One-Producer (3) Microeconomics • The consumer maximizes the utility function u(x1, x2): max x1,x2≥0 u(x1, x2) s.t. px2 ≤ w(¯L − x1) + π(p, w). This results in the Walrasian demand x1(p, w) and x2(p, w). Labor supply corresponds to ¯L − x1(p, w). • See MWG, Figure 15.C.1 on these optimization problems. 430 One-Consumer, One-Producer (4) Microeconomics • Walrasian equilibrium is attained at a pair (p∗ , w∗ ) where x2(p∗ , w∗ ) = q(p∗ , w∗ ) and z(p∗ , w∗ ) = ¯L − x1(p∗ , w∗ ). • See MWG, Figure 15.C.1 on these optimization problems. See MWG, Figure 15.C.2 for an equilibrium. 431 One-Consumer, One-Producer (5) Microeconomics • Remark: A particular consumption-leisure combination can arise in a competitive equilibrium if and only if it maximizes the consumer’s utility subject to the technological and endowment constraints. • ⇒ A Walrasian equilibrium allocation is the same as if a social planner would maximize the consumer’s utility given the technological constraints of the economy. A Walrasian equilibrium is Pareto optimal. 432 One-Consumer, One-Producer (6) Microeconomics • Remark on Non-convexity: Suppose the the production set is not convex, then we can construct examples where the price system does not support the allocation x∗ . • See MWG, Figure 15.C.3 (a). 433 General vs. Partial Equilibrium (1) Microeconomics • Bradford’s (1978) example on taxation: • Consider an economy with N large towns. Each town has a single price taking firm producing a consumption good by means of a strictly concave production function f(z). The consumption good is identical. • The overall economy has M units of labor, inelastically supplied. Utility is derived from consuming the output. • Workers are free to move to another town. Hence the equilibrium wage must be the same, i.e. w1, . . . , wN = ¯w. • Without loss of generality the price of the output is normalized, i.e. p = 1. 434 General vs. Partial Equilibrium (2) Microeconomics • By the symmetric construction of the model we get: each firm hires M/N works, the output of each firm is f(M/N). • Due to price taking we get ¯w = f = ∂f(M/N) ∂(M/N) . • The equilibrium profits are: f(M/N) − ∂f(M/N) ∂(M/N) (M/N). 435 General vs. Partial Equilibrium (3) Microeconomics • Suppose that town 1 levies a tax on labor, the tax rate is t > 0. • Given the wage w1 and the tax rate t we arrive at a labor demand z1, which is implicitly given by f (z1) = t + w1. 436 General vs. Partial Equilibrium (4) Microeconomics • Partial equilibrium argument: N is large, in impact on the other wage rates can be neglected. Hence ¯w remains the same. • Since labor moves freely, we get w1 = ¯w. The supply correspondence is 0 at w1 < ¯w and ∞ at w1 > ¯w. It is [0, ∞] at w1 = ¯w. • Then f (z1) = t + ¯w. z1 falls by our assumptions on f(.), labor moves to other towns. • The incomes of the workers and the profits in towns 2, . . . , N remain the same. The profit of firm 1 decreases, the firms completely bear the tax burden. 437 General vs. Partial Equilibrium (5) Microeconomics • General equilibrium argument: Since labor moves freely, w1, . . . , wN = w still has to hold. All M units of labor are employed by the structure of f(.). • w(t) denotes the equilibrium wage rate when the tax rate is t. By symmetry z2(t) = · · · = zN(t) = z(t). z1(t) is the labor demand in town 1. • Then equilibrium demands for: z1(t)+(N −1)z(t) = M , f (z(t)) = w(t) , f (z1(t)) = w(t)+t. 438 General vs. Partial Equilibrium (6) Microeconomics • Next, f (z1(t)) = w(t) + t = f (M − (N − 1)z(t)) = w(t) + t. By taking the first derivative w.r.t. to t and evaluating at t = 0 (where z1(0) = z(0) = M/N) yields f (M/N)[−(N − 1)]z (0) = w (0) + 1. 439 General vs. Partial Equilibrium (7) Microeconomics • The derivative of f (z(t)) = w(t) w.r.t. to t yields f (M/N)z (0) = w(0) such that w (0) = − 1 N . • Hence, the wage rates in all towns decrease due to the tax in town 1. Only if N goes to infinity this effect becomes zero. • In addition, when we consider the profits of the firms, we observe: π ( ¯w)(w (0) + 1) + (N − 1)π ( ¯w)w (0) = π ( ¯w) − N − 1 N + N − 1 N = 0. Hence, aggregate profit remains constant. The complete burden is attributed to the workers. 