Foundation of Mathematical Economics Solutions
Short Description
Foundation of Mathematical Economics by Michael Carter Solution manual The solutions manual contains detailed answers t...
Description
Solutions Manual Foundations of Mathematical Economics Michael Carter November 15, 2002
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 1: Sets and Spaces 1.1 { 1, 3, 5, 7 . . . } or { ð â ð : ð is odd } 1.2 Every ð¥ â ðŽ also belongs to ðµ. Every ð¥ â ðµ also belongs to ðŽ. Hence ðŽ, ðµ have precisely the same elements. 1.3 Examples of ï¬nite sets are â the letters of the alphabet { A, B, C, . . . , Z } â the set of consumers in an economy â the set of goods in an economy â the set of players in a game. Examples of inï¬nite sets are â the real numbers â â the natural numbers ð â the set of all possible colors â the set of possible prices of copper on the world market â the set of possible temperatures of liquid water. 1.4 ð = { 1, 2, 3, 4, 5, 6 }, ðž = { 2, 4, 6 }. 1.5 The player set is ð = { Jenny, Chris }. Their action spaces are ðŽð = { Rock, Scissors, Paper }
ð = Jenny, Chris
1.6 The set of players is ð = {1, 2, . . . , ð }. The strategy space of each player is the set of feasible outputs ðŽð = { ðð â â+ : ðð †ðð } where ðð is the output of dam ð. 1.7 The player set is ð = {1, 2, 3}. There are 23 = 8 coalitions, namely ð«(ð ) = {â
, {1}, {2}, {3}, {1, 2}, {1, 3}, {2, 3}, {1, 2, 3}} There are 2
10
coalitions in a ten player game.
/ ð ⪠ð . This implies ð¥ â / ð and ð¥ â / ð, 1.8 Assume that ð¥ â (ð ⪠ð )ð . That is ð¥ â or ð¥ â ð ð and ð¥ â ð ð. Consequently, ð¥ â ð ð â© ð ð . Conversely, assume ð¥ â ð ð â© ð ð . This implies that ð¥ â ð ð and ð¥ â ð ð . Consequently ð¥ â / ð and ð¥ â / ð and therefore ð¥â / ð ⪠ð . This implies that ð¥ â (ð ⪠ð )ð . The other identity is proved similarly. 1.9
âª
ð=ð
ðâð
â©
ð=â
ðâð
1
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
1
-1
ð¥2
0
1
ð¥1
-1 Figure 1.1: The relation { (ð¥, ðŠ) : ð¥2 + ðŠ 2 = 1 } 1.10 The sample space of a single coin toss is { ð», ð }. The set of possible outcomes in three tosses is the product { {ð», ð } à {ð», ð } à {ð», ð } = (ð», ð», ð»), (ð», ð», ð ), (ð», ð, ð»), } (ð», ð, ð ), (ð, ð», ð»), (ð, ð», ð ), (ð, ð, ð»), (ð, ð, ð ) A typical outcome is the sequence (ð», ð», ð ) of two heads followed by a tail. 1.11 ð â© âð+ = {0} where 0 = (0, 0, . . . , 0) is the production plan using no inputs and producing no outputs. To see this, ï¬rst note that 0 is a feasible production plan. Therefore, 0 â ð . Also, 0 â âð+ and therefore 0 â ð â© âð+ . To show that there is no other feasible production plan in âð+ , we assume the contrary. That is, we assume there is some feasible production plan y â âð+ â {0}. This implies the existence of a plan producing a positive output with no inputs. This technological infeasible, so that ðŠ â / ð. 1.12
1. Let x â ð (ðŠ). This implies that (ðŠ, âx) â ð . Let xⲠ⥠x. Then (ðŠ, âxâ² ) †(ðŠ, âx) and free disposability implies that (ðŠ, âxâ² ) â ð . Therefore xâ² â ð (ðŠ).
2. Again assume x â ð (ðŠ). This implies that (ðŠ, âx) â ð . By free disposal, (ðŠ â² , âx) â ð for every ðŠ Ⲡ†ðŠ, which implies that x â ð (ðŠ â² ). ð (ðŠ â² ) â ð (ðŠ). 1.13 The domain of â ðŠ1 or ð¥1 = ðŠ1 and ð¥2 > ðŠ2 Since all elements ð¥ â â2 are comparable, â» is complete; it is a total order. 1.35 Assume â¿ð is complete for every ð. Then for every ð¥, ðŠ â ð and for all ð = 1, 2, . . . , ð, either ð¥ð â¿ð ðŠð or ðŠð â¿ð ð¥ð or both. Either ð¥ð âŒð ðŠð for all ð Then deï¬ne ð¥ ⌠ðŠ. ð¥ð ââŒð ðŠð for some ð Let ð be the ï¬rst individual with a strict preference, that is ð = minð (ð¥ð â⌠ðŠð ). (Completeness of â¿ð ensures that ð is deï¬ned). Then deï¬ne ð¥ â» ðŠ if ð¥ð â»ð ðŠð ðŠ â» ð¥ otherwise 1.36 Let ð, ð and ð be subsets of a ï¬nite set ð. Set inclusion â is reï¬exive since ð â ð. transitive since ð â ð and ð â ð implies ð â ð . anti-symmetric since ð â ð and ð â ð implies ð = ð Therefore â is a partial order. 1.37 Assume ð¥ and ðŠ are both least upper bounds of ðŽ. That is ð¥ â¿ ð for all ð â ðŽ and ðŠ â¿ ð for all ð â ðŽ. Further, if ð¥ is a least upper bound, ðŠ â¿ ð¥. If ðŠ is a least upper bound, ð¥ â¿ ðŠ. By anti-symmetry, ð¥ = ðŠ. 1.38 ð¥ ⌠ðŠ =â ð¥ â¿ ðŠ and ðŠ â¿ ð¥ which implies that ð¥ = ðŠ by antisymmetry. Each equivalence class ⌠(ð¥) = { ðŠ â ð : ðŠ ⌠ð¥ } comprises just a single element ð¥. 1.39 max ð«(ð) = ð and min ð«(ð) = â
. 1.40 The subset {2, 4, 8} forms a chain. More generally, the set of integer powers of a given number { ð, ð2 , ð3 , . . . } forms a chain. 1.41 Assume ð¥ and ðŠ are maximal elements of the chain ðŽ. Then ð¥ â¿ ð for all ð â ðŽ and in particular ð¥ â¿ ðŠ. Similarly, ðŠ â¿ ð for all ð â ðŽ and in particular ðŠ â¿ ð¥. Since â¿ is anti-symmetric, ð¥ = ðŠ. 1.42
1. By assumption, for every ð¡ â ð â ð , âº(ð¡) is a nonempty ï¬nite chain. Hence, it has a unique maximal element, ð(ð¡).
2. Let ð¡ be any node. Either ð¡ is an initial node or ð¡ has a unique predecessor ð(ð¡). Either ð(ð¡) is an initial node, or it has a unique predecessor ð(ð(ð¡)). Continuing in this way, we trace out a unique path from ð¡ back to an initial node. We can be sure of eventually reaching an initial node since ð is ï¬nite. 1.43 (1, 2) âš (3, 1) = (3, 2) and (1, 2) ⧠(3, 2) = (1, 2) 6
Solutions for Foundations of Mathematical Economics 1.44
c 2001 Michael Carter â All rights reserved
1. ð¥âšðŠ is an upper bound for { ð¥, ðŠ }, that is xâšy â¿ ð¥ and xâšy â¿ ðŠ. Similarly, ð¥ âš ðŠ is a lower bound for { ð¥, ðŠ }.
2. Assume ð¥ â¿ ðŠ. Then ð¥ is an upper bound for { ð¥, ðŠ }, that is ð¥ â¿ ð¥ âš ðŠ. If ð is any upper bound for { ð¥, ðŠ }, then ð â¿ ð¥. Therefore, ð¥ is the least upper bound for { ð¥, ðŠ }. Similarly, ðŠ is a lower bound for { ð¥, ðŠ }, and is greater than any other lower bound. Conversely, assume ð¥ âš ðŠ = ð¥. Then ð¥ is an upper bound for { ð¥, ðŠ }, that is ð¥ â¿ ðŠ. 3. Using the preceding equivalence ð¥ â¿ ð¥ ⧠ðŠ =â ð¥ âš (ð¥ ⧠ðŠ) = ð¥ ð¥ âš ðŠ â¿ ð¥ =â (ð¥ âš ðŠ) ⧠ð¥ = ð¥ 1.45 A chain ð is a complete partially ordered set. For every ð¥, ðŠ â ð with ð¥ â= ðŠ, either ð¥ â» ðŠ or ðŠ â» ð¥. Therefore, deï¬ne the meet and join by { ðŠ if ð¥ â» ðŠ ð¥â§ðŠ = ð¥ if ðŠ â» ð¥ { ð¥ if ð¥ â» ðŠ ð¥âšðŠ = ðŠ if ðŠ â» ð¥ ð is a lattice with these operations. 1.46 Assume ð1 and ð2 are lattices, and let ð = ð1 à ð2 . Consider any two elements x = (ð¥1 , ð¥2 ) and y = (ðŠ1 , ðŠ2 ) in ð. Since ð1 and ð2 are lattices, ð1 = ð¥1 âš ðŠ1 â ð1 and ð2 = ð¥2 âš ðŠ2 â ð2 , so that b = (ð1 , ð2 ) = (ð¥1 âš ðŠ1 , ð¥2 âš ðŠ2 ) â ð. Furthermore b â¿ x and b â¿ y in the natural product order, so that b is an upper bound for the Ë = (Ëð1 , Ëð2 ) of {x, y} must have ðð â¿ð ð¥ð and ðð â¿ð ðŠð , {x, y}. Every upper bound b Ë â¿ b. Therefore, b is the least upper bound of {x, y}, that is b = x âš y. so that b Similarly, x ⧠y = (ð¥1 ⧠ðŠ1 , ð¥2 ⧠ðŠ2 ). 1.47 Let ð be a subset of ð and let ð â = { ð¥ â ð : ð¥ â¿ ð for every ð â ð } be the set of upper bounds of ð. Then ð¥â â ð â â= â
. By assumption, ð â has a greatest lower bound ð. Since every ð â ð is a lower bound of ð â , ð â¿ ð for every ð â ð. Therefore ð is an upper bound of ð. Furthermore, ð is the least upper bound of ð, since ð ⟠ð¥ for every ð¥ â ð â . This establishes that every subset of ð also has a least upper bound. In particular, every pair of elements has a least upper and a greatest lower bound. Consequently ð is a complete lattice. 1.48 Without loss of generality, we will prove the closed interval case. Let [ð, ð] be an interval in a lattice ð¿. Recall that ð = inf[ð, ð] and ð = sup[ð, ð]. Choose any ð¥, ðŠ in [ð, ð] â ð¿. Since ð¿ is a lattice, ð¥ âš ðŠ â ð¿ and ð¥ âš ðŠ = sup{ ð¥, ðŠ } ⟠ð Therefore ð¥ âš ðŠ â [ð, ð]. Similarly, ð¥ ⧠ðŠ â [ð, ð]. [ð, ð] is a lattice. Similarly, for any subset ð â [ð, ð] â ð¿, sup ð â ð¿ if ð¿ is complete. Also, sup ð ⟠ð = sup[ð, ð]. Therefore sup ð â [ð, ð]. Similarly inf ð â [ð, ð] so that [ð, ð] is complete.
7
Solutions for Foundations of Mathematical Economics 1.49
c 2001 Michael Carter â All rights reserved
1. The strong set order â¿ð is antisymmetric Let ð1 , ð2 â ð with ð1 â¿ð ð2 and ð2 â¿ð ð1 . Choose ð¥1 â ð1 and ð¥2 â ð2 . Since ð1 â¿ð ð2 , ð¥1 âš ð¥2 â ð1 and ð¥1 ⧠ð¥2 â ð2 . On the other hand, since ð2 â¿ ð1 , ð¥1 = (ð¥1 âš (ð¥1 ⧠ð¥2 ) â ð2 and ð¥2 = ð¥2 ⧠(ð¥1 âš ð¥2 ) â ð1 (Exercise 1.44. Therefore ð1 = ð2 and â¿ð is antisymmetric. transitive Let ð1 , ð2 , ð3 â ð with ð1 â¿ð ð2 and ð2 â¿ð ð3 . Choose ð¥1 â ð1 , ð¥2 â ð2 and ð¥3 â ð3 . Since ð1 â¿ð ð2 and ð2 â¿ð ð3 , ð¥1 âš ð¥2 and ð¥2 ⧠ð¥3 are in ð2 . Therefore ðŠ2 = ð¥1 âš (ð¥2 ⧠ð¥3 ) â ð2 which implies ) ( ð¥1 âš ð¥3 = ð¥1 âš (ð¥2 ⧠ð¥3 ) âš ð¥3 ( ) = ð¥1 âš (ð¥2 ⧠ð¥3 ) âš ð¥3 = ðŠ2 âš ð¥3 â ð3 since ð2 â¿ð ð3 . Similarly ð§2 = (ð¥1 âš ð¥2 ) ⧠ð¥3 â ð2 and ( ) ð¥1 ⧠ð¥3 = ð¥1 ⧠(ð¥1 âš ð¥2 ) ⧠ð¥3 ) ( = ð¥1 ⧠(ð¥1 âš ð¥2 ) ⧠ð¥3 = ð¥1 ⧠ð§2 â ð1 Therefore, ð1 â¿ð ð3 .
2. ð â¿ð ð if and only if, for every ð¥1 , ð¥2 â ð, ð¥1 âš ð¥2 â ð and ð¥1 ⧠ð¥2 â ð, which is the case if and only if ð is a sublattice. 3. Let ð¿(ð) denote the set of all sublattices of ð. We have shown that â¿ð is reï¬exive, transitive and antisymmetric on ð¿(ð). Hence, it is a partial order on ð¿(ð). 1.50 Assume ð1 â¿ð ð2 . For any ð¥1 â ð1 and ð¥2 â ð2 , ð¥1 âš ð¥2 â ð1 and ð¥1 ⧠ð¥2 â ð2 . Therefore sup ð1 â¿ ð¥1 âš ð¥2 â¿ ð¥2
for every ð¥2 â ð2
which implies that sup ð1 â¿ sup ð2 . Similarly inf ð2 ⟠ð¥1 ⧠ð¥2 ⟠ð¥1
for every ð¥1 â ð1
which implies that inf ð2 ⟠inf ð1 . Note that completeness ensures the existence of sup ð and inf ð respectively. 1.51 An argument analogous to the preceding exercise establishes =â . (Completeness is not required, since for any interval ð = inf[ð, ð] and ð = sup[ð, ð]). To establish the converse, assume that ð1 = [ð1 , ð1 ] and ð2 = [ð2 , ð2 ]. Consider any ð¥1 â ð1 and ð¥2 â ð2 . There are two cases. Case 1. ð¥1 â¿ ð¥2 Since ð is a chain, ð¥1 âš ð¥2 = ð¥1 â ð1 . ð¥1 ⧠ð¥2 = ð¥2 â ð2 . Case 2. ð¥1 ⺠ð¥2 Since ð is a chain, ð¥1 âš ð¥2 = ð¥2 . Now ð1 ⟠ð¥1 ⺠ð¥2 ⟠ð2 ⟠ð2 . Therefore, ð¥2 = ð¥1 âš ð¥2 â ð1 . Similarly ð2 ⟠ð1 ⟠ð¥1 ⺠ð¥2 ⟠ð2 . Therefore ð¥1 ⧠ð¥2 = ð¥1 â ð2 . We have shown that ð1 â¿ð ð2 in both cases. 1.52 Assume that â¿ is a complete relation on ð. This means that for every ð¥, ðŠ â ð, either ð¥ â¿ ðŠ or ðŠ â¿ ð¥. In particular, letting ð¥ = ðŠ, ð¥ â¿ ð¥ for ð¥ â ð. â¿ is reï¬exive. 8
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.53 Anti-symmetry implies that each indiï¬erence class contains a single element. If the consumerâs preference relation was anti-symmetric, there would be no baskets of goods between which the consumer was indiï¬erent. Each indiï¬erence curve which consist a single point. 1.54 We previously showed (Exercise 1.27) that every best element is maximal. To prove the converse, assume that ð¥ is maximal in the weakly ordered set ð. We have to show that ð¥ â¿ ðŠ for all ðŠ â ð. Assume otherwise, that is assume there is some ðŠ â ð for which ð¥ ââ¿ ðŠ. Since â¿ is complete, this implies that ðŠ â» ð¥ which contradicts the assumption that ð¥ is maximal. Hence we conclude that ð¥ â¿ ðŠ for ðŠ â ð and ð¥ is a best element. 1.55 False. A chain has at most one maximal element (Exercise 1.41). Here, uniqueness is ensured by anti-symmetry. A weakly ordered set in which the order is not antisymmetric may have multiple maximal and best elements. For example, ð and ð are both best elements in the weakly ordered set {ð ⌠ð â» ð}. 1.56
1. For every ð¥ â ð, either ð¥ â¿ ðŠ =â ð¥ â â¿(ðŠ) or ðŠ â¿ ð¥ =â ð¥ â âŸ(ðŠ) since â¿ is complete. Consequently, â¿(ðŠ) ⪠âº(ðŠ) = ð If ð¥ â â¿(ðŠ) â© âŸ(ðŠ), then ð¥ â¿ ðŠ and ðŠ â¿ ð¥ so that ð¥ ⌠ðŠ and ð¥ â ðŒðŠ .