440 General vs. Partial Equilibrium (8) Microeconomics • For N large the partial equilibrium approximation regarding prices and quantities is correct. However, the distributional effects remain wrong. • The derivative of f (z(t)) = w(t) w.r.t. to t yields f (M/N)z (0) = w(0) such that w (0) = − 1 N . • Hence, the wage rates in all towns decrease due to the tax in town 1. Only if N goes to infinity this effect becomes zero. 441 General vs. Partial Equilibrium (9) Microeconomics • In addition, when we consider the profits of the firms, we observe: π ( ¯w)(w (0) + 1) + (N − 1)π ( ¯w)w (0) = π ( ¯w) − N − 1 N + N − 1 N = 0. • Hence, aggregate profit remains constant. The complete burden is attributed to the workers. • For N large the partial equilibrium approximation regarding prices and quantities is correct. However, the distributional effects remain wrong. 442 General Equilibrium Microeconomics • First Fundamental Theorem of Welfare Economics • Second Fundamental Theorem of Welfare Economics MWG, Chapter 16 443 Notation (1) Microeconomics • Consider I consumers, indexed i = 1, . . . , I. Xi ⊂ RL are the consumption sets. The preferences are i. i is complete and transitive (rationale consumers). • J firms, indexed j = 1, . . . , J. The production possibility sets are Yj ∈ RL . Yj is non-empty and closed. The production vectors are yj. • L goods, indexed = 1, . . . , L. 444 Notation (2) Microeconomics • The initial endowment of good is ¯ω ∈ RL . The total endowments are ¯ω = (¯ω1, . . . , ¯ωL) ∈ RL . • Basis data of the economy: ([Xi, i]I i=1, [Yj]J j=1, ¯ω). • The wealth of consumer i is wi(p) = p · ωi. 445 Notation (3) Microeconomics • Definition - Economic Allocation [D 16.B.1]: An economic allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is a specification of a consumption vector xi ∈ Xi for each consumer i = 1, . . . , I and a production vector yj ∈ Yj for each firm j = 1, . . . , J. The allocation is feasible if I i=1 x i = ¯ω + J j=1 y j for = 1, . . . , L. This is I i=1 xi = ¯omega + J j=1 yj. We denote the set of feasible allocations by A := {(x, y) ∈ X1×· · ·×XI×Y1×· · ·×YJ : I i=1 xi = ¯ω + J j=1 yj} ⊂ R L(I+J) . 446 Notation (4) Microeconomics • Definition - Pareto Optimality [D 16.B.2]: A feasible allocation (x, y) = (x1, . . . , xI, y1, . . . , yJ) is Pareto optimal (efficient) if there is no other feasible allocation (x , y ) ∈ A that Pareto dominates it. This is, if there is no feasible allocation (x , y ) such that xi i xi for all i = 1, . . . , I and and xi i xi for some i. 447 Notation (5) Microeconomics • Suppose that consumer i initially owns ω i, where ¯ω = I i=1 ω i for = 1, . . . , L, ωi = (ωi1, . . . , ωiL). • Consumer i owns the shares θi = (θi1, . . . , θij, . . . , θiJ), where I i=1 θij = 1 for j = 1, . . . , J. • Markets exist for all L goods and all firms are price takers; the prices are p = (p1, . . . , pL). 448 Notation (6) Microeconomics • Definition [D 16.B.3] (Walrasian/Competitive Equilibrium) – Given a private ownership economy by ([Xi, i]I i=1, [Yj]J j=1, ¯ω, θ). An allocation (x∗ , y∗ ) and the price vector p ∈ RL constitute a competitive (Walrasian) equilibrium if the following conditions are met: ∗ Profit maximization: For each firm j, y∗ j solves the profit maximization problem, i.e. p · yj ≤ p · y ∗ j for all yj ∈ Yj. ∗ Preference maximization: For each consumer i, x∗ i is maximal for i in the budget set {xi ∈ Xi : p · xi ≤ p · ωi + J j=1 θijp · y ∗ j }. ∗ Market clearing: For each good = 1, . . . , L: I i=1 x ∗ i = ¯ω + J j=1 y ∗ i or I i=1 x ∗ i = ¯ω + J j=1 y ∗ j . 449 Notation (7) Microeconomics • Definition [D 16.B.4] (Price Equilibrium with Transfers) – Given a private ownership economy by ([Xi, i]I i=1, [Yj]J j=1, ¯ω, θ). An allocation (x∗ , y∗ ) and the price vector p ∈ RL constitute a price equilibrium with transfers if there is an assignment of wealth levels (w1, . . . , wI) with I i=1 wi = p · ¯ω + j p · y∗ j such that ∗ For each firm j, y∗ j solves the profit maximization problem, i.e. p · yj ≤ p · y ∗ j for all yj ∈ Yj. ∗ For each consumer i, x∗ i is maximal for i in the budget set {xi ∈ Xi : p · xi ≤ p · wi}. ∗ Market clearing: I i=1 x∗ i = ¯ω + J j=1 y∗ j . 450 First Fundamental Theorem of Welfare Economics (1) Microeconomics • Proposition [16.C.1] (First Fundamental Theorem of Welfare Economics) – If the preference relations i are locally nonsatiated and if (x∗ , y∗ , p) is a price equilibrium with transfers, then the allocation x∗ , y∗ is Pareto optimal. In particular, any Walrasian equilibrium is Pareto optimal. • Proof: See MWG page 549. 451 First Fundamental Theorem of Welfare Economics (2) Microeconomics • The First Fundamental Theorem of Welfare Economics is on Pareto optimality. • Recall - Local Nonsatiation: For all x ∈ X and for all ε > 0 there exists some y ∈ X such that ||x − y|| ≤ ε and y x. [D 3.B.3] • Note that markets are complete and price taking is assumed. 452 Second Fundamental Theorem of Welfare Economics (1) Microeconomics • First theorem: Given some assumptions and a price equilibrium with transfers ⇒ Pareto. • Consider a competitive economy with transfers. Given some Pareto efficient allocation (x, y). Does there exist a price system p which supports this Pareto efficient allocation? • Problem I: Convexity - see MWG, Figure 15.C.3 (a). • Problem II: Minimum wealth problem - see MWG, Figure 15.B.10 (a). Let xi ∈ R2 and consider the sequence pn = (1/n, 1), n = 1, 2, . . . . Let wi = (1, 0). Then pn · wi → 0. This is the minimum wealth you can have on X = R2 . In the limit p = (0, 1) and the budget set is the horizontal axis intersected with R2 +. So everything of x1 costs nothing. The budget correspondence is not continuous so Berge’s maximum theorem does not apply. • First investigate convexity. To do this we consider the concept of a quasi-equilibrium. 453 Second Fundamental Theorem of Welfare Economics (2) Microeconomics • Definition [16.D.1] (Price Quasi-equilibrium with Transfers) – Given a private ownership economy by ([Xi, i]I i=1, [Yj]J j=1, ¯ω). An allocation (x∗ , y∗ ) and the price vector p = 0 constitute a price quasi-equilibrium with transfers if there is an assignment of wealth levels (w1, . . . , wI) with wi = p · ¯ω + j p · y∗ j such that ∗ For each firm j, y∗ j solves the profit maximization problem, i.e. p · yj ≤ p · y∗ j for all yj ∈ Yj. ∗ For each consumer i: If xi i x∗ i , then p · xi ≥ wi. ∗ Market clearing: I i=1 x∗ i = ¯ω + J j=1 y∗ j . 454 Second Fundamental Theorem of Welfare Economics (3) Microeconomics • With local nonsatiation the second condition becomes: If xi x∗ i then p · xi ≥ p · x∗ i . • I.e. with local non-satiation, x∗ i minimizes the expenditures given {xi : xi x∗ i }. 455 Second Fundamental Theorem of Welfare Economics (4) Microeconomics • Proposition [16.D.1] (Second Fundamental Theorem of Welfare Economics) – Consider an economy specified by ([Xi, i]I i=1, [Yj]J j=1, ¯ω), and suppose that every Yj is convex and every preference relation i is convex (the set {xi ∈ Xi : xi i xi} is convex for every xi ∈ Xi) and locally non-satiated. Then for every Pareto optimal allocation (x∗ , y∗ ) there exists a price vector p = 0 such that (x∗ , y∗ , p) is a price quasi-equilibrium with transfers. • Proof: See MWG, page 553. 456 Second Fundamental Theorem of Welfare Economics (5) Microeconomics • When is a price quasi-equilibrium with transfers a price equilibrium with transfers? • The example considered in MWG, Figure 15.B.10 (a) and on page 554, is a quasi-equilibrium but not an equilibrium. • In this example the wealth of consumer 1 is zero (hence, zero wealth problem). • We need a sufficient condition under which which ”xi x∗ i ⇒ p · xi ≥ wi” implies ”xi x∗ i ⇒ p · xi>wi”. 