2. For every ð¥ â ð, either ð¥ â¿ ðŠ =â ð¥ â â¿(ðŠ) or ðŠ â» ð¥ =â ð¥ â âº(ðŠ) since â¿ is complete. Consequently, â¿(ðŠ) ⪠âº(ðŠ) = ð and â¿(ðŠ) â© âº(ðŠ) = â
. 3. For every ðŠ â ð, â»(ðŠ) and ðŒðŠ partition â¿(ðŠ) and therefore â»(ðŠ), ðŒðŠ and âº(ðŠ) partition ð. 1.57 Assume ð¥ â¿ ðŠ and ð§ â â¿(ð¥). Then ð§ â¿ ð¥ â¿ ðŠ by transitivity. Therefore ð§ â â¿(ðŠ). This shows that â¿(ð¥) â â¿(ðŠ). Similarly, assume ð¥ â» ðŠ and ð§ â â»(ð¥). Then ð§ â» ð¥ â» ðŠ by transitivity. Therefore ð§ â â»(ðŠ). This shows that â¿(ð¥) â â¿(ðŠ). To show that â¿(ð¥) â= â¿(ðŠ), observe that ð¥ â â»(ðŠ) but that ð¥ â / â»(ð¥) 1.58 Every ï¬nite ordered set has a least one maximal element (Exercise 1.28). 1.59 Kreps (1990, p.323), Luenberger (1995, p.170) and Mas-Colell et al. (1995, p.313) adopt the weak Pareto order, whereas Varian (1992, p.323) distinguishes the two orders. Osborne and Rubinstein (1994, p.7) also distinguish the two orders, utilizing the weak order in deï¬ning the core (Chapter 13) but the strong Pareto order in the Nash bargaining solution (Chapter 15). 1.60 Assume that a group ð is decisive over ð¥, ðŠ â ð. Let ð, ð â ð be two other states. We have to show that ð is decisive over ð and ð. Without loss of generality, assume for all individuals ð â¿ð ð¥ and ðŠ â¿ð ð. Then, the Pareto order implies that ð â» ð¥ and ðŠ â» ð. Assume that for every ð â ð, ð¥ â¿ð ðŠ. Since ð is decisive over ð¥ and ðŠ, the social order ranks ð¥ â¿ ðŠ. By transitivity, ð â¿ ð. By IIA, this holds irrespective of individual preferences on other alternatives. Hence, ð is decisive over ð and ð. 1.61 Assume that ð is decisive. Let ð¥, ðŠ and ð§ be any three alternatives and assume ð¥ â¿ ðŠ for every ð â ð. Partition ð into two subgroups ð1 and ð2 so that ð¥ â¿ð ð§ for every ð â ð1 and ð§ â¿ð ðŠ for every ð â ð2 Since ð is decisive, ð¥ â¿ ðŠ. By completeness, either ð¥ â¿ ð§ in which case ð1 is decisive over ð¥ and ð§. By the ï¬eld expansion lemma (Exercise 1.60), ð1 is decisive. 9
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ð§ â» ð¥ which implies ð§ â¿ ðŠ. In this case, ð2 is decisive over ðŠ and ð§, and therefore (Exercise 1.60) decisive. 1.62 Assume â» is a social order which is Pareto and satisï¬es Independence of Irrelevant Alternatives. By the Pareto principle, the whole group is decisive over any pair of alternatives. By the previous exercise, some proper subgroup is decisive. Continuing in this way, we eventually arrive at a decisive subgroup of one individual. By the Field Expansion Lemma (Exercise 1.60), that individual is decisive over every pair of alternatives. That is, the individual is a dictator. 1.63 Assume ðŽ is decisive over ð¥ and ðŠ and ðµ is decisive over ð€ and ð§. That is, assume ð¥ â»ðŽ ðŠ =â ð¥ â» ðŠ ð€ â»ðµ ð§ =â ð€ â» ð§ Also assume ðŠ â¿ð ð€
for every ð
ð§ â¿ð ð¥
for every ð
This implies that ðŠ â¿ ð€ and ð§ â¿ ð¥ (Pareto principle). Combining these preferences, transitivity implies that ð¥â»ðŠâ¿ð€â»ð§ which contradicts the assumption that ð§ â¿ ð¥. Therefore, the implied social ordering is intransitive. 1.64 Assume ð¥ â core. In particular this implies that there does not exist any ðŠ â ð (ð ) such that ðŠ â» ð¥. Therefore ð¥ â Pareto. 1.65 No state will accept a cost share which exceeds what it can achieve on its own, so that if ð¥ â core then ð¥ðŽð †1870 ð¥ð ð †5330 ð¥ðŽð †860 Similarly, the combined share of the two states AP and TN should not exceed 6990, which they could achieve by proceeding without KM, that is ð¥ðŽð + ð¥ð ð †6990 Similarly ð¥ðŽð + ð¥ðŸð †1960 ð¥ð ð + ð¥ðŸð †5020 Finally, the sum of the shares should equal the total cost ð¥ðŽð + ð¥ð ð + ð¥ðŸð = 6530 The core is the set of all allocations of the total cost which satisfy the preceding inequalities. For example, the allocation (ð¥ðŽð = 1500, ð¥ð ð = 5000, ð¥ðŸð = 30) does not belong to the core, since TN and KM will object to their combined share of 5030; since they can meet their needs jointly at a total cost of 5020. One the other hand, no group can object to the allocation (ð¥ðŽð = 1510, ð¥ð ð = 5000, ð¥ðŸð = 20), which therefore belongs to the core. 10
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
1.66 The usual way to model a cost allocation problem as a TP-coalitional game is to regard the potential cost savings from cooperation as the sum to be allocated. In this example, the total joint cost of 6530 represents a potential saving of 1530 over the aggregate cost of 8060 if each region goes its own way. This potential saving of 1530 measures ð€(ð ). Similarly, undertaking a joint development, AP and TN could satisfy their combined requirements at a total cost of 6890. This compares with the standalone costs of 7100 (= 1870 (AP) + 5330 (TN)). Hence, the potential cost savings from their collaboration are 210 (= 7100 - 6890), which measures ð€(ðŽð, ð ð ). By similar calculations, we can compute the worth of each coalition, namely ð€(ðŽð ) = 0 ð€(ð ð ) = 0
ð€(ðŽð, ð ð ) = 210 ð€(ðŽð, ðŸð ) = 770
ð€(ðŸð ) = 0
ð€(ðŸð, ð ð ) = 1170
ð€(ð ) = 1530
An outcome in this game is an allocation of the total cost savings ð€(ð ) = 1530 amongst the three players. This can be translated into ï¬nal cost shares by subtracting each players share of the cost savings from their standalone cost. For example, a speciï¬c outcome in this game is (ð¥ðŽð = 370, ð¥ð ð = 330, ð¥ðŸð = 830), which corresponds to ï¬nal cost shares of 1500 for AP, 5000 for TN and 30 for KM. 1.67 Let ð¶ = {x â ð :
â
ð¥ð ⥠ð€(ð) for every ð â ð }
ðâð
/ core. This implies there exists some 1. ð¶ â core Assume that x â ð¶. Suppose x â coalition ð and outcome y â ð€(ð) such that y â»ð x for every ð â ð. â â y â ð€(ð) implies ðâð ðŠð †ð€(ð) while â y â»ð x for every ð â ð implies ðŠð > ð¥ð for every ð â ð. Summing, this implies â â ðŠð > ð¥ð ⥠ð€(ð) ðâð
ðâð
This contradiction establishes that x â core. 2. core â ð¶ Assume that x â core. Suppose x â / ð¶. This implies there exists some â â coalition ð such that ðâð ð¥ð < ð€(ð). Let ð = ð€(ð) â ðâð ð¥ð and consider the allocation y obtained by reallocating ð from ð ð to ð, that is { ð¥ð + ð/ð ðâð ðŠð = ð¥ð â ð/(ð â ð ) ð â /ð where ð = â£ð⣠is the number of players in ð and ð = â£ð ⣠is the number in ð . Then â ðŠð > ð¥ð forâ every ð â ð so that y â»ð x for every ð â ð. Further, y â ð€(ð) since ðâð ðŠð = ðâð ð¥ð + ð = ð€(ð) and y â ð since â ðâð
ðŠð =
â ðâð
(ð¥ð + ð/ð ) +
â
(ð¥ð â ð/(ð â ð )) =
ðâð /
â
ð¥ð = ð€(ð )
ðâð
This contradicts our assumption that x â / core, establishing that x â ð¶.
11
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.68 The 7 unanimity games for the player set ð = {1, 2, 3} are { 1 S = {1}, {1,2}, {1,3}, N ð¢{1} (ð) = 0 otherwise { 1 S = {2}, {1,2}, {2,3}, N ð¢{2} (ð) = 0 otherwise { 1 S = {3}, {1,3}, {2,3}, N ð¢{3} (ð) = 0 otherwise { 1 S = {1,2}, N ð¢{1,2} (ð) = 0 otherwise { 1 S = {1,3}, N ð¢{1,3} (ð) = 0 otherwise { 1 S = {2,3}, N ð¢{2,3} (ð) = 0 otherwise { 1 S=N ð¢ð (ð) = 0 otherwise 1.69 Firstly, consider a simple game which is a unanimity game with essential coalition ð and let ð¥ be an outcome in which ð¥ð ⥠0
for every ð â ð
ð¥ð = 0
for every ð â /ð
and â
ð¥ð = 1
ðâð
We claim that ð¥ â core. Winning coalitions If ð is winning coalition, then ð€(ð) = 1. Furthermore, if it is a winning coalition, it must contain ð , that is ð â ð and â â ð¥ð ⥠ð¥ð = 1 = ð€(ð) ðâð
ðâð
Losing coalitions If ð is a losing coalition, ð€(ð) = 0 and â ð¥ð ⥠0 = ð€(ð) ðâð
Therefore ð¥ â core and so core â= â
. Conversely, consider a simple game which is not a unanimity game. Suppose there exists an outcome ð¥ â core. Then â ð¥ð ð€(ð ) = 1 (1.15) ðâð
12
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Since there are no veto players (ð = â
), ð€(ð â {ð}) = 1 for every player ð â ð and â ð¥ð ⥠ð€(ð â {ð}) = 1 ðâ=ð
which implies that ð¥ð = 0 for every ð â ð contradicting (1.15). Thus we conclude that core = â
. 1.70 The excesses of the proper coalitions at x1 and x2 are x1 -180 -955 -395 -365 -365 -180
{AP} {KM} {TN} {AP, KM} {AP, TN} {KM, TN}
x2 -200 -950 -380 -380 -370 -160
Therefore ð(x1 ) = (â180, â180, â365, â365, â395, â955) and ð(x2 ) = (â160, â200, â370, â380, â380, â950) d(x1 ) âºð¿ d(x2 ) which implies x1 â»ð x2 . 1.71 It is a weak order on ð, that is â¿ is reï¬exive, transitive and complete. Reï¬exivity ð and transitivity ï¬ow from the corresponding properties of â¿ð¿ on â2 . Similarly, for ð any x, y â ð, either d(x) âŸð¿ d(y) or d(y) âŸð¿ d(x) since â¿ð¿ is complete on â2 . Consequently either x â¿ y or y â¿ x (or both). â¿ is not a partial order since it is not antisymmetric d(x) âŸð¿ d(y) and d(y) âŸð¿ d(x) does not imply x = y 1.72 ð(ð, x) = ð€(ð) â
â
ð¥ð
ðâð
so that ð(ð, x) †0 ââ
â
ð¥ð ⥠ð€(ð)
ðâð
1.73 Assume to the contrary that x â Nu but that x â / core. Then, there exists a coalition ð with a positive deï¬cit ð(ð, x) > 0. Since core â= â
, there exists some y â ð such that ð(ð, y) †0 for every ð â Nu. Consequently, d(y) ⺠d(x) and y â» x, so that x â / Nu. This contradiction establishes that Nu â core. 1.74 For player 1, ðŽ1 = {ð¶, ð } and (ð¶, ð¶) â¿1 (ð¶, ð¶) (ð¶, ð¶) â¿1 (ð, ð¶) Similarly for player 2 (ð¶, ð¶) â¿2 (ð¶, ð¶) (ð¶, ð¶) â¿2 (ð¶, ð ) Therefore, (ð¶, ð¶) satisï¬es the requirements of the deï¬nition of a Nash equilibrium (Example 1.51). 13
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.75 If aâð is the best element in (ðŽð , â¿â²ð ) for every player ð, then (ðâð , aâð ) â»ð (ðð , aâð ) for every ðð â ðŽð and aâð â ðŽâð for every ð â ð . Therefore, aâ is a Nash equilibrium. ¯ is another Nash equilibrium. Then for every To show that it is unique, assume that a player ð â ð ¯âð ) â¿ð (ðð , a ¯âð ) for every ðð â ðŽð (¯ ðð , a ¯ is a maximal element of â¿â²ð . To see this, assume not. That is, which implies that a assume that there exists some ðËð â ðŽð such that ðËð â»â²ð ð ¯ð which implies ðð , aâð ) for every aâð â ðŽâð (Ë ðð , aâð ) â»ð (¯ In particular ¯âð ) â»ð (ðâð , aââð ) (Ë ðð , a ¯ is anwhich contradicts the assumption that aâ is a Nash equilibrium. Therefore, a other Nash equilibrium, then ð ¯ð is maximal in â¿â²ð and hence also a best element of â¿â²ð (Exercise 1.54), which contradicts the assumption that ðâð is the unique best element. Consequently, we conclude that aâ is the unique Nash equilibrium of the game. 1.76 We show that ð(ð¥, ðŠ) = â£ð¥ â ðŠâ£ satisï¬es the requirements of a metric, namely 1. â£ð¥ â ðŠâ£ ⥠0. 2. â£ð¥ â ðŠâ£ = 0 if and only if ð¥ = ðŠ. 3. â£ð¥ â ðŠâ£ = â£ðŠ â ð¥â£. To establish the triangle inequality, we can consider various cases. For example, if ð¥â€ðŠâ€ð§ â£ð¥ â ð§â£ + â£ð§ â ðŠâ£ ⥠â£ð¥ â ð§â£ = ð§ â ð¥ ⥠ðŠ â ð¥ = â£ð¥ â ðŠâ£ If 𥠆𧠆ðŠ â£ð¥ â ð§â£ + â£ð§ â ðŠâ£ = ð§ â ð¥ + ðŠ â ð§ = ðŠ â ð¥ = â£ð¥ â ðŠâ£ and so on. 1.77 We show that ðâ ð¥, ðŠ = maxðð=1 â£ð¥ð â ðŠð ⣠satisï¬es the requirements of a metric, namely 1. maxðð=1 â£ð¥ð â ðŠð ⣠⥠0 2. maxðð=1 â£ð¥ð â ðŠð ⣠= 0 if and only if ð¥ð = ðŠð for all ð. 3. maxðð=1 â£ð¥ð â ðŠð ⣠= maxðð=1 â£ðŠð â ð¥ð ⣠4. For every ð, â£ð¥ð â ðŠð ⣠†â£ð¥ð â ð§ð ⣠+ â£ð§ð â ðŠð ⣠from previous exercise. Therefore max â£ð¥ð â ðŠð ⣠†max (â£ð¥ð â ð§ð ⣠+ â£ð§ð â ðŠð â£) †max â£ð¥ð â ð§ð ⣠+ max â£ð§ð â ðŠð ⣠1.78 For any ð, any neighborhood of 1/ð contains points of ð (namely 1/ð) and points not in ð (1/ð + ð). Hence every point in ð is a boundary point. Also, 0 is a boundary point. Therefore b(ð) = ð ⪠{0}. Note that ð â b(ð). Therefore, ð has no interior points. 14
Solutions for Foundations of Mathematical Economics 1.79
c 2001 Michael Carter â All rights reserved
1. Let ð¥ â int ð. Thus ð is a neighborhood of ð¥. Therefore, ð â ð is a neighborhood of ð¥, so that ð¥ is an interior point of ð .
2. Clearly, if ð¥ â ð, then ð¥ â ð â ð . Therefore, assume ð¥ â ð â ð which implies that ð¥ is a boundary point of ð. Every neighborhood of ð¥ contains other points of ð â ð . Hence ð¥ â ð . 1.80 Assume that ð is open. Every ð¥ â ð has a neighborhood which is disjoint from ð ð . Hence no ð¥ â ð is a closure point of ð ð . ð ð contains all its closure points and is therefore closed. Conversely, assume that ð is closed. Let ð¥ be a point its complement ð ð . Since ð is closed and ð¥ â / ð, ð¥ is not a boundary point of ð. This implies that ð¥ has a neighborhood ð which is disjoint from ð, that is ð â ð ð . Hence, ð¥ is an interior point of ð ð . This implies that ð ð contains only interior points, and hence is open. 1.81 Clearly ð¥ is a neighborhood of every point ð¥ â ð, since ðµð (ð¥) â ð for every ð > 0. Hence, every point ð¥ â ð is an interior point of ð¥. Similarly, every point ð¥ â â
is an interior point (there are none). Since ð¥ and â
are open, there complements â
and ð¥ are closed. Alternatively, â
has no boundary points, and is therefore is open. Trivialy, on the other hand, â
contains all its boundary points, and is therefore closed. 1.82 Let ð be a metric space. Assume ð is the union of two disjoint closed sets ðŽ and ðµ, that is ð =ðŽâªðµ
ðŽâ©ðµ =â
Then ðŽ = ðµ ð is open as is ðµ = ðŽð . Therefore ð is not connected. Conversely, assume that ð is not connected. Then there exist disjoint open sets ðŽ and ðµ such that ð = ðŽ ⪠ðµ. But ðŽ = ðµ ð is also closed as is ðµ = ðŽð . Therefore ð is the union of two disjoint closed sets. 1.83 Assume ð is both open and closed, â
â ð â ð. We show that we can represent ð as the union of two disjoint open sets, ð and ð ð . For any ð â ð, ð = ð ⪠ð ð and ð â© ð ð = â
. ð is open by assumption. It complement ð ð is open since ð is closed. Therefore, ð is not connected. Conversely, assume that ð is not connected. That is, there exists two disjoint open sets ð and ð such that ð = ð ⪠ð . Now ð = ð ð , which implies that ð is closed since ð is open. Therefore ð is both open and closed. 1.84 Assume that ð is both open and closed. Then so is ð ð and ð is the disjoint union of two closed sets ð¥ = ð ⪠ðð so that b(ð) = ð â© ð ð = ð â© ð ð = â
Conversely, assume that b(ð) = ð â© ð ð = â
. This implies that Consider any ð¥ â ð. Since ð â© ð ð = â
, ð¥ â / ð ð . A fortiori, x â / ð ð which implies that ð¥ â ð and therefore ð â ð. ð is closed. Similarly we can show that ð ð â ð ð so that ð ð is closed and therefore ð is open. ð is both open and closed.
15
Solutions for Foundations of Mathematical Economics 1.85
c 2001 Michael Carter â All rights reserved
1. Let {ðºð } be a (possibly inï¬nite) collection of open sets. Let ðº = âªð ðºð . Let ð¥ be a point in ðº. Then there exists some particular ðºð which contains ð¥. Since ðºð is open, ðºð is a neighborhood of ð¥. Since ðºð â ðº, ð¥ is an interior point of ðº. Since ð¥ is an arbitrary point in ðº, we have shown that every ð¥ â ðº is an interior point. Hence, ðº is open. What happens if every ðºð is empty? In this case, ðº = â
and is open (Exercise 1.81). The other possibility is that the collection {ðºð } is empty. Again ðº = â
which is open. Suppose { ðº1 , ðº2 , . . . , ðºð } is a ï¬nite collection of open sets. Let ðº = â©ð ðºð . If ðº = â
, then it is trivially open. Otherwise, let ð¥ be a point in ðº. Then ð¥ â ðºð for all ð = 1, 2, . . . , ð. Since the sets ðºð are open, for every ð, there exists an open ball ðµ(ð¥, ðð ) â ðºð about ð¥. Let ð be the smallest radius of these open balls, that is ð = min{ ð1 , ð2 , . . . , ðð }. Then ðµð (ð¥) â ðµ(ð¥, ðð ), so that ðµð (ð¥) â ðºð for all i. Hence ðµð (ð¥) â ðº. ð¥ is an interior point of ðº and ðº is open. To complete the proof, we need to deal with the trivial case in which the collection is empty. In that case, ðº = â©ð ðºð = ð and hence is open.
2. The corresponding properties of closed sets are established analogously. 1.86
1. Let ð¥0 be an interior point of ð. This implies there exists an open ball ðµ â ð about ð¥0 . Every ð¥ â ðµ is an interior point of ð. Hence ðµ â int ð. ð¥0 is an interior point of int ð which is therefore open. Let ðº be any open subset of ð and ð¥ be a point in ðº. ðº is neighborhood of ð¥, which implies that ð â ðº is also neighborhood of ð¥. Therefore ð¥ is an interior point of ð. Therefore int ð contains every open subset ðº â ð, and hence is the largest open set in ð.
2. Let ð denote the closure of the set ð. Clearly, ð â ð. To show the converse, let ð¥ be a closure point of ð and let ð be a neighborhood of ð¥. Then ð contains some other point ð¥â² â= ð which is a closure point of ð. ð is a neighborhood of ð¥â² which intersects ð. Hence ð¥ is a closure point of ð. Consequently ð = ð which implies that ð is closed. Assume ð¹ is a closed subset of containing ð. Then ðâð¹ =ð¹ since ð¹ is closed. Hence, ð is a subset of every closed set containing ð. 1.87 Every ð¥ â ð is either an interior point or a boundary point. Consequently, the interior of ð is the set of all ð¥ â ð which are not boundary points int ð = ð â b(ð) 1.88 Assume that ð is closed, that is ð = ð ⪠b(ð) = ð This implies that b(ð) â ð. ð contains its boundary. Assume that ð contains its boundary, that is ð â b(ð). Then ð = ð ⪠b(ð) = ð ð is closed. 16
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.89 Assume ð is bounded, and let ð = ð(ð). Choose any ð¥ â ð. For all ðŠ â ð, ð(ð¥, ðŠ) †ð < ð + 1. Therefore, ðŠ â ðµ(ð¥, ð + 1). ð is contained in the open ball ðµ(ð¥, ð + 1). Conversely, assume ð is contained in the open ball ðµð (ð¥). Then for any ðŠ, ð§ â ð ð(ðŠ, ð§) †ð(ðŠ, ð¥) + ð(ð¥, ð§) < 2ð by the triangle inequality. Therefore ð(ð) < 2ð and the set is bounded. 1.90 Let ðŠ â ð â© ðµð (ð¥0 ). For every ð¥ â ð, ð(ð¥, ðŠ) < ð and therefore ð(ð¥, ð¥0 ) †ð(ð¥, ðŠ) + ð(ðŠ, ð¥0 ) < ð + ð = 2ð so that ð¥ â ðµ2ð (ð¥0 ). 1.91 Let y0 â ð . For any ð > 0, let yâ² = y â ð be the production plan which is ð units less in every commodity. Then, for any y â ðµð (yâ² ) ðŠð â ðŠðⲠ†ðâ (y, yâ² ) < ð
for every ð
and therefore y < y0 . Thus ðµð (yâ² ) â ð and so yâ² â int ð â= â
. 1.92 For any ð¥ â ð1 ðð¥ = ð(ð¥, ð2 ) > 0 Similarly, for every ðŠ â ð2 ððŠ = ð(ðŠ, ð1 ) > 0 Let ð1 =
âª
ðµðð¥ /2 (ð¥)
ð¥âð1
ð2 =
âª
ðµððŠ /2 (ð¥)
ðŠâð2
Then ð1 and ð2 are open sets containing ð1 and ð2 respectively. To show that ð1 and ð2 are disjoint, suppose to the contrary that ð§ â ð1 â© ð2 . Then, there exist points ð¥ â ð1 and ðŠ â ð2 such that ð(ð¥, ð§) < ðð¥ /2,
ð(ðŠ, ð§) < ððŠ /2
Without loss of generality, suppose that ð𥠆ððŠ and therefore ð(ð¥, ðŠ) †ð(ð¥, ð§) + ð(ðŠ, ð§) < ðð¥ /2 + ððŠ /2 †ððŠ which contradicts the deï¬nition of ððŠ and shows that ð1 â© ð2 = â
. 1.93 By Exercise 1.92, there exist disjoint open sets ð1 and ð2 such that ð1 â ð1 and ð2 â ð2 . Since ð2 â ð2 , ð2 â© ð2ð = â
. ð2ð is a closed set which contains ð1 , and therefore ð2 â© ð1 = â
. ð = ð1 is the desired set. 1.94 See Figure 1.2.
17
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
ð
1
ðµ1/2 ((2, 0)) 1
2
Figure 1.2: Open ball about (2, 0) relative to ð 1.95 Assume ð is connected. Suppose ð is not an interval. This implies that there exists numbers ð¥, ðŠ, ð§ such that ð¥ < ðŠ < ð§ and ð¥, ð§ â ð while ðŠ â / ð. Then ð = (ð â© (ââ, ðŠ)) ⪠(ð â© (ðŠ, â)) represents ð as the union of two disjoint open sets (relative to ð), contradicting the assumption that ð is connected. Conversely, assume that ð is an interval. Suppose that ð is not connected. That is, ð = ðŽ ⪠ðµ where ðŽ and ðµ are nonempty disjoint closed sets. Choose ð¥ â ðŽ and ð§ â ðµ. Since ðŽ and ðµ are disjoint, ð¥ â= ð§. Without loss of generality, we may assume ð¥ < ð§. Since ð is an interval, [ð¥, ð§] â ð = ðŽ ⪠ðµ. Let ðŠ = sup{ [ð¥, ð§] â© ð } Clearly ð¥ †ðŠ †ð§ so that ðŠ â ð. Now ðŠ belongs to either ðŽ or ðµ. Since ðŽ is closed in ð, [ð¥, ð§] â© ðŽ is closed and ðŠ = sup{ [ð¥, ð§] â© ð } â ðŽ. This implies the ðŠ < ð§. Consequently, ðŠ + ð â ðµ for every ð > 0 such that ðŠ + ð †ð§. Since ðµ is closed, ðŠ â ðµ. This implies that ðŠ belongs to both ðŽ and ðµ contradicting the assumption that ðŽ â© ðµ = â
. We conclude that ð must be connected. 1.96 Assume ð¥ð â ð¥ and also ð¥ð â ðŠ. We have to show that ð¥ = ðŠ. Suppose not, that is suppose ð¥ â= ðŠ (see Figure 1.3). Then ð(ð¥, ðŠ) = ð
> 0. Let ð = ð
/3 > 0. Since ð¥ð â ð¥, there exists some ðð¥ such that ð¥ð â ðµð (ð¥) for all ð ⥠ðð¥ . Since ð¥ð â ðŠ, there exists some ððŠ such that ð¥ð â ðµð (ðŠ) for all ð ⥠ððŠ . But these statements are contradictory since ðµð (ð¥) â© ðµ(ðŠ, ð) = â
. We conclude that the successive terms of a convergent sequence cannot get arbitrarily close to two distinct points, so that the limit a convergent sequence is unique. 1.97 Let (ð¥ð ) be a sequence which converges to ð¥. There exists some ð such that ð(ð¥ð â ð¥) < 1 for all ð ⥠ð . Let ð
= max{ ð(ð¥1 â ð¥), ð(ð¥2 â ð¥), . . . , ð(ð¥ð â1 â ð¥), 1 } Then for all ð, ð(ð¥ð âð¥) †ð
. That is every element ð¥ð in the sequence (ð¥ð ) belongs to ðµ(ð¥, ð
+ 1), the open ball about ð¥ of radius ð
+ 1. Therefore the sequence is bounded. 18
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics ðµ(ð¥, ð)
ð
ðµ(ðŠ, ð)
ð
ð
ð¥
ðŠ
Figure 1.3: A convergent sequence cannot have two distinct limits 1.98 The share ð ð of the ðth guest is ð ð =
1ð 2
lim ð ð = 0 However, ð ð > 0 for all ð. There is no limit to the number of guests who will get a share of the cake, although the shares will get vanishingly small for large parties. 1.99 Suppose ð¥ð â ð¥. That is, there exists some ð such that ð(ð¥ð , ð¥) < ð/2 for all ð ⥠ð . Then, for all ð, ð ⥠ð ð(ð¥ð , ð¥ð ) †ð(ð¥ð , ð¥) + ð(ð¥, ð¥ð ) < ð/2 + ð/2 = ð 1.100 Let (ð¥ð ) be a Cauchy sequence. There exists some ð such that ð(ð¥ð â ð¥ð ) < 1 for all ð ⥠ð . Let ð
= max{ ð(ð¥1 â ð¥ð ), ð(ð¥2 â ð¥ð ), . . . , ð(ð¥ð â1 â ð¥ð ), 1 } Every ð¥ð belongs to ðµ(ð¥ð , ð
+ 1), the ball of radius ð
+ 1 centered on ð¥ð . 1.101 Let (ð¥ð ) be a bounded increasing sequence in â and let ð = { ð¥ð } be the set of elements of (ð¥ð ). Let ð be the least upper bound of ð. We show that ð¥ð â ð. First observe that ð¥ð †ð for every ð (since ð is an upper bound). Since ð is the least upper bound, for every ð > 0 there exists some element ð¥ð such that ð¥ð > ð â ð. Since (ð¥ð ) is increasing, we must have ð â ð < ð¥ð †ð for every ð ⥠ð That is, for every ð > 0 there exists an ð such that ð(ð¥ð , ð¥) < ð for every ð ⥠ð ð¥ð â ð. 1.102 If ðœ > 1, the sequence ðœ, ðœ 2 , ðœ 3 , . . . is unbounded. Otherwise, if 𜠆1, ðœ ð †ðœ ðâ1 and the sequence is decreasing and bounded by 𜠆1. Therefore the sequence converges (Exercise 1.101). Let ð¥ = limðââ . Then ðœ ð+1 = ðœðœ ð and therefore ð¥ = lim ðœ ð+1 = ðœ lim ðœ ð = ðœð¥ ðââ
ðââ
which can be satisï¬ed if and only if 19
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics â ðœ = 1, in which case ð¥ = lim 1ð = 1 â ð¥ = 0 when 0 †ðœ < 1 Therefore ðœ ð â 0 ââ ðœ < 1 1.103
1. For every ð¥ â â (ð¥ â
Expanding
â 2 2) ⥠0
â ð¥2 â 2 2ð¥ + 2 ⥠0
â ð¥2 + 2 ⥠2 2ð¥
Dividing by ð¥ â 2 â¥2 2 ð¥
ð¥+ for every ð¥ > 0. Therefore 1 2
(
2 ð¥+ ð¥
) â¥
â 2
2. Let (ð¥ð ) be the sequence deï¬ned in Example 1.64. That is ) ( 1 2 ð¥ð = ð¥ðâ1 + ðâ1 2 ð¥ Starting from ð¥0 = 2, it is clear that ð¥ð ⥠0 for all ð. Substituting in ( ) â 1 2 ð¥+ ⥠2 2 ð¥ 1 ð¥ = 2 ð
That is ð¥ð â¥
( ð¥ðâ1 +
2 ð¥ðâ1
) â¥
â 2
â 2 for every ð. Therefore for every ð ) ( 1 2 ð¥ð â ð¥ð+1 = ð¥ð â ð¥ð + ð 2 ð¥ ) ( 1 2 = ð¥ð â ð 2 ð¥ ( ) 1 2 ⥠ð¥ð â â 2 2 â ð =ð¥ â 2 â¥0
â This implies that ð¥ð+1 †ð¥ð . Consequently 2 †ð¥ð †2 for every ð. (ð¥ð ) is a bounded monotone sequence. By Exercise 1.101, ð¥ð â ð¥. The limit ð¥ satisï¬es the equation ( ) 1 2 ð¥= ð¥+ 2 ð¥ â Solving, this implies ð¥2 = 2 or ð¥ = 2 as required. 20
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.104 The following sequence approximates the square root of any positive number ð ð¥1 = ð 1( ð ) ð¥ð+1 = ð¥ð + ð 2 ð¥ 1.105 Let ð¥ â ð. If ð¥ â ð, then ð¥ is the limit of the sequence (ð¥, ð¥, ð¥, . . . ). If ð¥ â / ð, then ð¥ is a boundary point of ð. For every ð, the ball ðµ(ð¥, 1/ð) contains a point ð¥ð â ð. From the sequence of open balls ðµ(ð¥, 1/ð) for ð = 1, 2, 3, . . . , we can generate of a sequence of points ð¥ð which converges to ð¥. Conversely, assume that ð¥ is the limit of a sequence (ð¥ð ) of points in ð. Either ð¥ â ð and therefore ð¥ â ð. Or ð¥ â / ð. Since ð¥ð â ð¥, every neighborhood of ð¥ contains points ð ð¥ of the sequence. Hence, ð¥ is a boundary point of ð and ð¥ â ð. 1.106 ð is closed if and only if ð = ð. The result follows from Exercise 1.105. 1.107 Let ð be a closed subset of a complete metric space ð. Let (ð¥ð ) be a Cauchy sequence in ð. Since ð is complete, ð¥ð â ð¥ â ð. Since ð is closed, ð¥ â ð (Exercise 1.106). 1.108 Since ð(ð ð ) â 0, ð cannot contain more than one point. Therefore, it suï¬ces to show that ð is nonempty. Choose some ð¥ð from each ð ð . Since ð(ð ð ) â 0, (ð¥ð ) is a Cauchy sequence. Since ð is complete, there exists some ð¥ â ð such that ð¥ð â ð¥. Choose some ð. Since the sets are nested, the subsequence { ð¥ð : ð ⥠ð } â ð ð . Since ð ð is closed, ð¥ â ð ð (Exercise 1.106). Since ð¥ â ð ð for every ð ð¥â
â â©
ðð
ð=1
1.109 If player 1 picks closed balls whose radius decreases by at least half after each pair of moves, then { ð 1 , ð 3 , ð 5 , . . . } is a nested sequence of closed sets which has a nonempty intersection (Exercise 1.108). 1.110 Let (ð¥ð ) be a sequence in ð â ð with ð closed and ð compact. Since ð is compact, there exists a convergent subsequence ð¥ð â ð¥ â ð . Since ð is closed, we must have ð¥ â ð (Exercise 1.106). Therefore (ð¥ð ) contains a subsequence which converges in ð, so that ð is compact. 1.111 Let (ð¥ð ) be a Cauchy sequence in a metric space. For every ð > 0, there exists ð such that ð(ð¥ð , ð¥ð ) < ð/2 for all ð, ð ⥠ð Trivially, if (ð¥ð ) converges, it has a convergent subsequence (the whole sequence). Conversely, assume that (ð¥ð ) has a subsequence (ð¥ð ) which converges to ð¥. That is, there exists some ð such that ð(ð¥ð , ð¥) < ð/2 for all ð ⥠ð Therefore, by the triangle inequality ð(ð¥ð , ð¥) †ð(ð¥ð , ð¥ð ) + ð(ð¥ð , ð¥) < ð/2 + ð/2 = ð for all ð ⥠max ð, ð
21
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.112 We proceed sequentially as follows. Choose any ð¥1 in ð. If the open ball ðµ(ð¥1 , ð) contains ð, we are done. Otherwise, choose some ð¥2 â / ðµ(ð¥1 , ð) and consider the set âª2 ðµ(ð¥ð , ð). If this set contains ð, we are done. Otherwise, choose some ð¥3 â / âªð=1 âª3 2 ðµ(ð¥ , ð) and consider ðµ(ð¥ , ð) ð ð ð=1 ð=1 The process must terminate with a ï¬nite number of open balls. Otherwise, if the process could be continued indeï¬nitely, we could construct an inï¬nite sequence (ð¥1 , ð¥2 , ð¥3 , . . . ) which had no convergent subsequence. The would contradict the compactness of ð. 1.113 Assume ð is compact. The previous exercise showed that ð is totally bounded. Further, since every sequence has a convergent subsequence, every Cauchy sequence converges (Exercise 1.111). Therefore ð is complete. Conversely, assume that ð is complete and totally bounded and let ð1 = { ð¥11 , ð¥21 , ð¥31 , . . . } be an inï¬nite sequence of points in ð. Since ð is totally bounded, it is covered by a ï¬nite collection of open balls of radius 1/2. ð1 has a subsequence ð2 = { ð¥12 , ð¥22 , ð¥32 , . . . } all of whose points lie in one of the open balls. Similarly, ð2 has a subsequence ð3 = { ð¥13 , ð¥23 , ð¥33 , . . . } all of whose points lie in an open ball of radius 1/3. Continuing in this fashion, we construct a sequence of subsequences, each of which lies in a ball of smaller and smaller radius. Consequently, successive terms of the âdiagonalâ subsequence { ð¥11 , ð¥22 , ð¥33 , . . . } get closer and closer together. That is, ð is a Cauchy sequence. Since ð is complete, ð converges in ð and ð1 has a convergent subsequence ð. Hence, ð is compact. 1.114