457 Second Fundamental Theorem of Welfare Economics (6) Microeconomics • Proposition [16.D.2] – Assume that Xi is convex and i is continuous. Suppose also that the consumption vector x∗ i ∈ Xi, the price vector p and the wealth level wi are such that xi i x∗ i implies p · xi≥wi. Then, if there is a consumption vector xi ∈ Xi such that p · xi < wi [a cheaper consumption for (p, wi)], it follows that xi x∗ i implies p · xi>wi. • Proof: See MWG page 555. See also MWG, Figure 16.D.3 (right). 458 Second Fundamental Theorem of Welfare Economics (7) Microeconomics • Proposition [16.D.3] – Suppose that for every i = 1, . . . , L, Xi is convex and i is continuous. Then, any price quasi-equilibrium with transfers that has (w1, . . . , wL) 0 is a price equilibrium with transfers. • Proof: See MWG page 556. 459 Quasiconcave Functions Motivation (1) Microeconomics • Motivation: Sufficient conditions for maximum for Kuhn Tucker problem: Suppose that there are no nonlinear equality constraints and each inequality constraint is given by a quasiconvex function. Suppose that the objective function satisfies f(x)(x − x) > 0 for any x and x with f(x ) > f(x). If x∗ satisfied the Kuhn-Tucker conditions, then x∗ is a global maximizer. (see Mas-Colell, Theorem [M.K.3]). • Jehle, Reny: Chapter A 1.4, 1.4. • Mas-Colell, Chapter M.C 460 Quasiconcave Functions Concave Functions (1) Microeconomics • Consider a convex subset A of Rn . • Definition - Concave Function: A function f : A → R is concave if f(νx + (1 − ν)x) ≥ νf(x ) + (1 − ν)f(x) , ν ∈ [0, 1]. If strict > holds then f is strictly concave; ν ∈ (0, 1) and x = x . This last equation can be rewritten with z = x − x and α = ν: f(x + αz) ≥ αf(x ) + (1 − α)f(x) . • If f is (strictly) concave then −f is (strictly) convex. 461 Quasiconcave Functions Concave Functions (2) Microeconomics • Theorem - Tangents and Concave Functions: If f is continuously differentiable and concave, then f(x ) ≤ f(x) + f(x)) · (x − x) (and vice versa). < holds if f is strict concave for all x = x . [Theorem M.C.1] • For the univariate case this implies that the tangent line is above the function graph of f(x); strictly for x = x with strict concave functions. 462 Quasiconcave Functions Concave Functions (3) Microeconomics Proof: • ⇒: For α ∈ (0, 1] the definition of a concave function implies: f(x ) = f(x + z) ≤ f(x) + f(x + αz) − f(x) α If f is differentiable the limit of the last term exists such that f(x + z) ≤ f(x) + f(x) · z 463 Quasiconcave Functions Concave Functions (4) Microeconomics Proof: • ⇐: Suppose that f(x + z) − f(x) ≤ f(x) · z for any non-concave function. Since f(.) is not concave f(x + z) − f(x) > f(x + αz) − f(x) α for some x, z and α ∈ (0, 1]. • Taking the limit results in f(x + z) − f(x) > f(x) · z, i.e. we arrive at a contradiction. 464 Quasiconcave Functions Concave Functions (5) Microeconomics • Theorem - Hessian and Concave Functions: If f is twice continuously differentiable and concave, then the Hessian matrix D2 f(x) is negative semidefinite; negative definite for strict concave functions (and vice versa). [Theorem M.C.2] 465 Quasiconcave Functions Concave Functions (6) Microeconomics Proof: • ⇒: A Taylor expansion of f(x ) around the point α = 0 results in f(x + αz) = f(x) + f(x) · (αz) + α2 2 (z · D2 (f(x + β(α)z))z) By the former theorem we know that f(x + αz) − f(x) − f(x) · (αz) ≤ 0 for concave functions ⇒ z D2 (f(x + β(α)z))z ≤ 0. For arbitrary small α we get z D2 (f(x))z ≤ 0. 466 Quasiconcave Functions Concave Functions (7) Microeconomics Proof: • ⇐: If the right hand side of f(x +αz)−f(x)− f(x)·(αz) = 0.5α2 (z D2 (f(x +β(α)z))z) is ≤ 0 then the left hand side. By the former theorem f is concave. 467 Quasiconcave Functions Quasiconcave Functions (1) Microeconomics • Definition - Quasiconcave Function: A function f : A → R is quasiconcave if f(νx + (1 − ν)x) ≥ min{f(x ), f(x)} , ν ∈ [0, 1]. If > holds it is said to be strict quasiconcave; ν ∈ (0, 1) and x = x . • Quasiconvex is defined by f(νx + (1 − ν)x) ≤ max{f(x ), f(x)}. If f is quasiconcave than −f is quasiconvex. • If f is concave then f is quasiconcave but not vice versa. E.g. f(x) = √ x for x > 0 is concave and also quasiconcave. x3 is quasiconcave but not concave. 