1. Every big set ð â ⬠has a least two distinct points. 0 for every ð â â¬.
Hence ð(ð ) >
2. Otherwise, there exists ð such that ð(ð ) ⥠1/ð for every ð â ⬠and therefore ð¿ = inf ð â⬠ð(ð ) ⥠1/ð > 0. 3. Choose a point ð¥ð in each ðð . Since ð is compact, the sequence (ð¥ð ) has a convergent subsequence (ð¥ð ) which converges to some point ð¥0 â ð. 4. The point ð¥0 belongs to at least one ð0 in the open cover ð. Since ð0 is open, there exists some open ball ðµð (ð¥0 ) â ð0 . 5. Consider the concentric ball ðµð/2 (ð¥0 ). Since (ð¥ð ) is a convergent subsequence, there exists some ð such that ð¥ð â ðµð/2 (ð¥) for every ð ⥠ð . 6. Choose some ð0 ⥠min{ ð, 2/ð }. Then 1/ð0 < ð/2 and ð(ðð0 ) < 1/ð0 < ð/2. ð¥ð0 â ðð0 â© ðµð/2 (ð¥) and therefore (Exercise 1.90) ðð0 â ðµð (ð¥) â ð 0 . This contradicts the assumption that ðð is a big set. Therefore, we conclude that ð¿ > 0. 1.115
1. ð is totally bounded (Exercise 1.112). Therefore, for every ð > 0, there exists a ï¬nite number of open balls ðµð (ð¥ð ) such that ð=
ð âª
ðµð (ð¥ð )
ð=1
2. ð(ðµð (ð¥ð )) = 2ð < ð¿. By deï¬nition of the Lebesgue number, every ðµð (ð¥ð ) is contained in some ðð â ð. 3. The collection of open balls {ðµð (ð¥ð )} covers ð. Therefore, fore every ð¥ â ð, there exists ð such that ð¥ â ðµð (ð¥ð ) â ðð 22
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Therefore, the ï¬nite collection ð1 , ð2 , . . . , ðð covers ð. 1.116 For any family of subsets ð ⩠⪠ð = â
ââ ðð = ð ðâð
ðâð
Suppose to the contrary that ð is a collection of closed sets with the ï¬nite intersection â© property, but that ðâð ð = â
. Then { ð ð : ð â ð } is a open cover of ð which does not have a ï¬nite subcover. Consequently ð cannot be compact. Conversely, assume every collection of closed sets with the ï¬nite intersection property has a nonempty intersection. Let ⬠be an open cover of ð. Let ð = { ð â ð : ðð â ⬠} That is
âª
ð ð = ð which implies
ðâð
â©
ð=â
ðâð
Consequently, ð does not have the ï¬nite intersection property. There exists a ï¬nite subcollection { ð1 , ð2 , . . . , ðð } such that ð â©
ðð = â
ð=1
which implies that ð âª
ððð = ð
ð=1
{ ð1ð , ð2ð , . . . , ððð } is a ï¬nite subcover of ð. Thus, ð is compact. 1.117 Every ï¬nite collection of nested (nonempty) sets has the ï¬nite intersection property. By Exercise 1.116, the sequence has a non-empty intersection. (Note: every set ðð is a subset of the compact set ð1 .) 1.118 (1) =â (2) Exercises 1.114 and 1.115. (2) =â (3) Exercise 1.116 (3) =â (1) Let ð be a metric space in which every collection of closed subsets with the ï¬nite intersection property has a ï¬nite intersection. Let (ð¥ð ) be a sequence in ð. For any ð, let ðð be the tail of the sequence minus the ï¬rst ð terms, that is ðð = { ð¥ð : ð = ð + 1, ð + 2, . . . } The collection (ðð ) has the ï¬nite intersection property since, for any ï¬nite set of integers { ð1 , ð2 , . . . , ðð } ð â©
ððð â ððŸ â= â
ð=1
where ðŸ = max{ ð1 , ð2 , . . . , ðð }. Therefore â â© ð=1
23
ðð â= â
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
âªâ Choose any ð¥ â ð=1 ðð . That is, ð¥ â ðð for each ð = 1, 2, . . . . Thus, for every ð > 0 and ð = 1, 2, . . . , there exists some ð¥ð â ðµð (ð¥) â© ðð We construct a subsequence as follows. For ð = 1, 2, . . . , let ð¥ð be the ï¬rst term in ðð which belongs to ðµ1/ð (ð¥). Then, (ð¥ð ) is a subsequence of (ð¥ð ) which converges to ð¥. We conclude that every sequence has a convergent subsequence. 1.119 Assume (ð¥ð ) is a bounded sequence in â. Without loss of generality, we can assume that { ð¥ð } â [0, 1]. Divide ðŒ 0 = [0, 1] into two sub-intervals [0, 1/2] and [1/2, 1]. At least one of the sub-intervals must contain an inï¬nite number of terms of the sequence. Call this interval ðŒ 1 . Continuing this process of subdivision, we obtain a nested sequence of intervals ðŒ0 â ðŒ1 â ðŒ2 â . . . each of which contains an inï¬nite number of terms of the sequence. Consequently, we can construct a subsequence (ð¥ð ) with ð¥ð â ðŒ ð . Furthermore, the intervals get smaller and smaller with ð(ðŒ ð ) â 0, so that (ð¥ð ) is a Cauchy sequence. Since â is complete, the subsequence (ð¥ð ) converges to ð¥ â â. Note how we implicitly called on the Axiom of Choice (Remark 1.5) in choosing a subsequence from the nested sequence of intervals. 1.120 Let (ð¥ð ) be a Cauchy sequence in â. That is, for every ð > 0, there exists ð such that â£ð¥ð â ð¥ð ⣠< ð for all ð, ð ⥠ð . (ð¥ð ) is bounded (Exercise 1.100) and hence by the Bolzano-Weierstrass theorem, it has a convergent subsequence (ð¥ð ) with ð¥ð â ð¥ â â. Choose ð¥ð from the convergent subsequence such that ð ⥠ð and â£ð¥ð â ð¥â£ < ð/2. By the triangle inequality â£ð¥ð â ð¥â£ †â£ð¥ð â ð¥ð ⣠+ â£ð¥ð â ð¥â£ < ð/2 + ð/2 = ð Hence the sequence (ð¥ð ) converges to ð¥ â â. 1.121 Since ð1 and ð2 are linear spaces, x1 + y1 â ð1 and x2 + y2 â ð2 , so that (x1 + y1 , x2 + y2 ) â ð1 à ð2 . Similarly (ðŒx1 , ðŒx2 ) â ð1 à ð2 for every (x1 , x2 ) â ð1 à ð2 . Hence, ð = ð1 à ð2 is closed under addition and scalar multiplication. With addition and scalar multiplication deï¬ned component-wise, ð inherits the arithmetic properties (like associativity) of its constituent spaces. Verifying this would proceed identically as for âð . It is straightforward though tedious. The zero element in ð is 0 = (01 , 02 ) where 01 is the zero element in ð1 and 02 is the zero element in ð2 . Similarly, the inverse of x = (x1 , x2 ) is âx = (âx1 , âx2 ). 1.122
1. x+y =x+z âx + (x + y) = âx + (x + z) (âx + x) + y = (âx + x) + z 0+y =0+z y=z
24
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. ðŒx = ðŒy 1 1 (ðŒx) = (ðŒy) ðŒ ( ) (ðŒ ) 1 1 ðŒ x= ðŒ y ðŒ ðŒ x=y 3. ðŒx = ðœx implies (ðŒ â ðœ)x = ðŒx â ðœx = 0 Provided x = 0, we must have (ðŒ â ðœ)x = 0x That is ðŒ â ðœ = 0 which implies ðŒ = ðœ. 4. (ðŒ â ðœ)x = (ðŒ + (âðœ))x = ðŒx + (âðœ)x = ðŒx â ðœx 5. ðŒ(x â y) = ðŒ(x + (â1)y) = ðŒx + ðŒ(â1)y = ðŒx â ðŒy 6. ðŒ0 = ðŒ(x + (âx)) = ðŒx + ðŒ(âx) = ðŒx â ðŒx =0 1.123 The linear hull of the vectors {(1, 0), (0, 2)} is { ( ) ( )} 1 0 lin {(1, 0), (0, 2)} = ðŒ1 + ðŒ2 0 2 ) ( { ðŒ1 } = ðŒ2 = â2 The linear hull of the vectors {(1, 0), (0, 2)} is the whole plane â2 . Figure 1.4 illustrates how any vector in â2 can be obtained as a linear combination of {(1, 0), (0, 2)}. 1.124
1. From the deï¬nition of ðŒ, ðŒð = ð€(ð) â
â ð âð
25
ðŒð
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
(â2, 3)
3 2 1
-2
-1
0
1
Figure 1.4: Illustrating the span of { (1, 0), (0, 2) }. for every ð â ð . Rearranging ð€(ð) = ðŒð + =
â
â
ðŒð +
ð =ð
=
â
ðŒð
ð âð
â
ðŒð
ð âð
ðŒð
ð âð
2.
â
ðŒð ð€ð (ð) =
ð âð
â
ðŒð ð€ð (ð) +
ð âð
=
â
ðŒð 1 +
ð âð
=
â
â
â
ðŒð ð€ð (ð)
ð ââð
ðŒð 0
ð ââð
ðŒð 1
ð âð
= ð€(ð) 1.125
1. Choose any x â ð. By homogeneity 0x = ð â ð.
2. For every x â ð, âx = (â1)x â ð by homogeneity. 1.126 Examples of subspaces in âð include: 1. The set containing just the null vector {0} is subspace. 2. Let x be any element in âð and let ð be the set of all scalar multiples of x ð = { ðŒx : ðŒ â â } ð is a line through the origin in âð and is a subspace. 3. Let ð be the set of all ð-tuples with zero ï¬rst coordinate, that is ð = { (ð¥1 , ð¥2 , . . . , ð¥ð ) : ð¥1 = 0, ð¥ð â â, ð â= 1 } For any x, y â ð x + y = (0, ð¥2 , ð¥3 , . . . , ð¥ð ) + (0, ðŠ2 , ðŠ3 , . . . , ðŠð ) = (0, ð¥2 + ðŠ2 , ð¥3 + ðŠ3 , . . . , ð¥ð + ðŠð ) â ð 26
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Similarly ðŒx = ðŒ(0, ð¥2 , ð¥3 , . . . , ð¥ð ) = (0, ðŒð¥2 , ðŒð¥3 , . . . , ðŒð¥ð ) â ð Therefore ð is a subspace of âð . Generalizing, any set of vectors with one or more coordinates identically zero is a subspace of âð . 4. We will meet some more complicated subspaces in Chapter 2. 1.127 No, âx â / âð+ if x â âð+ unless x = 0. âð+ is an example of a cone (Section 1.4.5). 1.128 lin ð is a subspace Let x, y be two elements in lin ð. x is a linear combination of elements of ð, that is x = ðŒ1 ð¥1 + ðŒ2 ð¥2 + . . . ðŒð ð¥ð Similarly y = ðœ1 ð¥1 + ðœ2 ð¥2 + . . . ðœð ð¥ð and x + y = (ðŒ1 + ðœ1 )ð¥1 + (ðŒ2 + ðœ2 )ð¥2 + â
â
â
+ (ðŒð + ðœð )ð¥ð â lin ð and ðŒx = ðŒðŒ1 ð¥1 + ðŒðŒ2 ð¥2 + â
â
â
+ ðŒðŒð ð¥ð â lin ð This shows that lin ð is closed under addition and scalar multiplication and hence is a subspace. lin ð is the smallest subspace containing ð Let ð be any subspace containing ð. Then ð contains all linear combinations of elements in ð, so that lin ð â ð . Hence lin ð is the smallest subspace containing S. 1.129 The previous exercise showed that lin ð is a subspace. Therefore, if ð = lin ð, ð is a subspace. Conversely, assume that ð is a subspace. Then ð is the smallest subspace containing ð, and therefore ð = lin ð (again by the previous exercise). 1.130 Let x, y â ð = ð1 â© ð2 . Hence x, y â ð1 and for any ðŒ, ðœ â â, ðŒx + ðœy â ð1 . Similarly ðŒx + ðœy â ð2 and therefore ðŒx + ðœy â ð. ð is a subspace. 1.131 Let ð = ð1 + ð2 . First note that 0 = 0 + 0 â ð. Suppose x, y belong to ð. Then there exist s1 , t1 â ð1 and s2 , t2 â ð2 such that x = s1 + s2 and y = t1 + t2 . For any ðŒ, ðœ â â, ðŒx + ðœy = ðŒ(s1 + s2 ) + ðœ(t1 + t2 ) = (ðŒs1 + ðœt1 ) + (ðŒs2 + ðœt2 ) â ð since ðŒs1 + ðœt1 â ð1 and ðŒs2 + ðœt2 â ð2 . 1.132 Let ð1 = { ðŒ(1, 0) : ðŒ â â } ð2 = { ðŒ(0, 1) : ðŒ â â } 27
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ð1 and ð2 are respectively the horizontal and vertical axes in â2 . Their union is not a subspace, since for example ( ) ( ) ( ) 1 1 0 = + â / ð1 ⪠ð2 1 0 1 However, any vector in â2 can be written as the sum of an element of ð1 and an element of ð2 . Therefore, their sum is the whole space â2 , that is ð 1 + ð 2 = â2 1.133 Assume that ð is linearly dependent, that is there exists x1 , . . . , xð â ð and ðŒ2 , . . . , ðŒð â ð
such that x1 = ðŒ2 x2 + ðŒ3 x3 + . . . , ðŒð xð Rearranging, this implies 1x1 â ðŒ2 x2 â ðŒ3 x3 â . . . ðŒð xð = 0 Conversely, assume there exist x1 , x2 , . . . , xð â x and ðŒ1 , ðŒ2 , . . . , ðŒð â â such that ðŒ1 x1 + ðŒ2 x2 . . . + ðŒð xð = 0 Assume without loss of generality that ðŒ1 â= 0. Then x1 = â
ðŒ2 ðŒ3 ðŒð x2 â x3 â . . . â xð ðŒ1 ðŒ1 ðŒ1
which shows that x1 â lin ð â {x1 } 1.134 Assume {(1, 1, 1), (0, 1, 1), (0, 0, 1)} are linearly dependent. Then there exists ðŒ1 , ðŒ2 , ðŒ3 such that â â â â â â â â 1 0 0 0 ðŒ1 â1â + ðŒ2 â1â + ðŒ3 â0â = â0â 1 1 1 0 or equivalently ðŒ1 = 0 ðŒ1 + ðŒ2 = 0 ðŒ1 + ðŒ2 + ðŒ3 = 0 which imply that ðŒ1 = ðŒ2 = ðŒ3 = 0 Therefore {(1, 1, 1), (0, 1, 1), (0, 0, 1)} are linearly independent. 1.135 Suppose on the contrary that ð is linearly dependent. That is, there exists a set of games { ð¢ð1 , ð¢ð2 , . . . , ð¢ðð } and nonzero coeï¬cients (ðŒ1 , ðŒ2 , . . . , ðŒð ) such that (Exercise 1.133) ðŒ1 ð¢ð1 + ðŒ2 ð¢ð2 + . . . + ðŒð ð¢ðð = 0 28
(1.16)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Assume that the coalitions are ordered so that ð1 has the smallest number of players of any of the coalitions ð1 , ð2 , . . . , ðð . This implies that no coalition ð2 , ð3 , . . . , ðð is a subset of ð1 and ð¢ðð (ð1 ) = 0
for every ð = 2, 3, . . . , ð
(1.17)
Using (1.39), ð¢ð1 can be expressed as a linear combination of the other games, ð¢ð1 = â1/ðŒ1
ð â
ðŒð ð¢ðð
(1.18)
ð=2
Substituting (1.40) this implies that ð¢ð1 (ð1 ) = 0 whereas ð¢ð (ð ) = 1
for every ð
by deï¬nition. This contradiction establishes that the set ð is linearly independent. 1.136 If ð is a subspace, then 0 â ð and ðŒx1 = 0 with ðŒ â= 0 and x1 = 0 (Exercise 1.122). Therefore ð is linearly dependent (Exercise 1.133). 1.137 Suppose x has two representations, that is x = ðŒ1 x1 + ðŒ2 x2 + . . . + ðŒð xð x = ðœ1 x1 + ðœ2 x2 + . . . + ðœð xð Subtracting 0 = (ðŒ1 â ðœ1 )x1 + (ðŒ2 â ðœ2 )x2 + . . . + (ðŒð â ðœð )xð
(1.19)
Since {x1 , x2 , . . . , , xð } is linearly independent, (1.19) implies that ðŒð = ðœð = 0 for all ð (Exercise 1.133) 1.138 Let ð be the set of all linearly independent subsets of a linear space ð. ð is partially ordered by inclusion. Every chain ð¶ = {ððŒ } â ð has an upper bound, namely ⪠ð. By Zornâs lemma, ð has a maximal element ðµ. We show that ðµ is a basis ðâð¶ for ð. ðµ is linearly independent since ðµ â ð . Suppose that ðµ does not span ð so that lin ðµ â ð. Then there exists some x â ð â lin ðµ. The set ðµ ⪠{x} is a linearly independent and contains ðµ, which contradicts the assumption that ðµ is the maximal element of ð . Consequently, we conclude that ðµ spans ð and hence is a basis. 1.139 Exercise 1.134 established that the set ðµ = { (1, 1, 1), (0, 1, 1), (0, 0, 1)} is linearly independent. Since dim ð
3 = 3, any other vectors must be linearly dependent on ðµ. That is lin ðµ = â3 . ðµ is a basis. By a similar argument to exercise 1.134, it is readily seen that {(1, 0, 0), (0, 1, 0), (0, 0, 1)} is linearly independent and hence constitutes a basis.
29
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.140 Let ðŽ = {a1 , a2 , . . . , að } and ðµ = {b1 , b2 , . . . , bð } be two bases for a linear space ð. Let ð1 = {ð1 } ⪠ðŽ = {b1 , a1 , a2 , . . . , að } ð is linearly dependent (since ð1 â lin ðŽ) and spans ð. ðŒ1 , ðŒ2 , . . . , ðŒð and ðœ1 such that
Therefore, there exists
ðœ1 b1 + ðŒ1 a1 + ðŒ2 a2 + . . . + ðŒð að = 0 At least one ðŒð â= 0. Deleting the corresponding element að , we obtain another set ð1â² of ð elements ð1â² = {b1 , a1 , a2 , . . . , aðâ1 , að+1 , . . . , að } which is also spans ð. Adding the second element from ðµ, we obtain the ð + 1 element set ð2 = {b1 , b2 , a1 , a2 , . . . , aðâ1 , að+1 , . . . , að } which again is linearly dependent and spans ð. Continuing in this way, we can replace ð vectors in ðŽ with the ð vectors from ðµ while maintaining a spanning set. This process cannot eliminate all the vectors in ðŽ, because this would imply that ðµ was linearly dependent. (Otherwise, the remaining bð would be linear combinations of preceding elements of ðµ.) We conclude that necessarily ð †ð. Reversing the process and replacing elements of ðµ with elements of ðŽ establishes that ð †ð. Together these inequalities imply that ð = ð and ðŽ and ðµ have the same number of elements. 1.141 Suppose that the coalitions are ordered in some way, so that ð«(ð ) = {ð0 , ð1 , ð2 , . . . , ð2ð â1 } with ð0 = â
. There are 2ð coalitions. Each game ðº â ð¢ ð corresponds to a unique list of length 2ð of coalitional worths v = (ð£0 , ð£1 , ð£2 , . . . , ð£2ð â1 ) ð
with ð£0 = 0. That is, each game deï¬nes a vector ð£ = (0, ð£1 , . . . , ð£2ð â1 ) â â2 and ð conversely each vector ð£ â â2 (with ð£0 = 0) deï¬nes a game. Therefore, the space of ð all games ð¢ ð is formally identical to the subspace of â2 in which the ï¬rst component ð is identically zero, which in turn is equivalent to the space â2 â1 . Thus, ð¢ ð is a 2ð â 1-dimensional linear space. 1.142 For illustrative purposes, we present two proofs, depending upon whether the linear space is assumed to be ï¬nite dimensional or not. In the ï¬nite dimensional case, a constructive proof is possible, which forms the basis for practical algorithms for constructing a basis. Let ð be a linearly independent set in a linear space ð. ð is ï¬nite dimensional Let ð = dim ð. Assume ð has ð elements and denote it ðð . If lin ðð = ð, then ðð is a basis and we are done. Otherwise, there exists some xð+1 â ð â lin ðð . Adding xð+1 to ðð gives a new set of ð + 1 elements ðð+1 = ðð ⪠{ xð+1 } 30
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
which is also linearly independent ( since xð+1 â / lin ðð ). If lin ðð+1 = ð, then ðð+1 is a basis and we are done. Otherwise, there exists some xð+2 â ð â lin ðð+1 . Adding xð+2 to ðð+1 gives a new set of ð + 2 elements ðð+2 = ðð+1 ⪠{ xð+2 } which is also linearly independent ( since xð+2 â / lin ðð+2 ). Repeating this process, we can construct a sequence of linearly independent sets ðð , ðð+1 , ðð+2 . . . such that lin ðð â« lin ðð+1 â« lin ðð+2 â
â
â
â ð. Eventually, we will reach a set which spans ð and hence is a basis. ð is possibly inï¬nite dimensional For the general case, we can adapt the proof of the existence of a basis (Exercise 1.138), restricting ð to be the class of all linearly independent subsets of ð containing ð. 1.143 Otherwise (if a set of ð + 1 elements was linearly independent), it could be extended to basis at least ð + 1 elements (exercise 1.142). This would contradict the fundamental result that all bases have the same number of elements (Exercise 1.140). 1.144 Every basis is linearly independent. Conversely, let ðµ = {x1 , x2 , . . . , xð } be a set of linearly independent elements in an ð-dimensional linear space ð. We have to show that lin ðµ = ð. Take any x â ð. The set ðµ ⪠{x} = {x1 , x2 , . . . , xð , x } must be linearly dependent (Exercise 1.143). That is there exists numbers ðŒ1 , ðŒ2 , . . . , ðŒð , ðŒ, not all zero, such that ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð + ðŒx = 0
(1.20)
Furthermore, it must be the case that ðŒ â= 0 since otherwise ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð = 0 which contradicts the linear independence of ðŽ. Solving (1.20) for x, we obtain x=
1 ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð ðŒ
Since x was an arbitrary element of ð, we conclude that ðµ spans ð and hence ðµ is a basis. 1.145 A basis spans ð. To establish the converse, assume that ðµ = {x1 , x2 , . . . , xð } is a set of ð elements which span ð. If ð is linearly dependent, then one element is linearly dependent on the other elements. Without loss of generality, assume that x1 â lin ðµ â {x1 }. Deleting x1 the set ðµ â {x1 } = {x2 , x3 , . . . , xð } also spans ð. Continuing in this fashion by eliminating dependent elements, we ï¬nish with a linearly independent set of ð < ð elements which spans ð. That is, we can ï¬nd a basis of ð < ð elements, which contradicts the assumption that the dimension of ð is ð (Exercise 1.140). Thus any set of ð vectors which spans ð must be linearly independent and hence a basis. 31
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.146 We have previously shown â that the set ð is linearly independent (Exercise 1.135). â the space ð¢ ð has dimension 2ðâ1 (Exercise 1.141). There are 2ðâ1 distinct T-unanimity games ð¢ð in ð . Hence ð spans the 2ðâ1 space ð¢ ð . Alternatively, note that any game ð€ â ð¢ ð can be written as a linear combination of T-unanimity games (Exercise 1.75). 1.147 Let ðµ = {x1 , x2 , . . . , xð } be a basis for ð. Since ðµ is linearly independent, ð †ð (Exercise 1.143). There are two possibilities. Case 1: ð = ð. ðµ is a set of ð linearly independent elements in an ð-dimensional space ð. Hence ðµ is a basis for ð and ð = lin ðµ = ð. Case 2: ð < ð. Since ðµ is linearly independent but cannot be a basis for the ðdimensional space ð, we must have ð = lin ðµ â ð. Therefore, we conclude that if ð â ð is a proper subspace, it has a lower dimension than ð. 1.148 Let ðŒ1 , ðŒ2 , ðŒ3 be the coordinates of (1, 1, 1) for the basis {(1, 1, 1), (0, 1, 1), (0, 0, 1)}. That is â â â â â â â â 1 0 0 1 â1â = ðŒ1 â1â + ðŒ2 â1â + ðŒ3 â0â 1 1 1 1 which implies that ðŒ1 = 1, ðŒ2 = ðŒ3 = 0. Therefore (1, 0, 0) are the required coordinates of the (1, 1, 1) with respect to the basis {(1, 1, 1), (0, 1, 1), (0, 0, 1)}. (1, 1, 1) are the coordinates of the vector (1, 1, 1) with respect to the standard basis. 1.149 A subset ð of a linear space ð is a subspace of ð if ðŒx + ðœy â ð for every x, y â ð and for every ðŒ, ðœ â â Letting ðœ = 1 â ðŒ, this implies that ðŒx + (1 â ðŒ)y â ð
for every x, y â ð and ðŒ â â
ð is an aï¬ne set. Conversely, suppose that ð is an aï¬ne set containing 0, that is ðŒx + (1 â ðŒ)y â ð
for every x, y â ð and ðŒ â â
Letting y = 0, this implies that ðŒx â ð
for every x â ð and ðŒ â â
so that ð is homogeneous. Now letting ðŒ = 12 , for every x and y in ð, 1 1 x+ y âð 2 2 and homogeneity implies
( x+y =2
1 1 x+ y 2 2
ð is also additive. Hence ð is subspace. 32
) âð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.150 For any x â ð, let ð = ð âx = {v â ð : v +x â ð } ð is an aï¬ne set For any v1 , v2 â ð , there exist corresponding s1 , s2 â ð such that v1 = s1 â x and v2 = s2 â x and therefore ðŒv1 + (1 â ðŒ)v2 = ðŒ(s1 â x) + (1 â ðŒ)(s1 â x) = ðŒs1 + (1 â ðŒ)s2 â ðŒx + (1 â ðŒ)x =sâx where s = ðŒð 1 + (1 â ðŒ)ð 2 â ð. There ð is an aï¬ne set. ð is a subspace Since x â ð, 0 = x â x â ð . Therefore ð is a subspace (Exercise 1.149). ð is unique Suppose that there are two subspaces ð 1 and ð 2 such that ð = ð 1 + x1 and ð = ð 2 + x2 . Then ð1 + x1 = ð2 + x2 ð1 = ð2 + (x2 â x1 ) = ð2 + x where x = x2 â x1 â ð. Therefore ð1 is parallel to ð2 . Since ð1 is a subspace, 0 â ð1 which implies that âx â ð2 . Since ð2 is a subspace, this implies that x â ð2 and ð2 + x â ð2 . Therefore ð1 = ð2 + x â ð2 . Similarly, ð2 â ð1 and hence ð1 = ð2 . Therefore the subspace ð is unique. 1.151 Let ð ⥠ð denote the relation ð is parallel to ð , that is ð ⥠ð ââ ð = ð + x for some x â ð The relation ⥠is reï¬exive ð ⥠ð since ð = ð + 0 transitive Assume ð = ð + x and ð = ð + y. Then ð = ð + (x + y) symmetric ð = ð + x =â ð = ð + (âx) Therefore ⥠is an equivalence relation. 1.152 See exercises 1.130 and 1.162. 1.153
1. Exercise 1.150
2. Assume x0 â ð . For every x â ð» x = x0 + v = w â ð which implies that ð» â ð . Conversely, assume ð» = ð . Then x0 = 0 â ð since ð is a subspace. 3. By deï¬nition, ð» â ð. Therefore ð = ð» â x â ð. / ð . Suppose to the contrary 4. Let x1 â lin {x1 , ð } = ð â² â ð
33
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Then ð» â² = x0 + ð â² is an aï¬ne set (Exercise 1.150) which strictly contains ð». This contradicts the deï¬nition of ð» as a maximal proper aï¬ne set. 5. Let x1 â / ð . By the previous part, x â lin {x1 , ð }. That is, there exists ðŒ â â such that x = ðŒx1 + v for some v â ð To see that ðŒ is unique, suppose that there exists ðœ â â such that x = ðœx1 + vâ² for some vâ² â ð Subtracting 0 = (ðŒ â ðœ)x1 + (v â vâ² ) / ð. which implies that ðŒ = ðœ since x1 â 1.154 Assume x, y â ð. That is, x, y â âð and â â ð¥ð = ðŠð = ð€(ð ) ðâð
ðâð
ð
For any ðŒ â â, ðŒx + (1 â ðŒ)y â â and â â â ðŒð¥ð + (1 â ðŒ)ðŠð = ðŒ ð¥ð + (1 â ðŒ) ðŠð ðâð
ðâð
ðâð
= ðŒð€(ð ) + (1 â ðŒ)ð€(ð ) = ð€(ð ) Hence ð is an aï¬ne subset of âð . 1.155 See Exercise 1.129. 1.156 No. A straight line through any two points in âð+ extends outside âð+ . Put diï¬erently, the aï¬ne hull of âð+ is the whole space âð . 1.157 Let ð = aï¬ ð â x1 = aï¬ {0, x2 â x1 , x3 â x1 , . . . , xð â x1 } ð is a subspace (0 â ð ) and aï¬ ð = ð + x1 and dim aï¬ ð = dim ð Note that the choice of x1 is arbitrary. ð is aï¬nely dependent if and only if there exists some xð â ð such that xð â â x1 . aï¬ (ð â {xð }). Since the choice of x1 is arbitrary, we assume that xð = xð â aï¬ (ð â {xð }) ââ xð â (ð + x1 ) â {xð } ââ xð â x1 â ð â {xð â x1 } ââ xð â x1 â lin {x2 â x1 , x3 â x1 , . . . , xðâ1 â x1 , . . . , xð+1 â x1 , . . . , xð â x1 } 34
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
Therefore, ð is aï¬nely dependent if and only if {x2 â x1 , x3 â x1 , . . . , xð â x1 } is linearly independent. 1.158 By the previous exercise, the set ð = {x1 , x2 , . . . , xð } is aï¬nely dependent if and only if the set {x2 â x1 , x3 â x1 , . . . , xð â x1 } is linearly dependent, so that there exist numbers ðŒ2 , ðŒ3 , . . . , ðŒð , not all zero, such that ðŒ2 (x2 â x1 ) + ðŒ3 (x3 â x1 ) + â
â
â
+ ðŒð (xð â x1 ) = 0 or ðŒ2 x2 + ðŒ3 x3 + â
â
â
+ ðŒð xð â
ð â
ðŒð x1 = 0
ð=2
Let ðŒ1 = â
âð
ð=2
ðŒð . Then ðŒ1 x1 + ðŒ2 x2 + . . . + ðŒð xð = 0
and ðŒ1 + ðŒ2 + . . . + ðŒð = 0 as required. 1.159 Let ð = aï¬ ð â x1 = aï¬ { 0, x2 â x1 , x3 â x1 , . . . , xð â x1 } Then aï¬ ð = x1 + ð If ð is aï¬nely independent, every x â aï¬ ð has a unique representation as x = x1 + v,
vâð
with v = ðŒ2 (x2 â x1 ) + ðŒ3 (x3 â x1 ) + â
â
â
+ ðŒð (xð â x1 ) so that x = x1 + ðŒ2 (x2 â x1 ) + ðŒ3 (x3 â x1 ) + â
â
â
+ ðŒð (xð â x1 ) âð Deï¬ne ðŒ1 = 1 â ð=2 ðŒð . Then x = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð with ðŒ1 + ðŒ2 + â
â
â
+ ðŒð = 1 x is a unique aï¬ne combination of the elements of ð. 1.160 Assume that ð¥, ðŠ â (ð, ð) â â. This means that ð < ð¥ < ð and ð < ðŠ < ð. For every 0 †ðŒ †1 ðŒð¥ + (1 â ðŒ)ðŠ > ðŒð + (1 â ðŒ)ð 35
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
and ðŒð¥ + (1 â ðŒ)ðŠ < ðŒð + (1 â ðŒ)ð Therefore ð < ðŒð¥+(1âðŒ)ðŠ < ð and ðŒð¥+(1âðŒ)ðŠ â (ð, ð). (ð, ð) is convex. Substituting †for < demonstrates that [ð, ð] is convex. Let ð be an arbitrary convex set in â. Assume that ð is not an interval. This implies that there exist numbers ð¥, ðŠ, ð§ such that ð¥ < ðŠ < ð§ and ð¥, ð§ â ð while ðŠ â / ð. Deï¬ne ðŒ=
ð§âðŠ ð§âð¥
so that 1âðŒ=
ðŠâð¥ ð§âð¥
Note that 0 †ðŒ †1 and that ðŒð¥ + (1 â ðŒ)ð§ =
ðŠâð¥ ð§âðŠ ð¥+ ð§=ðŠâ /ð ð§âð¥ ð§âð¥
which contradicts the assumption that ð is convex. We conclude that every convex set in â is an interval. Note that ð may be a hybrid interval such (ð, ð] or [ð, ð) as well as an open (ð, ð) or closed [ð, ð] interval. 1.161 Let (ð, ð€) be a TP-coalitional game. If core(ð, ð€) = â
then it is trivially convex. Otherwise, assume core(ð, ð€) is nonempty and let x1 and x2 belong to core(ð, ð€). That is â ð¥1ð ⥠ð€(ð) for every ð â ð ðâð
â
ð¥1ð = ð€(ð )
ðâð
and therefore for any 0 †ðŒ †1 â ðŒð¥1ð ⥠ðŒð€(ð)
for every ð â ð
ðâð
â
ðŒð¥1ð = ðŒð€(ð )
ðâð
Similarly â
(1 â ðŒ)ð¥2ð ⥠(1 â ðŒ)ð€(ð)
for every ð â ð
ðâð
â
(1 â ðŒ)ð¥2ð = (1 â ðŒ)ð€(ð )
ðâð
Summing these two systems â ðŒð¥1ð + (1 â ðŒ)ð¥2ð ⥠ðŒð€(ð) + (1 â ðŒ)ð€(ð) = ð€(ð) ðâð
â
ðŒð¥1ð + (1 â ðŒ)ð¥2ð = ðŒð€(ð ) + (1 â ðŒ)ð€(ð ) = ð€(ð )
ðâð
That is, ðŒð¥1ð + (1 â ðŒ)ð¥2ð belongs to core(ð, ð€). 36
for every ð â ð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
â© 1.162 Let â be a collection of convex sets and let x, y belong to ðââ ð. for every ð â â, x, y â ð and therefore â© ðŒx + (1 â ðŒ)y â ð for all 0 †ðŒ †1 (since ð is convex). Therefore ðŒx + (1 â ðŒ)y â ðââ ð. 1.163 Fix some output ðŠ. Assume that x1 , x2 â ð (ðŠ). This implies that both (ðŠ, âx1 ) and (ðŠ, âx2 ) belong to the production possibility set ð . If ð is convex ðŒ(ðŠ, âx1 ) + (1 â ðŒ)(ðŠ, âx2 ) = (ðŒðŠ + (1 â ðŒ)ðŠ, ðŒx1 + (1 â ðŒ)x2 ) = (ðŠ, ðŒx1 + (1 â ðŒ)x2 ) â ð for every ðŒ â [0, 1]. This implies that ðŒx1 + (1 â ðŒ)x2 â ð (ðŠ). Since the choice of ðŠ was arbitrary, this implies that ð (ðŠ) is convex for every ðŠ. 1.164 Assume ð1 and ð2 are convex sets. Let ð = ð1 + ð2 . Suppose x, y belong to ð. Then there exist s1 , t1 â ð1 and s2 , t2 â ð2 such that x = s1 + s2 and y = t1 + t2 . For any ðŒ â [0, 1] ðŒx + (1 â ðŒ)y = ðŒs1 + s2 + (1 â ðŒ)t1 + t2 = ðŒs1 + (1 â ðŒ)t1 + ðŒs2 + (1 â ðŒ)t2 â ð since ðŒs1 + (1 â ðŒ)t1 â ð1 and ðŒs2 + (1 â ðŒ)t2 â ð2 . The argument readily extends to any ï¬nite number of sets. 1.165 Without loss of generality, assume that ð = 2. Let ð = ð1 à ð2 â ð = ð1 à ð2 . Suppose x = (ð¥1 , ð¥2 ) and y = (ðŠ1 , ðŠ2 ) belong to ð. Then ðŒx + (1 â ðŒ)y = ðŒ(ð¥1 , ð¥2 ) + (1 â ðŒ)(ðŠ1 , ðŠ2 ) = (ðŒð¥1 , ðŒð¥2 ) + ((1 â ðŒ)ðŠ1 , (1 â ðŒ)ðŠ2 ) = (ðŒð¥1 + (1 â ðŒ)ðŠ1 , ðŒð¥2 + (1 â ðŒ)ðŠ2 ) â ð 1.166 Let ðŒx, ðŒy be points in ðŒð so that x, y â ð. Since ð is convex, ðœx+ (1 â ðœ)y â ð for every 0 †ðœ †1. Multiplying by ðŒ ðŒ(ðœx + (1 â ðœ)y) = ðœ(ðŒx) + (1 â ðœ)(ðŒy) â ðŒð Therefore, ðŒð is convex. 1.167 Combine Exercises 1.164 and 1.166. 1.168 The inclusion ð â ðŒð + (1 â ðŒ)ð is true for any set (whether convex or not), since for every x â ð x = ðŒx + (1 â ðŒ)x â ðŒð + (1 â ðŒ)ð The reverse inclusion ðŒð +(1âðŒ)ð â ð follows directly from the deï¬nition of convexity. 1.169 Given any two convex sets ð and ð in a linear space, the largest convex set contained in both is ð â© ð ; the smallest convex set containing both is conv ð ⪠ð . Therefore, the set of all convex sets is a lattice with ð â§ð =ð â©ð ð âš ð = conv ð ⪠ð The lattice is complete since every collection {ðð } has a least upper bound conv ⪠ðð and a greatest lower bound â©ðð . 37