468 Quasiconcave Functions Quasiconcave Functions (2) Microeconomics • Transformation property: Positive monotone transformations of quasiconcave functions result in a quasiconcave function. • Definition - Superior Set: S(x) := {x ∈ A|f(x ) ≥ f(x)} is called superior set of x (upper contour set of x). • Note that if f(xν ) ≥ min{f(x ), f(x )}, then if f(x ) ≥ t and f(x ) ≥ t this implies that f(xν ) ≥ t; where t = f(x). 469 Quasiconcave Functions Quasiconcave Functions (3) Microeconomics • Theorem - Quasiconcave Function and Convex Sets: The function f is quasiconcave if and only if S(x) is convex for all x ∈ A. 470 Quasiconcave Functions Quasiconcave Functions (4) Microeconomics Proof: • Sufficient condition ⇒: If f is quasiconcave then S(x) is convex. Consider x1 and x2 in S(x). We need to show that f(xν ) in S(x); f(x) = t. • Since f(x1 ) ≥ t and f(x2 ) ≥ t, the quasiconcave f implies f(xν ) ≥ min{f(x1 ), f(x2 )} ≥ t. • Therefore f(xν ) ∈ S(x); i.e. the set S(x) is convex. 471 Quasiconcave Functions Quasiconcave Functions (5) Microeconomics Proof: • Necessary condition ⇐: If S(x) is convex then f(x) has to be quasiconcave. W.l.g. assume that f(x1 ) ≥ f(x2 ), x1 and x2 in A. • By assumption S(x) is convex, such that S(x2 ) is convex. Since f(x1 ) ≥ f(x2 ), we get x1 ∈ S(x2 ) and xν ∈ S(x2 ). • From the definition of S(x2 ) we conclude that f(xν ) ≥ f(x2 ) = min{f(x1 ), f(x2 )}. • Therefore f(x) has to be quasiconcave. 472 Quasiconcave Functions Quasiconcave Functions (6) Microeconomics • Theorem - Gradients and Quasiconcave Functions: If f is continuously differentiable and quasiconcave, then f(x) · (x − x) ≥ 0 whenever f(x ) ≥ f(x) (and vice versa). [Theorem M.C.3] • If f(x) · (x − x) > 0 whenever f(x ) ≥ f(x) and x = x then f(x) is strictly quasiconcave. If f(x) is strictly quasiconcave and if f(x) = 0 for all x ∈ A, then f(x) · (x − x) > 0 whenever f(x ) ≥ f(x) and x = x . 473 Quasiconcave Functions Quasiconcave Functions (7) Microeconomics Proof: • ⇒: For f(x ) ≥ f(x) and α ∈ (0, 1] the definition of a quasiconcave function implies: f(x + α(x − x)) − f(x) α ≥ 0 If f is differentiable, then the limit exists such that f(x) · z ≥ 0 474 Quasiconcave Functions Quasiconcave Functions (8) Microeconomics Proof: • ⇐: Suppose that f(x) · z ≥ 0 holds but f is not quasiconcave. Then f(x + αz) − f(x) < 0 for some x, z and α ∈ (0, 1]. Such that (f(x + αz) − f(x))/α < 0. Taking the limit results in a contradiction. 475 Quasiconcave Functions Quasiconcave Functions (9) Microeconomics • Theorem - Hessian Matrix and Quasiconcave Functions: Suppose f is twice continuously differentiable. f(x) is quasiconcave if and only if D2 (f(x)) is negative semidefinite in the subspace {z| f(x) · z = 0}. I.e. z D2 (f(x))z ≤ 0 whenever f(x) · z = 0. [Theorem M.C.4] • If the Hessian D2 (f(x)) is negative definite in the subspace {z| f(x) · z = 0} for every x ∈ A then f(x) is strictly quasiconcave. 476 Quasiconcave Functions Quasiconcave Functions (10) Microeconomics Proof: • ⇒: If f is quasiconcave then whenever f(xν ) ≥ f(x), so f(x) · (αz) ≥ 0 has to hold. • Thus f(x1 ) − f(x) ≤ 0 and the above theorem imply: f(x) · (z) ≤ 0, where z = x1 − x. • A first order Taylor series expansion of f in α (at α = 0) results in f(x + αz) = f(x) + f(x)αz + α2 2 · z D2 f(x + β(α)z)z . 477 Quasiconcave Functions Quasiconcave Functions (11) Microeconomics Proof: • Apply this to x1 , x with f(x1 ) ≤ f(x): f(x + αz) − f(x) − f(x)αz = α2 2 · z D2 f(x + β(α)z)z. • If z = x1 − x fulfills f(x)(x1 − x) = 0 the above inequality still has to hold. • This implies α2 /2z D2 f(x + β(α)z)z ≤ 0. 