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.170 If a set contains all convex combinations of its elements, it contains all convex combinations of any two points, and hence is convex. Conversely, assume that ð is convex. Let x be a convex combination of elements in ð, that is let x = ðŒ1 x1 + ðŒ2 x2 + . . . + ðŒð xð where x1 , x2 , . . . , xð â ð and ðŒ1 , ðŒ2 , . . . , ðŒð â â+ with ðŒ1 + ðŒ2 + . . . + ðŒð = 1. We need to show that x â ð. We proceed by induction of the number of points ð. Clearly, x â ð if ð = 1 or ð = 2. To show that it is true for ð = 3, let x = ðŒ1 x1 + ðŒ2 x2 + ðŒ3 x3 where x1 , x2 , x3 â ð and ðŒ1 , ðŒ2 , ðŒ3 â â+ with ðŒ1 + ðŒ2 + ðŒ3 = 1. Assume that ðŒð > 0 for all ð (otherwise ð = 1 or ð = 2) so that ðŒ1 < 1. Rewriting x = ðŒ1 x1 + ðŒ2 x2 + ðŒ3 x3 ( ) ðŒ2 ðŒ2 = ðŒ1 x1 + (1 â ðŒ1 ) x2 + x3 1 â ðŒ1 1 â ðŒ1 = ðŒ1 x1 + (1 â ðŒ1 )y where ( y=
ðŒ2 ðŒ2 x2 + x3 1 â ðŒ1 1 â ðŒ1
)
y is a convex combination of two elements x2 and x3 since ðŒ2 ðŒ2 ðŒ2 + ðŒ3 + = =1 1 â ðŒ1 1 â ðŒ1 1 â ðŒ1 and ðŒ2 + ðŒ3 = 1 â ðŒ1 . Hence y â ð. Therefore x is a convex combination of two elements x1 and ðŠ and is also in ð. Proceeding in this fashion, we can show that every convex combination belongs to ð, that is conv ð â ð. 1.171 This is precisely analogous to Exercise 1.128. We observe that 1. conv ð is a convex set. 2. if ð¶ is any convex set containing ð, then conv ð â ð¶. Therefore, conv ð is the smallest convex set containing S. 1.172 Note ï¬rst that ð â conv ð for any set ð. The converse for convex sets follows from Exercise 1.170. 1.173 Assume x â conv (ð1 + ð2 ). Then, there exist numbers ðŒ1 , ðŒ2 , . . . , ðŒð and vectors x1 , x2 , . . . , xð in ð1 + ð2 such that x = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð For every xð , there exists x1ð â ð1 and x2ð â ð2 such that xð = x1ð + x2ð
38
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
and therefore ð â
x=
ðŒð x1ð +
ð=1
ð â
ðŒð x2ð
ð=1
= x1 + x2 âð âð where x1 = ð=1 ðŒð x1ð â ð1 and x2 = ð=1 ðŒð x2ð â ð2 . Therefore x â conv ð1 + conv ð2 . Conversely, assume that x â conv ð1 + conv ð2 . Then x = x1 + x2 , where x1 =
ð â
ðŒð ð¥1ð ,
x1ð â ð1
ðœð ð¥2ð ,
x2ð â ð2
ð=1
x2 =
ð â ð=1
and x = x1 + x2 =
ð â
ðŒð ð¥1ð +
ð=1
ð â
ðœð ð¥2ð â conv (ð1 + ð2 )
ð=1
since x1ð , x2ð â ð1 + ð2 for every ð and ð. 1.174 The dimension of the input requirement set ð (ðŠ) is ð. Its aï¬ne hull is âð . 1.175
1. Let x = ðŒ1 x1 + ðŒ2 x2 + . . . + ðŒð xð
(1.21)
If ð > dim ð +1, the elements x1 , x2 , . . . , xð â ð are aï¬nely dependent (Exercise 1.157 and therefore there exist numbers ðœ1 , ðœ2 , . . . , ðœð , not all zero, such that (Exercise 1.158) ðœ1 x1 + ðœ2 x2 + . . . + ðœð xð = 0
(1.22)
and ðœ1 + ðœ2 + . . . + ðœð = 0 2. Combining (1.21) and (1.22) x = x â ð¡0 ð ð â â ðŒð xð â ð¡ ðœð xð = ð=1
=
ð â
ð=1
(ðŒð â ð¡ðœð )xð
ð=1
for any ð¡ â â. } { 3. Let ð¡ = minð ðŒðœðð : ðœð > 0 =
ðŒð ðœð
We note that â ð¡ > 0 since ðŒð > 0 for every ð. 39
(1.23)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
â If ðœð > 0, then ðŒð /ðœð ⥠ðŒð /ðœð ⥠ð¡ and therefore ðŒð â ð¡ðœð ⥠0 â If ðœð †0 then ðŒð â ð¡ðœð > 0 for every ð¡ > 0. â Therefore ðŒð â ð¡ðœð ⥠0 for every ð¡ and â ðŒð â ð¡ðœð = 0 for ð = ð. Therefore, (1.23) represents x as a convex combination of only ð â 1 points. 4. This process can be repeated until x is represented as a convex combination of at most dim ð + 1 elements. 1.176 Assume x is not an extreme point of ð. Then there exists distinct x1 and x2 in S such that x = ðŒx1 + (1 â ðŒ)x2 Without loss of generality, assume ðŒ †1/2 and let y = x2 â x. Then x + y = x2 â ð. Furthermore x â y = x â x2 + x = 2x â x2 = 2(ðŒx1 + (1 â ðŒ)x2 ) â x2 = 2ðŒx1 + (1 â 2ðŒ)x2 â ð since ðŒ †1/2. 1.177
1. For any x = (ð¥1 , ð¥2 ) â ð¶2 , there exists some ðŒ1 â [0, 1] such that ð¥1 = ðŒ1 ð + (1 â ðŒ1 )(âð) = (2ðŒ1 â 1)ð In fact, ðŒ1 is deï¬ned by ðŒ1 = Therefore (see Figure 1.5) ) ( ( ð¥1 = ðŒ1 ð ( ) ( ð¥1 = ðŒ1 âð
ð¥1 + ð 2ð
) âð + (1 â ðŒ1 ) ð ( ) ) âð ð + (1 â ðŒ1 ) âð âð ð ð
)
(
Similarly ð¥2 = ðŒ2 ð + (1 â ðŒ2 )(âð) where ðŒ2 =
ð¥2 + ð 2ð
Therefore, for any x â ð¶2 , ( ( ) ) ) ( ð¥1 ð¥1 ð¥1 x= + (1 â ðŒ2 ) = ðŒ2 ð¥2 ð âð ( ) ( ) ð âð = ðŒ1 ðŒ2 + (1 â ðŒ1 )ðŒ2 ð ð ( ) ( ) ð âð + ðŒ1 (1 â ðŒ2 ) + (1 â ðŒ1 )(1 â ðŒ2 ) âð âð ( ) ( ) ( ) ( ) ð âð ð âð = ðœ1 + ðœ2 + ðœ3 + ðœ4 ð ð âð âð 40
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
(ð¥1 , c) x ð¥1
(ð¥1 , -c) Figure 1.5: A cube in â2 where 0 †ðœð †1 and ðœ1 + ðœ2 + ðœ3 + ðœ4 = ðŒ1 ðŒ2 + (1 â ðŒ1 )ðŒ2 + ðŒ1 (1 â ðŒ2 ) + (1 â ðŒ1 )(1 â ðŒ2 ) = ðŒ1 ðŒ2 + ðŒ2 â ðŒ1 ðŒ2 + ðŒ1 â ðŒ1 ðŒ2 + 1 â ðŒ1 â ðŒ2 + ðŒ1 ðŒ2 =1 That is
{( ð¥ â conv
ð ð
) ( ) ( ) ( )} âð ð âð , , , ð âð âð
2. (a) For any point (ð¥1 , ð¥2 , . . . , ð¥ðâ1 , ð) which lies on face of the cube ð¶ð , (ð¥1 , ð¥2 , . . . , ð¥ðâ1 ) â ð¶ðâ1 and therefore (ð¥1 , ð¥2 , . . . , ð¥ðâ1 ) â conv { ±ð, ±ð, . . . , ±ð) } â âðâ1 so that x â conv { (±ð, ±ð, . . . , ±ð, ð) } â âð Similarly, any point (ð¥1 , ð¥2 , . . . , ð¥ðâ1 , âð) on the opposite face lies in the convex hull of the points { (±ð, ±ð, . . . , ±ð, âð) }. (b) For any other point x = (ð¥1 , ð¥2 , . . . , ð¥ð ) â ð¶ð , let ðŒð =
ð¥ð + ð 2ð
so that ð¥ð = ðŒð ð + (1 â ðŒð )(âð) Then
â
â â â â â ð¥1 ð¥1 ð¥1 â ð¥2 â â ð¥2 â â ð¥2 â â â â â â â â â â â â x = â . . . â = ðŒð â . . . â + (1 â ðŒð ) â â ... â âð¥ðâ1 â âð¥ðâ1 â âð¥ðâ1 â ð¥ð ð âð 41
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Hence
x â conv { (±ð, ±ð, . . . , ±ð) } â âð In other words ð¶ð â conv { (±ð, ±ð, . . . , ±ð) } â âð 3. Let ðž denote the set of points of the form { (±ð, ±ð, . . . , ±ð) } â âð . Clearly, every point in ðž is an extreme point of ð¶ð . Conversely, we have shown that ð¶ð â conv ðž. Therefore, no point x â ð¶ ð â ðž can be an extreme point of ð¶ ð . ðž is the set of extreme points of ð¶ ð . 4. Since ð¶ ð is convex, and ðž â ð¶ð , conv ðž â ð¶ ð . Consequently, ð¶ ð = conv ðž. 1.178 Let x, y belong to ð â ð¹ is convex. For any ðŒ â [0, 1] â ðŒx + (1 â ðŒ)y â ð since ð convex â ðŒx + (1 â ðŒ)y â / ð¹ since ð¹ is a face Thus ðŒx + (1 â ðŒ)y â ð â ð¹ which is convex. 1.179
1. Trivial.
⪠2. Let {ð¹ð } be a collection of faces of ð and let ð¹ = ð¹ð . Choose any x, y â ð. If the line segment between x and y intersects ð¹ , then ⪠it intersects some face ð¹ð which implies that x, y â ð¹ð . Therefore, x, y â ð¹ = ð¹ð . â© 3. Let {ð¹ð } be a collection of faces of ð and let ð¹ = ð¹ð . Choose any x, y â ð. if the line segment between x and y intersects ð¹ , then it intersects ⪠every face ð¹ð which implies that x, y â ð¹ð for every ð. Therefore, x, y â ð¹ = ð¹ð . 4. Let ð be the collection of all faces of ð. This is partially ordered by inclusion. By â© the previous result, every nonempty subcollection ð has a least upper bound ( ð¹ âð ð¹ ). Hence ð is a complete lattice (Exercise 1.47).
1.180 Let ð be a polytope. Then ð = conv { x1 , x2 , . . . , xð }. Note that every extreme point belongs to { x1 , x2 , . . . , xð }. Now choose the smallest subset whose convex hull is still ð, that is delete elements which can be written as convex combinations of other elements. Suppose the minimal subset is { x1 , x2 , . . . , xð }. We claim that each of these elements is an extreme point of ð, that is { x1 , x2 , . . . , xð } = ðž. Assume not, that is assume that xð is not an extreme point so that there exists x, y â ð with xð = ðŒx + (1 â ðŒ)y
with 0 < ðŒ < 1
(1.24)
Since x, y â conv {x1 , x2 , . . . , xð } x=
ð â
ðŒð xð
y=
ð=1
ð â
ðœxð
ð=1
Substituting in (1.24), we can write xð as a convex combination of {x1 , x2 , . . . , xð }. xð =
ð ð â â ( ) ðŒðŒð + (1 â ðŒ)ðœð xð = ðŸð xð ð=1
ð=1
where ðŸð = ðŒðŒð + (1 â ðŒ)ðœð 42
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
Note that 0 †ðŸð †1, so that either ðŸð < 1 or ðŸð = 1. We show that both cases lead to a contradiction. â ðŸð < 1. Then ðâ1 â( ) 1 ðŒðŒð + (1 â ðŒ)ðœð xð 1 â ðŸð ð=1
xð =
which contradicts the minimality of the set {x1 , x2 , . . . , xð }. â ðŸð = 1. Then ðŸð = 0 for every ð â= ð. That is ðŒðŒð + (1 â ðŒ)ðœð = 0 which implies that ðŒð = ðœð
for every ð â= ð
for every ð â= ð and therefore x = y.
Therefore, if {x1 , x2 , . . . , xð } is a minimal spanning set, every point must be an extreme point. 1.181 Assume to the contrary that one of the vertices is not an extreme point of the simplex. Without loss of generality, assume this is x1 . Then, there exist distinct y, z â ð and 0 < ðŒ < 1 such that x1 = ðŒy + (1 â ðŒ)z
(1.25)
Now, since y â ð, there exist ðœ1 , ðœ2 , . . . , ðœð such that y=
ð â
ðœð xð ,
ð=1
ð â
ðœð = 1
ð=1
Similarly, there exist ð¿1 , ð¿2 , . . . , ð¿ð such that z=
ð â
ð â
ð¿ð xð ,
ð=1
ð¿ð = 1
ð=1
Substituting in (1.25) x1 = ðŒ =
ð â
ðœð xð + (1 â ðŒ)
ð=1 ð â
ð â
ð¿ð xð
ð=1
( ) ðŒðœð + (1 â ðŒ)ð¿ð xð
ð=1
Since
âð
ð=1
( ) â â ðŒðœð + (1 â ðŒ)ð¿ð = ðŒ ðð=1 ðœð + (1 â ðŒ) ð=1 ð¿ð = 1 x1 =
ð â ( ) ðŒðœð + (1 â ðŒ)ð¿ð xð ð=1
Subtracting, this implies 0=
ð â ( ) ðŒðœð + (1 â ðŒ)ð¿ð (xð â x1 ) ð=2
This establishes that the set {x2 â x1 , x3 â x1 , . . . , xð â x1 } is linearly dependent and therefore ðž = {x1 , x2 , . . . , xð } is aï¬nely dependent (Exercise 1.157). This contradicts the assumption that ð is a simplex. 43
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.182 Let ð be the dimension of a convex set ð in a linear space ð. Then ð = dim aï¬ ð and there exists a set { x1 , x2 , . . . , xð+1 } of aï¬nely independent points in ð. Deï¬ne ð â² = conv { x1 , x2 , . . . , xð+1 } Then ð â² is an ð-dimensional simplex contained in ð. 1.183 Let w = (ð€({1}), ð€({2}), . . . , ð€({ð})) denote the vector of individual worths and let ð denote the surplus to be distributed, that is â ð = ð€(ð ) â ð€({ð}) ðâð
ð > 0 if the game is essential. For each player ð = 1, 2, . . . , ð, let yð = w + ð eð be the outcome in which player ð receives the entire surplus. (eð is the ðth unit vector.) Note that { ð€({ð}) + ð ð = ð ð ðŠð = ð€({ð}) ð â= ð Each yð is an imputation since ðŠðð ⥠ð€({ð}) and â â ðŠðð = ð€({ð}) + ð = ð€(ð ) ðâð
ðâð
Therefore {y1 , y2 , . . . , yð } â ðŒ. Since ðŒ is convex (why ?), ð = conv {y1 , y2 , . . . , yð } â ðŒ. Further, for every ð, ð â ð the vectors yð â yð = ð (eð â eð ) are linearly independent. Therefore ð is an ð â 1-dimensional simplex in âð . For any x â ðŒ deï¬ne ðŒð =
ð¥ð â ð€({ð}) ð
so that ð¥ð = ð€({ð}) + ðŒð ð Since x is an imputation â ðŒð ⥠0 (â ) â â â ðâð ðŒð = ðâð ð¥ð â ðâð ð€({ð}) /ð = 1 â We claim that x = ðâð ðŒð yð since for each ð = 1, 2, . . . , ð â â ðŒð ðŠðð = ðŒð ð€({ð}) + ðŒð ð ðâð
ðâð
= ð€({ð}) + ðŒð ð = ð¥ð Therefore x â conv {y1 , y2 , . . . , yð } = ð, that is ðŒ â ð. Since we previously showed that ð â ðŒ, we have established that ðŒ = ð, which is an ð â 1 dimensional simplex in âð . 44
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
ð¥2
ð¥2
ð¥2
ð¥1
ð¥1
ð¥1
1. A non-convex cone
2. A convex set
3. A convex cone
Figure 1.6: A cone which is not convex, a convex set and a convex cone 1.184 See Figure 1.6. 1.185 Let x = (ð¥1 , ð¥2 , . . . , ð¥ð ) belong to âð+ , which means that ð¥ð ⥠0 for every ð. For every ðŒ > 0 ðŒx = (ðŒx1 , ðŒx2 , . . . , ðŒxð ) and ðŒð¥ð ⥠0 for every ð. Therefore ðŒx â âð+ . âð+ is a cone in âð . 1.186 Assume ðŒx + ðœy â ð for every x, y â ð and ðŒ, ðœ â â+
(1.26)
Letting ðœ = 0, this implies that ðŒx â ð for every x â ð and ðŒ â â+ so that ð is a cone. To show that ð is convex, let x and y be any two elements in ð. For any ðŒ â [0, 1], (1.26) implies that ðŒx + (1 â ðŒ)y â ð Therefore ð is convex. Conversely, assume that ð is a convex cone. For any ðŒ, ðœ â â+ and x, y â ð ðœ ðŒ x+ yâð ðŒ+ðœ ðŒ+ðœ and therefore ðŒx + ðœy â ð 1.187 Assume ð satisï¬es 1. ðŒð â ð for every ðŒ ⥠0 2. ð + ð â ð
45
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
By (1), ð is a cone. To show that it is convex, let x and y belong to ð. By (1), ðŒx and (1 â ðŒ)y belong to ð, and therefore ðŒx + (1 â ðŒ)y belongs to ð by (2). ð is convex. Conversely, assume that ð is a convex cone. Then ðŒð â ð
for every ðŒ ⥠0
Let x and y be any two elements in ð. Since ð is convex, ð§ = ðŒ 12 x + (1 â ðŒ) 12 y â ð and since it is a cone, 2ð§ = x + y â ð. Therefore ð +ð âð 1.188 We have to show that ð is convex cone. By assumption, ð is convex. To show that ð is a cone, let y be any production plan in ð . By convexity ðŒy = ðŒy + (1 â ðŒ)0 â ð for every 0 †ðŒ †1 Repeated use of additivity ensures that ðŒy â ð for every ðŒ = 1, 2, . . . Combining these two conclusions implies that ðŒy â ð for every ðŒ ⥠0 1.189 Let ð® â ð¢ ð denote the set of all superadditive games. Let ð€1 , ð€2 â ð be two superadditive games. Then, for all distinct coalitions ð, ð â ð with ð â© ð = â
ð€1 (ð ⪠ð ) ⥠ð€1 (ð) + ð€1 (ð ) ð€2 (ð ⪠ð ) ⥠ð€2 (ð) + ð€2 (ð ) Adding (ð€1 + ð€2 )(ð ⪠ð ) = ð€1 (ð ⪠ð ) + ð€2 (ð ⪠ð ) ⥠ð€1 (ð) + ð€2 (ð) + ð€1 (ð ) + ð€2 (ð ) = (ð€1 + ð€2 )(ð) + (ð€1 + ð€2 )(ð ) so that ð€1 + ð€2 is superadditive. Similarly, we can show that ðŒð€1 is superadditive for all ðŒ â â+ . Hence ð® is a convex cone in ð¢ ð . â©ð 1.190 Let x belong to ð=1 ðð . Then x â ð â©ð for every ð. Since each ðð is a cone, ðŒx â ðð for every ðŒ ⥠0 and therefore ðŒx â ðð=1 ðð . Let ð = ð1 + ð2 + â
â
â
+ ðð and assume x belongs to ð. Then there exist xð â ðð , ð = 1, 2, . . . , ð such that x = x1 + x2 + â
â
â
+ xð For any ðŒ ⥠0 ðŒx = ðŒ(x1 + x2 + â
â
â
+ xð ) = ðŒx1 + ðŒx2 + â
â
â
+ ðŒxð â ð since ðŒxð â ðð for every ð.