478 Quasiconcave Functions Quasiconcave Functions (12) Microeconomics Proof: • To fulfill this requirement on the subspace {z| f(x) · z = 0}, where f(x)αz = 0, this requires a negative definite Hessian of f(x). • ⇐: In the above equation a negative semidefinite Hessian implies that . . . . 479 Envelope Theorem (1) Microeconomics • Consider f(x; q), x are variables in RN and q are parameters in RS . • We look at the constrained maximization problem max x f(x; q) s.t.gm(x; q) ≤ bm m = 1, . . . , M. • Assume that the solution of this optimization problem x∗ = x(q) is at least locally differentiable function (in a neighborhood of a ¯q considered). • v(q) = f(x(q); q) is the maximum value function associated with this problem. 480 Envelope Theorem (2) Microeconomics • With no constraints (M = 0) and S, N = 1 the chain rule yields: d dq v(¯q) = ∂f(x(¯q); ¯q) ∂x ∂x(¯q) ∂q + ∂f(x(¯q); ¯q) ∂q . • With an unconstrained maximization problem the first order condition ∂f(x(¯q);¯q) ∂x = 0 results in d dq v(¯q) = ∂f(x(¯q); ¯q) ∂q . 481 Envelope Theorem (3) Microeconomics [T. M.L.1] Consider the value function v(q) for the above constrained maximization problem. Assume that v(q) is differentiable at ¯q ∈ RS and (λ1, . . . , λM) are the Lagrange multipliers associated with the maximizer solution x(q) at ¯q. In addition the inequality constraints are remain unaltered in a neighborhood of ¯q. Then ∂v(¯q) ∂qs = ∂f(x(¯q); ¯q) ∂qs − M m=1 λm ∂gm(x(¯q); ¯q) ∂qs . For s = 1, . . . , S. 482 Envelope Theorem (4) Microeconomics Proof: • Let x(.) stand for the maximizer of the function f(.) and v(q) = f(x(q), q) for all q. The chain rule yields: ∂v(¯q) dqs = N n=1 ∂f(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • The first order conditions tell us ∂f(x(¯q); ¯q) ∂xn = M m=1 λm ∂gm(x(¯q); ¯q) ∂xn . 483 Envelope Theorem (5) Microeconomics Proof: • In addition we observe N n=1 ∂gm(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂gm(¯q) ∂qs = 0. if a constraint is binding; if not the multiplier λm is zero. 484 Envelope Theorem (6) Microeconomics Proof: • Plugging in and changing the order of summation results in : ∂v(¯q) dqs = M m=1 λm N n=1 ∂gm(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • and ∂v(¯q) dqs = − M m=1 λm ∂gm(x(¯q); ¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • Remark: remember that the Lagrangian of the problem is L(x, λ; q) = f(x; q) − m λmgm(x; q). Hence we get ∂v(¯q) dqs by means of the partial derivative of the Lagrangian with respect to ql, evaluated at ¯q. 485 Quasiconcave Functions Motivation (1) Microeconomics • Motivation: Sufficient conditions for maximum for Kuhn Tucker problem: Suppose that there are no nonlinear equality constraints and each inequality constraint is given by a quasiconvex function. Suppose that the objective function satisfies f(x)(x − x) > 0 for any x and x with f(x ) > f(x). If x∗ satisfied the Kuhn-Tucker conditions, then x∗ is a global maximizer. (see Mas-Colell, Theorem [M.K.3]). • Jehle, Reny: Chapter A 1.4, 1.4. • Mas-Colell, Chapter M.C 486 Quasiconcave Functions Concave Functions (1) Microeconomics • Consider a convex subset A of Rn . • Definition - Concave Function: A function f : A → R is concave if f(νx + (1 − ν)x) ≥ νf(x ) + (1 − ν)f(x) , ν ∈ [0, 1]. If strict > holds then f is strictly concave; ν ∈ (0, 1) and x = x . This last equation can be rewritten with z = x − x and α = ν: f(x + αz) ≥ αf(x ) + (1 − α)f(x) . • If f is (strictly) concave then −f is (strictly) convex. 487 Quasiconcave Functions Concave Functions (2) Microeconomics • Theorem - Tangents and Concave Functions: If f is continuously differentiable and concave, then f(x ) ≤ f(x) + f(x)) · (x − x) (and vice versa). < holds if f is strict concave for all x = x . [Theorem M.C.1] • For the univariate case this implies that the tangent line is above the function graph of f(x); strictly for x = x with strict concave functions. 488 Quasiconcave Functions Concave Functions (3) Microeconomics Proof: • ⇒: For α ∈ (0, 1] the definition of a concave function implies: f(x ) = f(x + z) ≤ f(x) + f(x + αz) − f(x) α If f is differentiable the limit of the last term exists such that f(x + z) ≤ f(x) + f(x) · z 489 Quasiconcave Functions Concave Functions (4) Microeconomics Proof: • ⇐: Suppose that f(x + z) − f(x) ≤ f(x) · z for any non-concave function. Since f(.) is not concave f(x + z) − f(x) > f(x + αz) − f(x) α for some x, z and α ∈ (0, 1]. • Taking the limit results in f(x + z) − f(x) > f(x) · z, i.e. we arrive at a contradiction. 490 Quasiconcave Functions Concave Functions (5) Microeconomics • Theorem - Hessian and Concave Functions: If f is twice continuously differentiable and concave, then the Hessian matrix D2 f(x) is negative semidefinite; negative definite for strict concave functions (and vice versa). [Theorem M.C.2] 491 Quasiconcave Functions Concave Functions (6) Microeconomics Proof: • ⇒: A Taylor expansion of f(x ) around the point α = 0 results in f(x + αz) = f(x) + f(x) · (αz) + α2 2 (z · D2 (f(x + β(α)z))z) By the former theorem we know that f(x + αz) − f(x) − f(x) · (αz) ≤ 0 for concave functions ⇒ z D2 (f(x + β(α)z))z ≤ 0. For arbitrary small α we get z D2 (f(x))z ≤ 0. 492 Quasiconcave Functions Concave Functions (7) Microeconomics Proof: • ⇐: If the right hand side of f(x +αz)−f(x)− f(x)·(αz) = 0.5α2 (z D2 (f(x +β(α)z))z) is ≤ 0 then the left hand side. By the former theorem f is concave. 493 Quasiconcave Functions Quasiconcave Functions (1) Microeconomics • Definition - Quasiconcave Function: A function f : A → R is quasiconcave if f(νx + (1 − ν)x) ≥ min{f(x ), f(x)} , ν ∈ [0, 1]. If > holds it is said to be strict quasiconcave; ν ∈ (0, 1) and x = x . • Quasiconvex is defined by f(νx + (1 − ν)x) ≤ max{f(x ), f(x)}. If f is quasiconcave than −f is quasiconvex. • If f is concave then f is quasiconcave but not vice versa. E.g. f(x) = √ x for x > 0 is concave and also quasiconcave. x3 is quasiconcave but not concave. 494 Quasiconcave Functions Quasiconcave Functions (2) Microeconomics • Transformation property: Positive monotone transformations of quasiconcave functions result in a quasiconcave function. • Definition - Superior Set: S(x) := {x ∈ A|f(x ) ≥ f(x)} is called superior set of x (upper contour set of x). • Note that if f(xν ) ≥ min{f(x ), f(x )}, then if f(x ) ≥ t and f(x ) ≥ t this implies that f(xν ) ≥ t; where t = f(x). 495 Quasiconcave Functions Quasiconcave Functions (3) Microeconomics • Theorem - Quasiconcave Function and Convex Sets: The function f is quasiconcave if and only if S(x) is convex for all x ∈ A. 496 Quasiconcave Functions Quasiconcave Functions (4) Microeconomics Proof: • Sufficient condition ⇒: If f is quasiconcave then S(x) is convex. Consider x1 and x2 in S(x). We need to show that f(xν ) in S(x); f(x) = t. • Since f(x1 ) ≥ t and f(x2 ) ≥ t, the quasiconcave f implies f(xν ) ≥ min{f(x1 ), f(x2 )} ≥ t. • Therefore f(xν ) ∈ S(x); i.e. the set S(x) is convex. 497 Quasiconcave Functions Quasiconcave Functions (5) Microeconomics Proof: • Necessary condition ⇐: If S(x) is convex then f(x) has to be quasiconcave. W.l.g. assume that f(x1 ) ≥ f(x2 ), x1 and x2 in A. • By assumption S(x) is convex, such that S(x2 ) is convex. Since f(x1 ) ≥ f(x2 ), we get x1 ∈ S(x2 ) and xν ∈ S(x2 ). • From the definition of S(x2 ) we conclude that f(xν ) ≥ f(x2 ) = min{f(x1 ), f(x2 )}. • Therefore f(x) has to be quasiconcave. 498 Quasiconcave Functions Quasiconcave Functions (6) Microeconomics • Theorem - Gradients and Quasiconcave Functions: If f is continuously differentiable and quasiconcave, then f(x) · (x − x) ≥ 0 whenever f(x ) ≥ f(x) (and vice versa). [Theorem M.C.3] • If f(x) · (x − x) > 0 whenever f(x ) ≥ f(x) and x = x then f(x) is strictly quasiconcave. If f(x) is strictly quasiconcave and if f(x) = 0 for all x ∈ A, then f(x) · (x − x) > 0 whenever f(x ) ≥ f(x) and x = x . 