46
Solutions for Foundations of Mathematical Economics 1.191
c 2001 Michael Carter â All rights reserved
1. Suppose that y â ð . Then, there exist ðŒ, ðŒ2 , . . . , ðŒ8 ⥠0 such that y=
8 â
ðŒð yð
ð=1
and for the ï¬rst commodity 8 â
y1 =
ðŒð ðŠð1
ð=1
If y â= 0, at least one of the ðŒð > 0 and hence y1 < 0 since ðŠð1 < 0 for ð = 1, 2, . . . , 8. 2. Free disposal requires that y â ð, yⲠ†y =â yâ² â ð . Consider the production plan yâ² = (â2, â2, â2, â2). Note that yⲠ†y3 and yⲠ†y6 . Suppose that yâ² â ð . Then there exist ðŒ1 , ðŒ2 , . . . , ðŒ8 ⥠0 such that y=
8 â
ðŒð yð
ð=1
For the third commodity (component), we have 4ðŒ1 + 3ðŒ2 + 3ðŒ3 + 3ðŒ4 + 12ðŒ5 â 2ðŒ6 + 5ðŒ8 = â2
(1.27)
and for the fourth commodity 2ðŒ2 â 1ðŒ3 + 1ðŒ4 + 5ðŒ6 + 10ðŒ7 â 2ðŒ8 = â2
(1.28)
Adding to (1.28) to (1.27) gives 4ðŒ1 + 5ðŒ2 + 2ðŒ3 + 4ðŒ4 + 12ðŒ5 + 3ðŒ6 + 10ðŒ7 + 3ðŒ8 = â4 / ð. which is impossible given that ðŒð ⥠0. Therefore, we conclude that yâ² â 3. y2 = (â7, â9, 3, 2) ⥠(â8, â13, 3, 1) = y4 3y1 = (â9, â18, 12, 0) ⥠(â11, â19, 12, 0) = y5 y7 = (â8, â5, 0, 10) ⥠(â8, â6, â4, 10) = 2y6 2y3 = (â2, â4, 6, â2) ⥠(â2, â4, 5, â2) = y8 4. 2y3 + y7 = (â2, â4, 6, â2) + (â8, â5, 0, 10) = (â10, â9, 6, 8) ⥠(â14, â18, 6, 4) = 2y2 20y3 + 2y7 = 20(â1, â2, 3, â1) + 2(â8, â5, 0, 10) = (â20, â40, 60, â20) + (â16, â10, 0, 20) = (â36, â50, 60, 0) ⥠(â45, â90, 60, 0) = 15y1 47
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5. Eï¬(ð ) = cone { y3 , y7 }. 1.192 This is precisely analogous to Exercise 1.128. We observe that 1. cone ð is a convex cone. 2. if ð¶ is any convex cone containing ð, then conv ð â ð¶. Therefore, cone ð is the smallest convex cone containing S. 1.193 For any set ð, ð â cone ð. If ð is a convex cone, Exercise 1.186 implies that cone ð â ð. 1.194
1. If ð > ð = dim cone ð = dim lin ð, the elements x1 , x2 , . . . , xð â ð are linearly dependent and therefore there exist numbers ðœ1 , ðœ2 , . . . , ðœð , not all zero, such that (Exercise 1.134) ðœ1 x1 + ðœ2 x2 + . . . + ðœð xð = 0
(1.29)
2. Combining (1.14) and (1.29) x = x â ð¡0 ð ð â â = ðŒð xð â ð¡ ðœð xð ð=1
=
ð â
ð=1
(ðŒð â ð¡ðœð )xð
(1.30)
ð=1
for any ð¡ â â. { } 3. Let ð¡ = minð ðŒðœðð : ðœð > 0 =
ðŒð ðœð
We note that â ð¡ > 0 since ðŒð > 0 for every ð. â If ðœð > 0, then ðŒð /ðœð ⥠ðŒð /ðœð ⥠ð¡ and therefore ðŒð â ð¡ðœð ⥠0. â If ðœð †0 then ðŒð â ð¡ðœð > 0 for every ð¡ > 0. â Therefore ðŒð â ð¡ðœð ⥠0 for every ð¡ and â ðŒð â ð¡ðœð = 0 for ð = ð. Therefore, (1.30) represents x as a nonnegative combination of only ð â 1 points. 4. This process can be repeated until x is represented as a convex combination of at most ð points. 1.195
1. The aï¬ne hull of ðË is parallel to the aï¬ne hull of ð. Therefore ( Since
0 0
)
dim ð = dim aï¬ ð = dim aï¬ ðË Ë â / aï¬ ð, dim cone ðË = dim aï¬ ðË + 1 = dim ð + 1
48
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
( ) x 2. For every x â conv ð, â conv ðË and there exist (Exercise 1.194) ð + 1 1 ( ) xð points â ðË such that 1 (
x 1
)
( â conv {
x1 1
) ( ) ( ) x2 xð+1 , ,... } 1 1
This implies that x â conv { x1 , x2 , . . . , xð+1 } with x1 , x2 , . . . , xð+1 â ð. 1.196 A subsimplex with precisely one distinguished face is completely labeled. Suppose a subsimplex has more than one distinguished face. This means that it has vertices labeled 1, 2, . . . , ð. Since it has ð + 1 vertices, one of these labels must be repeated (twice). The distinguished faces lie opposite the repeated vertices. There are precisely two distinguished faces. 1.197
1. ð(x, y) = â¥x â y⥠⥠0.
2. ð(x, y) = â¥x â y⥠= 0 if and only if x â y = 0, that is x = y. 3. Property (3) ensures that â¥âx⥠= â¥x⥠and therefore â¥x â y⥠= â¥y â x⥠so that ð(x, y) = â¥x â y⥠= â¥y â x⥠= ð(y, x) 4. For any z â ð ð(x, y) = â¥x â y⥠= â¥x â z + z â y⥠†â¥x â z⥠+ â¥z â y⥠= ð(x, z) + ð(z, y) Therefore ð(x, y) = â¥x â y⥠satisï¬es the properties required of a metric. 1.198 Clearly â¥xâ¥â ⥠0 and â¥xâ¥â = 0 if and only if x = 0. Thirdly ð
ð
ð=1
ð=1
â¥ðŒx⥠= max â£ðŒð¥ð ⣠= â£ðŒâ£ max â£ð¥ð ⣠= â£ðŒâ£ â¥x⥠To prove the triangle inequality, we note that for any ð¥ð , ðŠð â â max(ð¥ð + ðŠð ) †max ð¥ð + max ðŠð Therefore ð
ð
ð
ð=1
ð=1
ð=1
â¥x⥠= max(ð¥ð + ðŠð ) †max ð¥ð + max ðŠð = â¥x⥠+ â¥y⥠1.199 Suppose that producing one unit of good 1 requires two units of good 2 and three units of good 3. The production plan is (1, â2, â3) and the average net output, â2, is negative. A norm is required to be nonnegative. Moreover, theâquantities of inputs ð and outputs may balance out yielding a zero average. That is, ( ð=1 ðŠð )/ð = 0 does not imply that ðŠð = 0 for all ð.
49
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 1.200 â¥x⥠â â¥y⥠= â¥x â y + y⥠â â¥y⥠†â¥x â y⥠+ â¥y⥠â â¥y⥠= â¥x â y⥠1.201 Using the previous exercise â¥xð ⥠â â¥x⥠†â¥xð â x⥠â 0
1.202 First note that each term xð + yð â ð by linearity. Similarly, x + y â ð. Fix some ð > 0. There exists some ðx such that â¥xð â x⥠< ð for all ð ⥠ðx . Similarly, there exists some ðy such that â¥yð â y⥠< ð for all ð ⥠ðy . For all ð ⥠max{ ðx , ðy }, â¥(xð + yð ) â (x + y)⥠= â¥(xð â x) + (yð â y)⥠†â¥xð â x⥠+ â¥yð â y⥠0, there exists some ð such that â¥zð â zð ⥠= max{ â¥xð â xð ⥠, â¥yð â yð ⥠} < ð for every ð, ð ⥠ð . This implies that (xð ) and (yð ) are Cauchy sequences in ð and ð respectively. Since ð and ð are complete, both sequences converge. That is, there exists x â ð and y â ð such that â¥xð â x⥠â 0 and â¥yð â y⥠â 0. In other words, given ð > 0 there exists ð such that â¥xð â x⥠< ð and â¥yð â y⥠< ð for every ð ⥠ð . Let z = (x, y). Then, for every ð ⥠ð â¥zð â z⥠= max{ â¥xð â x⥠, â¥yð â y⥠} < ð zð â z. 1.210
1. By assumption, for every ð = 1, 2, . . . , there exists a point yð such that ( ð ) 1 â â¥y⥠< â£ðŒð ⣠ð ð=1 where y = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð Let ð ð =
âð
ð=1
â£ðŒð â£. By assumption ð ð > ð â¥yð ⥠⥠0. Deï¬ne xð =
1 ð y ð ð
Then xð = ðœ1ð x1 + ðœ2ð x2 + â
â
â
+ ðœðð xð âð 1 ð ð ð where ðœðð = ðŒð ð /ð , ð=1 â£ðœð ⣠= 1 and â¥x ⥠< ð for every ð = 1, 2, . . . . ð Consequently â¥x ⥠â 0. 51
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
âð 2. Since ð=1 â£ðœðð ⣠= 1, â£ðœðð ⣠†1 for every ð. Consequently, for every coordinate ð, the sequence (ðœðð ) is bounded. By the Bolzano-Weierstrass theorem (Exercise 1.119), the sequence (ðœ1ð ) has a convergent subsequence with ðœ1ð â ðœ1 . Let xð,1 denote the corresponding subsequence of xð . Similarly, ðœ2ð,1 has a subsequence converging to ðœ2 . Let (xð,2 ) denote the corresponding subsequence of (xð ). Proceeding coordinate by coordinate, we obtain a subsequence (xð,ð ) where each term is xð,ð = ðœ ð,ð x1 + ðœ ð,ð x2 + â
â
â
+ ðœ ð,ð xð and each coeï¬cient converges ðœðð,ð â ðœð . Let x = ðœ1 x1 + ðœ2 x2 + â
â
â
+ ðœ2 xð Then xð,ð â x (Exercise 1.202). âð âð ð 3. Since ð=1 â£ðœð ⣠= 1 for every ð, ð=1 â£ðœð ⣠= 1. Consequently, at least one of the coeï¬cients ðœð â= 0. Since x1 , x2 , . . . , xð are linearly independent, x â= 0 (Exercise 1.133) and therefore â¥x⥠â= 0. But (xð,ð ) is a subsequence of (xð ). This contradicts the earlier conclusion (part 1) that â¥xð ⥠â 0. 1.211
1. Let ð be a normed linear space ð of dimension ð and let { x1 , x2 , . . . , xð } be a basis for ð. Let (xð ) be a Cauchy sequence in ð. Each term xð has a unique representation ð ð xð = ðŒð 1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð
We show that each of the sequences ðŒð ð is a Cauchy sequence in â. Since xð is a Cauchy sequence, for every ð > 0 there exists an ð such that â¥xð â xð ⥠< ð for all ð, ð ⥠ð . Using Lemma 1.1, there exists ð > 0 such that ð ð â â ð ð ð ð ð â£ðŒð â ðŒð ⣠†(ðŒð â ðŒð )xð = â¥xð â xð ⥠< ð ð=1
ð=1
for all ð, ð ⥠ð . Dividing by ð > 0 we get for every ð ð â£ðŒð ð â ðŒð ⣠â€
ð â
ð â£ðŒð ð â ðŒð ⣠<
ð=1
ð ð
ðŒð ð
is a Cauchy sequence in â. Since â is complete, each Thus each sequence sequence converges to some limit ðŒð â â. 2. Let x = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð Then x â ð and
ð ð â â â¥xð â x⥠= (ðŒð â£ðŒð ð â ðŒð )xð †ð â ðŒð ⣠â¥xð ⥠ð=1
ð=1
ð ð Since ðŒð ð â ðŒð for every ð, â¥x â x⥠â 0 which implies that x â x.
3. Since (xð ) was an arbitrary Cauchy sequence, we have shown that every Cauchy sequence in ð converges. Hence ð is complete. 52
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.212 Let ð be an open set according to the â¥â
â¥ð and let x0 be a point in ð. Since ð is open, it contains an open ball in the â¥â
â¥ð topology about x0 , namely ðµð (x0 , ð) = { x â ð : â¥x â x0 â¥ð < ð } â ð Let ðµð (x0 , ð) = { x â ð : â¥x â x0 â¥ð < ð } be the open ball about x0 in the â¥â
â¥ð topology. The condition (1.15) implies that ðµð (x0 , ð) â ðµð (x0 , ð) â ð and therefore x0 â ðµð (x0 , ð) â ð ð is open in the â¥â
â¥ð topology. Similarly, any ð open in the â¥â
â¥ð topology is open in the â¥â
â¥ð topology. 1.213 Let ð be a normed linear space of dimension ð. and let { x1 , x2 , . . . , xð } be a basis for ð. Let â¥â
â¥ð and â¥â
â¥ð be two norms on ð. Every x â ð has a unique representation x = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð Repeated application of the triangle inequality gives â¥xâ¥ð = â¥ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð â¥ð ð â †â¥ðŒð xð â¥ð ð=1
=
ð â
â£ðŒð ⣠â¥xð â¥ð
ð=1 ð â
â€ð
â£ðŒð â£
ð=1
where ð = maxð â¥xð â¥. By Lemma 1.1, there is a positive constant ð such that ð â ð=1
â£ðŒð ⣠†â¥xâ¥ð /ð
Combining these two inequalities, we have â¥xâ¥ð †ð â¥xâ¥ð /ð Setting ðŽ = ð/ð > 0, we have shown ðŽ â¥xâ¥ð †â¥xâ¥ð The other inequality in (1.15) is obtained by interchanging the roles of â¥â
â¥ð and â¥â
â¥ð . 1.214 Assume xð â x = (ð¥1 , ð¥2 , . . . , ð¥ð ). Then, for every ð > 0, there exists some ð such that â¥xð â xâ¥â < ð. Therefore, for ð = 1, 2, . . . , ð â£ð¥ðð â ð¥ð ⣠†max â£ð¥ðð â ð¥ð ⣠= â¥xð â xâ¥â < ð ð
Therefore ð¥ðð â xð .
53
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Conversely, assume that (xð ) is a sequence in âð with ð¥ðð â ð¥ð for every coordinate ð. Choose some ð > 0. For every ð, there exists some integer ðð such that â£ð¥ðð â ð¥ð ⣠< ð for every ð ⥠ðð Let ð = maxð { ð1 , ð2 , . . . , ðð }. Then â£ð¥ðð â ð¥ð ⣠< ð for every ð ⥠ð and â¥xð â xâ¥â = max â£ð¥ðð â ð¥ð ⣠< ð for every ð ⥠ð ð
ð
That is, x â x. A similar proof can be given using the Euclidean norm â¥â
â¥2 , but it is slightly more complicated. This illustrates an instance where the sup norm is more tractable. 1.215
1. Let ð â ð be closed and bounded and let xð be a sequence in ð. Every term xð has a representation ð â
xð =
ðŒð ð xð
ð=1
Since ð is bounded, so is xð . That is, there exists ð such that â¥xð ⥠†ð for all ð. Applying Lemma 1.1, there is a positive constant ð such that ð
ð â
â£ðŒð ⣠†â¥xð ⥠†ð
ð=1
Hence, for every ð, the sequence of scalars ðŒðð is bounded. 2. By the Bolzano-Weierstrass theorem (Exercise 1.119), the sequence ðŒð 1 has a convergent subsequence with limit ðŒ1 . Let ð¥ð (1) be the corresponding subsequence of xð . ð 3. Similarly, ð¥ð (1) has a subsequence for which the corresponding scalars ðŒ2 converge to ðŒ2 . Repeating this process ð times (this is were ï¬niteness is important), we deduce the existence of a subsequence ð¥ð (ð) whose scalars converge to (ðŒ1 , ðŒ2 , . . . , ðŒð ).
4. Let x=
ð â
ðŒð xð
ð=1 ð ð Since ðŒð ð â ðŒð for every ð, â¥x â x⥠â 0 which implies that x â x.