499 Quasiconcave Functions Quasiconcave Functions (7) Microeconomics Proof: • ⇒: For f(x ) ≥ f(x) and α ∈ (0, 1] the definition of a quasiconcave function implies: f(x + α(x − x)) − f(x) α ≥ 0 If f is differentiable, then the limit exists such that f(x) · z ≥ 0 500 Quasiconcave Functions Quasiconcave Functions (8) Microeconomics Proof: • ⇐: Suppose that f(x) · z ≥ 0 holds but f is not quasiconcave. Then f(x + αz) − f(x) < 0 for some x, z and α ∈ (0, 1]. Such that (f(x + αz) − f(x))/α < 0. Taking the limit results in a contradiction. 501 Quasiconcave Functions Quasiconcave Functions (9) Microeconomics • Theorem - Hessian Matrix and Quasiconcave Functions: Suppose f is twice continuously differentiable. f(x) is quasiconcave if and only if D2 (f(x)) is negative semidefinite in the subspace {z| f(x) · z = 0}. I.e. z D2 (f(x))z ≤ 0 whenever f(x) · z = 0. [Theorem M.C.4] • If the Hessian D2 (f(x)) is negative definite in the subspace {z| f(x) · z = 0} for every x ∈ A then f(x) is strictly quasiconcave. 502 Quasiconcave Functions Quasiconcave Functions (10) Microeconomics Proof: • ⇒: If f is quasiconcave then whenever f(xν ) ≥ f(x), so f(x) · (αz) ≥ 0 has to hold. • Thus f(x1 ) − f(x) ≤ 0 and the above theorem imply: f(x) · (z) ≤ 0, where z = x1 − x. • A first order Taylor series expansion of f in α (at α = 0) results in f(x + αz) = f(x) + f(x)αz + α2 2 · z D2 f(x + β(α)z)z . 503 Quasiconcave Functions Quasiconcave Functions (11) Microeconomics Proof: • Apply this to x1 , x with f(x1 ) ≤ f(x): f(x + αz) − f(x) − f(x)αz = α2 2 · z D2 f(x + β(α)z)z. • If z = x1 − x fulfills f(x)(x1 − x) = 0 the above inequality still has to hold. • This implies α2 /2z D2 f(x + β(α)z)z ≤ 0. 504 Quasiconcave Functions Quasiconcave Functions (12) Microeconomics Proof: • To fulfill this requirement on the subspace {z| f(x) · z = 0}, where f(x)αz = 0, this requires a negative definite Hessian of f(x). • ⇐: In the above equation a negative semidefinite Hessian implies that . . . . 505 Envelope Theorem (1) Microeconomics • Consider f(x; q), x are variables in RN and q are parameters in RS . • We look at the constrained maximization problem max x f(x; q) s.t.gm(x; q) ≤ bm m = 1, . . . , M. • Assume that the solution of this optimization problem x∗ = x(q) is at least locally differentiable function (in a neighborhood of a ¯q considered). • v(q) = f(x(q); q) is the maximum value function associated with this problem. 506 Envelope Theorem (2) Microeconomics • With no constraints (M = 0) and S, N = 1 the chain rule yields: d dq v(¯q) = ∂f(x(¯q); ¯q) ∂x ∂x(¯q) ∂q + ∂f(x(¯q); ¯q) ∂q . • With an unconstrained maximization problem the first order condition ∂f(x(¯q);¯q) ∂x = 0 results in d dq v(¯q) = ∂f(x(¯q); ¯q) ∂q . 507 Envelope Theorem (3) Microeconomics [T. M.L.1] Consider the value function v(q) for the above constrained maximization problem. Assume that v(q) is differentiable at ¯q ∈ RS and (λ1, . . . , λM) are the Lagrange multipliers associated with the maximizer solution x(q) at ¯q. In addition the inequality constraints are remain unaltered in a neighborhood of ¯q. Then ∂v(¯q) ∂qs = ∂f(x(¯q); ¯q) ∂qs − M m=1 λm ∂gm(x(¯q); ¯q) ∂qs . For s = 1, . . . , S. 508 Envelope Theorem (4) Microeconomics Proof: • Let x(.) stand for the maximizer of the function f(.) and v(q) = f(x(q), q) for all q. The chain rule yields: ∂v(¯q) dqs = N n=1 ∂f(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • The first order conditions tell us ∂f(x(¯q); ¯q) ∂xn = M m=1 λm ∂gm(x(¯q); ¯q) ∂xn . 509 Envelope Theorem (5) Microeconomics Proof: • In addition we observe N n=1 ∂gm(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂gm(¯q) ∂qs = 0. if a constraint is binding; if not the multiplier λm is zero. 510 Envelope Theorem (6) Microeconomics Proof: • Plugging in and changing the order of summation results in : ∂v(¯q) dqs = M m=1 λm N n=1 ∂gm(x(¯q); ¯q) ∂xn ∂xn(¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • and ∂v(¯q) dqs = − M m=1 λm ∂gm(x(¯q); ¯q) ∂qs + ∂f(x(¯q); ¯q) ∂qs . • Remark: remember that the Lagrangian of the problem is L(x, λ; q) = f(x; q) − m λmgm(x; q). Hence we get ∂v(¯q) dqs by means of the partial derivative of the Lagrangian with respect to ql, evaluated at ¯q. 511