5. Since ð is closed, x â ð. 6. We have shown that every sequence in ð has a subsequence which converges in ð. ð is compact. 1.216 Let x and y belong to ðµ = { x : â¥x⥠< 1 }, the unit ball in the normed linear space ð. Then â¥x⥠, â¥y⥠< 1. By the triangle inequality â¥ðŒx + (1 â ðŒ)y⥠†ðŒ â¥x⥠+ (1 â ðŒ) â¥y⥠†ðŒ + (1 â ðŒ) = 1 Hence ðŒx + (1 â ðŒ)y â ðµ. 54
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.217 If int ð is empty, it is trivially convex. Therefore, assume int ð â= â
and let x, y â int ð. We must show that z = ðŒx + (1 â ðŒ)y â int ð. Since x, y â int ð, there exists some ð > 0 such that the open balls ðµ(x, ð) and ðµ(y, ð) are both contained in int ð. Let w be any vector with â¥w⥠< ð. The point z + w = ðŒ(x + w) + (1 â ðŒ)(y + w) â ð since x + w â ðµ(x, ð) â ð and y + w â ðµ(y, ð) â ð and ð is convex. Hence z is an interior point of ð. Similarly, if ð is empty, it is trivially convex. Therefore, assume ð â= â
and let x, y â ð. Choose some ðŒ. We must show that ð§ = ðŒx + (1 â ðŒ)y â ð. There exist sequences (xð ) and (yð ) in ð which converge to x and y respectively (Exercise 1.105). Since ð is convex, the sequence (ðŒxð + (1 â ðŒ)yð ) lies in ð and moreover converges to ðŒx + (1 â ðŒ)y = z (Exercise 1.202). Therefore ð§ is the limit of a sequence in ð and hence ð§ â ð. Therefore, ð is convex. ¯ = ðŒx1 + (1 â ðŒ)x2 for some ðŒ â (0, 1). Since x1 â ð, 1.218 Let x x1 â ð + ððµ ðŒx1 â ðŒ(ð + ððµ) ¯ of radius ð is The open ball about x ¯ + ððµ ðµ(¯ x, ð) = x = ðŒx1 + (1 â ðŒ)x2 + ððµ â ðŒ(ð + ððµ) + (1 â ðŒ)x2 + ððµ = ðŒð + (1 â ðŒ)x2 + (1 + ðŒ)ððµ ( ) 1+ðŒ = ðŒð + (1 â ðŒ) x2 + ððµ 1âðŒ Since x2 â int ð x2 +
( ) 1+ðŒ 1+ðŒ ððµ = ðµ x2 , ð âð 1âðŒ 1âðŒ
for suï¬ciently small ð. For such ð ðµ(¯ x, ð) â ðŒð + (1 â ðŒ)ð =ð ¯ â int ð. by Exercise 1.168. Therefore x 1.219 It is easy to show that ðâ
â©
ðð
ðâðŒ
To show the converse, choose any x â ð and let x0 â ðð for every ð â ðŒ. By Exercise 1.218, ðŒx + (1 â ðŒ)x0 â ðð for all 0 < ðŒ < 1. This implies that ðŒx + (1 â ðŒ)x0 â â©ðâðŒ ðð = ð for all 0 < ðŒ < 1, and therefore that x0 = limðŒâ0 ðŒx + (1 â ðŒ)x0 â ð. 1.220 Assume that x â int ð. Then, there exists some ð such that ðµ(x, ð) = x + ððµ â ð 55
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Let y be any element in the unit ball ðµ. Then ây â ðµ and x1 = x + ðy â ð x2 = x â ðy â ð so that x=
1 1 x1 + x2 2 2
x is not an extreme point. We have shown that no interior point is an extreme point; hence every extreme point must be a boundary point. 1.221 We showed in Exercise 1.220 that ext(ð) â b(ð). To show the converse, assume that x is a boundary point which is not an extreme point. That is, there exist x1 , x2 â ð such that x = ðŒx1 + (1 â ðŒ)x2
0 0 and ð¥ð < 1 1.229 Let ð = dim ð. By Exercise 1.182, ð contains a simplex ð ð of the same dimension. That is, there exist ð vertices v1 , v2 , . . . , vð such that ð ð = conv { v1 , v2 , . . . , vð } { = ðŒ1 v1 + ðŒ2 v2 + â
â
â
+ ðŒð vð : ðŒ1 , ðŒ2 , . . . , ðŒð ⥠0, ðŒ1 + ðŒ2 + . . . + ðŒð = 1
}
Analogous to the previous part, the relative interior of ð ð is ri ð ð = conv { v1 , v2 , . . . , vð } { = ðŒ1 v1 + ðŒ2 v2 + â
â
â
+ ðŒð vð : ðŒ1 , ðŒ2 , . . . , ðŒð > 0, ðŒ1 + ðŒ2 + . . . + ðŒð = 1
}
which is nonempty. Note, the proposition is trivially true for a set containing a single point (ð = 0), since this point is the whole aï¬ne space. 1.230 If int ð â= â
, then aï¬ ð = ð and ri ð = int ð. The converse follows from Exercise 1.229. 1.231 Since ð > inf
xâð
ð â ð=1
58
ðð ð¥ð
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics there exists some x â ð such that ð â
ðð ð¥ð †ð
ð=1
Therefore x â ð(p, ð) which is nonempty. Let ðË = minð ðð be the lowest price of the ð goods. Then ð(p, ð) â ðµ(0, ð/Ë ð) and so is bounded. (That is, no component of an aï¬ordable bundle can contain more than ð/Ë ð units.) To show that ð(p, ð) is closed, let (xð ) be a sequence of consumption bundles in ð(p, ð). Since ð(p, ð) is bounded, xð â x â ð. Furthermore ð1 ð¥ð1 + ð2 ð¥ð2 + â
â
â
+ ðð ð¥ðð †ð for every ð This implies that ð1 ð¥1 + ð2 ð¥2 + â
â
â
+ ðð ð¥ð †ð ð
so that x â x â ð(p, ð). Therefore ð(p, ð) is closed. We have shown that ð(p, ð) is a closed and bounded subset of âð . Hence it is compact (Proposition 1.4). 1.232 Let x, y â ð(p, ð). That is ð â
ðð ð¥ð †ð
ð=1 ð â
ðð ðŠ ð †ð
ð=1
For any ðŒ â [0, 1], the cost of the weighted average bundle z = ðŒx + (1 â ðŒ)y (where each component ð§ð = ðŒð¥ð + (1 â ðŒ)ðŠð ) is ð â ð=1
ðð ð§ ð =
ð â
ðð (ðŒð¥ð ð=1 ð â
=ðŒ
+ (1 â ðŒ)ðŠð
ðð ð¥ð + (1 â ðŒ)
ð=1
ð â
ðð ðŠ ð
ð=1
†ðŒð + (1 â ðŒ)ð =ð Therefore z â ð(p, ð). The budget set ð(p, ð) is convex. 1.233
1. Assume that â» is strongly monotone. Let x, y â ð with x ⥠y. Either x â© y so that x â» y by strong monotonicity or x = y so that x â¿ y by reï¬exivity. In either case, x â¿ y so that â¿ is weakly monotonic.
2. Again, assume that â¿ is strongly monotonic and let y â ð. ð is open (relative to itself). Therefore, there exists some ð such that ðµ(y, ð) = y + ððµ â ð Let x = y + ðe1 be the consumption bundle containing ð more units of good 1. Then e1 â ðµ, x â ðµ(y, ð) and therefore â¥x â y⥠< ð. Furthermore, x â© y and therefore x â» y. 59
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3. Assume â¿ is locally nonsatiated. Then, for every x â ð, there exists some y â ð such that y â» x. Therefore, there is no best element. 1.234 is assume that xâ â¿ x for every x â ðµ(p, ð) but that âð Assume otherwise, thatâ ð ð=1 ðð ð¥ð < ð. Let ð = ð â ð=1 ðð ð¥ð be the unspent income. Spending the residual on good 1, the commodity bundle x = xâ + ðð1 e1 is aï¬ordable ð â
ðð ð¥ð =
ð=1
ð â
ðð ð¥âð + ð1
ð=1
ð =ð ð1
Moreover, since x â© xâ , x â» xâ , which contradicts the assumption that xâ is the best element in ð(p, ð). 1.235 otherwise, that is assume that xâ â¿ x for every x â ðµ(p, ð) but that âð Assume â â ð=1 ðð ð¥ð < ð. This implies that x â int ð(p, ð). Therefore, there exists a neighâ borhood ð of x with ð â ð(p, ð). Within this neighborhood, there exists some x â ð â ð(p, ð) with x â» xâ , which contradicts the assumption that xâ is the best element in ð(p, ð). 1.236
1. Assume â¿ is continuous. Choose some y â ð. For any x0 in â»(y), x0 â» y and (since â¿ is continuous) there exists some neighborhood ð(x0 ) such that x â» y for every x â ð(x0 ). That is, ð(x0 ) â â»(y) and â»(y) is open. Similarly, for any x0 â âº(y), x0 ⺠y and there exists some neighborhood ð(x0 ) such that x ⺠y for every x â ð(x0 ). Thus ð(x0 ) â âº(y) and âº(y) is open.
2. Conversely, assume that the sets â»(y) = { x : x â» y } and âº(y) = { x : x ⺠y } are open in x. Assume x0 â» y0 . (a) Suppose there exists some y such that x0 â» y â» z0 . Then x0 â â»(y), which is open by assumption. That is, â»(y) is an open neighborhood of x0 and x â» y for every x â â»(y). Similarly, âº(y) is an open neighborhood of z0 for which y â» z for every z â âº(y). Therefore ð(x0 ) = â»(y) and ð(z0 ) = âº(y) are the required neighborhoods of x0 and z0 respectively such that xâ»yâ»z
for every x â ð(x0 ) and y â ð(z0 )
(b) Suppose there is no y such that x0 â» y â» z0 . i. By assumption â â»(z0 ) is open â x0 â» z0 which implies x0 â â»(z0 ), Therefore â»(z0 ) is an open neighborhood of x0 . ii. Since â¿ is complete, either y ⺠x0 or y â¿ x0 for every y â ð (Exercise 1.56. Since there is no y such that x0 â» y â» z0 y â» z0 =â y â⺠x0 =â y â¿ x0 Therefore â»(z0 ) = â¿(x0 ). iii. Since x â¿ x0 â» z0 for every x â â¿(x0 ) = â»(z0 ) x â» z0 for every x â â»(z0 )
60
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
iv. Therefore ð(x0 ) = â»(z0 ) is an open neighborhood of x0 such that x â» z0 for every x â ð(x0 ) Similarly, ð(z0 ) = âº(x0 ) is an open neighborhood of z0 such that z ⺠x0 for every z â ð(z0 ). Consequently xâ»z
for every x â ð(x0 ) and z â ð(z0 )
( )ð 3. â¿(y) = âº(y) (Exercise 1.56). Therefore, â¿(y) is closed if and only if âº(y) is open (Exercise 1.80). Similarly, âŸ(y) is closed if and only if â»(y) is open. 1.237
1. Let ð¹ = { (x, y) â ð Ãð : x â¿ y }. Let ((xð , yð )) be a sequence in ð¹ which converges to (x, y). Since (xð , yð ) â ð¹ , xð â¿ yð for every ð. By assumption, x â¿ y. Therefore, (x, y) â ð¹ which establishes that ð¹ is closed (Exercise 1.106) Conversely, assume that ð¹ is closed and let ((xð , yð )) be a sequence converging to (x, y) with xð â¿ yð for every ð. Then ((xð , yð )) â ð¹ which implies that (x, y) â ð¹ . Therefore x â¿ y.
2. Yes. Setting yð = y for every ð, their deï¬nition implies that for every sequence (xð ) in ð with xð â¿ y, x = lim xð â¿ y. That is, the upper contour set â¿(y) = { x : x â¿ y } is closed. Similarly, the lower contour set âŸ(y) is closed. Conversely, assume that the preference relation is continuous (in our deï¬nition). We show that the set ðº = { (x, y) : x ⺠y } is open. Let (x0 , y0 ) â ðº. Then x0 ⺠y0 . By continuity, there exists neighborhoods ð(x0 ) and ð(y0 ) of x0 and y0 such that x ⺠y for every x â ð(x0 ) and y â ð(y0 ). Hence, for every (x, y) â ð = ð(x0 ) à ð(y0 ), x ⺠y. Therefore ð â ðº which implies that ðº is open. Consequently ðºð = { (x, y) : x â¿ y } is closed. 1.238 Assume the contrary. That is, assume there is no y with x â» y â» z. Since â¿ is complete, either y ⺠x0 or y â¿ x0 for every y â ð (Exercise 1.56). Since there is no y such that x0 â» y â» z0 y â» z0 =â y â⺠x0 =â y â¿ x0 Therefore â»(z0 ) = â¿(x0 ). By continuity, â»(z0 ) is open and â¿(x0 ) is closed. Hence â»(z0 ) = â¿(x0 ) is both open and closed (Exercise 1.83). Alternatively, â¿(x0 ) and âŸ(z0 ) are both open sets which partition ð. This contradicts the assumption that ð is connected. 1.239 Let ð â denote the set of best elements. As demonstrated in the preceding proof â© ðâ = â¿(yð ) yâð
Therefore ð â is closed (Exercise 1.85) and hence compact (Exercise 1.110). 1.240 Assume for simplicity that ð1 = ð2 = 1 and that ð = 1. Then, the budget set is ðµ(1, 1) = { x â â2++ : ð¥1 + ð¥2 †1 } The consumer would like to spend as much as possible of her income on good 1. However, the point (1, 0) is not feasible, since (1, 0) â / ð.
61
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.241 Essentially, consumer theory (in economics) is concerned with predicting the way in which consumer purchases vary with changes in observable parameters such as prices and incomes. Predictions are deduced by assuming that the consumer will consistently choose the best aï¬ordable alternative in her budget set. The theory would be empty if there was no such optimal choice. 1.242
1. Let ð 0 = ð â© âð+ . Then ð 0 is compact and ð 1 â ð 0 . Deï¬ne the order x â¿1 y if and only if ð1 (x) †ð1 (y). Then â¿1 is continuous on ð and ð 1 = { x â ð : ð1 (x) †ð1 (y) for every y â ð } is the set of best elements in ð with respect to the order â¿1 . By Exercise 1.239, ð 1 is nonempty and compact.
2. Assume ð ðâ1 is compact. Deï¬ne the order x â¿ð y if and only if ðð (x) †ðð (y). Then â¿ð is continuous on ð ðâ1 and ð ð = { x â ð ðâ1 : ðð (x) †ðð (y) for every y â ð ðâ1 } is the set of best elements in ð ðâ1 with respect to the order â¿ð . By Exercise 1.239, ð ð is nonempty and compact. 3. Assume x â Nu. Then x â¿ y for every y â ð d(x) âŸð¿ d(y) for every y â ð For every ð = 1, 2, . . . , 2ð dð (x) †dð (y) for every y â ð ð
ð
which implies x â ð ð . In particular x â ð 2 . Therefore Nu â ð 2 . ð
ð
ð
Suppose Nu â ð 2 . Then there exists some x, y â ð, x â / ð 2 and y â ð 2 such ð / ð ð . Then ðð (x) > ðð (y). that x â¿ y. Let ð be the smallest integer such that x â ð But x â ð for every ð < ð and therefore ðð (x) = ðð (y) for ð = 1, 2, . . . , ð â 1. This means that d(y) âºð¿ d(x) so that x âºð y. This contradiction establishes ð that Nu = ð 2 . 1.243 Assume â¿ is convex. Choose any y â ð and let x â â¿(y). Then x â¿ y. Since â¿ is convex, this implies that ðŒx + (1 â ðŒ)y â¿ y
for every 0 †ðŒ †1
and therefore ðŒx + (1 â ðŒ)y â â¿(y)
for every 0 †ðŒ †1
Therefore â¿(y) is convex. To show the converse, assume that â¿(y) is convex for every y â ð. Choose x, y â ð. Interchanging x and y if necessary, we can assume that x â¿ y so that x â â¿(y). Of course, y â â¿(y). Since â¿(y) is convex ðŒx + (1 â ðŒ)y â â¿(y)
for every 0 †ðŒ †1
which implies ðŒx + (1 â ðŒ)y â¿ y
for every 0 †ðŒ †1 62
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.244 ð â may be empty, in which case it is trivially convex. Otherwise, let xâ â ð â . For every x â ð â x â¿ xâ which implies x â â¿(xâ ) Therefore ð â â â¿(xâ ). Conversely, by transitivity x â¿ xâ â¿ y for every y â ð for every x â â¿(xâ ) which implies â¿(xâ ) â ð â . Therefore, ð â = â¿(xâ ) which is convex. 1.245 To show that â¿ð is strictly convex, assume that x, y â ð are such d(x) = d(y) with x â= y. Suppose ) ( d(x) = ð(ð1 , x), ð(ð2 , x) . . . , ð(ð2ð , x) In the order ð1 , ð2 , . . . , ð2ð , let ðð be the ï¬rst coalition for which ð(ðð , x) â= ð(ðð , y). That is ð(ðð , x) = ð(ðð , y) for every ð < ð
(1.32)
Since ð(ðð , x) â= ð(ðð , y) and d(x) is listed in descending order, we must have ð(ðð , x) > ð(ðð , y)
(1.33)
ð(ðð , x) ⥠ð(ðð , y) for every ð > ð
(1.34)
and
Choose 0 < ðŒ < 1 and let z = ðŒx + (1 â ðŒ)y. For any coalition ð â ð(ð, z) = ð€(ð) â ð§ð ðâð
â( ) = ð€(ð) â ðŒð¥ð + (1 â ðŒ)ðŠð ðâð
= ð€(ð) â ðŒ (
â
ð¥ð â (1 â ðŒ)
ðâð
= ðŒ ð€(ð) â
â
ðâð
) ð¥ð
â
ðŠð (
+ (1 â ðŒ) ð€(ð) â
ðâð
â
) ðŠð
ðâð
= ðŒð(ð, x) + (1 â ðŒ)ð(ð, y) Using (1.55) to (1.57), this implies that ð(ðð , z) = ð(ðð , x),
ðð
for every 0 < ðŒ < 1, Therefore d(z) âºð¿ d(x). Thus z â»ð x, which establishes that â¿ is strictly convex. The set of feasible outcomes is convex. Assume x, y â Nu â ð, x â= y. Then d(x) = d(y) and z = ðŒx + (1 â ðŒ)y â»ð x for every 0 < ðŒ < 1 which contradicts the assumption that x â Nu. We conclude that the nucleolus contains only one element. 63
Solutions for Foundations of Mathematical Economics 1.246
c 2001 Michael Carter â All rights reserved
1. (a) Clearly âº(x0 ) â âŸ(x0 ) and â»(y0 ) â â¿(y0 ). Consequently âº(x0 ) ⪠â»(y0 ) â âŸ(x0 ) ⪠â¿(y0 ). We claim that these sets are in fact equal. Let z â âŸ(x0 ) ⪠â¿(y0 ). Suppose that z â âŸ(x0 ) but z â / âº(x0 ). Then z â¿ x0 . By transitivity, z â¿ x0 â» y0 which implies that z â â»(y0 ). Similarly z â â¿(y0 ) â â»(y0 ) implies z â âº(x0 ). Therefore âº(x0 ) ⪠â»(y0 ) = âŸ(x0 ) ⪠â¿(y0 ) (b) By continuity, âº(x0 ) ⪠â»(y0 ) is open and âŸ(x0 ) ⪠â¿(y0 ) = âº(x0 ) ⪠â»(y0 ) is closed. Further x0 â» y0 implies that x0 â â»(y0 ) so that âº(x0 ) ⪠â»(y0 ) â= â
. We have established that âº(x0 ) ⪠â»(y0 ) is a nonempty subset of ð which is both open and closed. Since ð is connected, this implies (Exercise 1.83) that âº(x0 ) ⪠â»(y0 ) = ð
2. (a) By deï¬nition, x â / âº(x). So âº(x) â© âº(y) = ð implies x â â»(y), that is x â¿ y contradicting the noncomparability of x and y. Therefore âº(x) â© âº(y) â= ð (b) By assumption, there exists at least one pair x0 , y0 such that x0 â» y0 . By the previous part âº(x0 ) ⪠â»(y0 ) = ð This implies either x ⺠x0 or x â» y0 . Without loss of generality, assume x â» y0 . Again using the previous part, we have âº(x) ⪠â»(y0 ) = ð Since x and y are not comparable, y â / âº(x) which implies that y â â»(y0 ). Therefore x â» y0 and y â» y0 or alternatively y0 â âº(x0 ) â© â»(y0 ) â= â
(c) Clearly âº(x) â âŸ(x) and â»(y) â â¿(y). Consequently âº(x) â© âº(y) â âŸ(x) â© âŸ(y) Let z â âŸ(x) â© âŸ(y). That is, z ⟠x. If x ⟠z, then transitivity implies x ⟠z ⟠y, which contradicts the noncomparability of x and y. Consequently x â⟠z which implies z ⺠x and z â âº(x). Similarly z â âº(y) and therefore âº(x) â© âº(y) = âŸ(x) â© âŸ(y) 3. If x and y are noncomparable, âº(x)â©âº(y) is a nonempty proper subset of ð. By continuity âº(x) â© âº(y) = âŸ(x) â© âŸ(y) is both open and closed which contradicts the assumption that ð is connected (Exercise 1.83). We conclude that â¿ must be complete. 1.247 Assume x â» y. Then x â â»(y). Since â»(y) is open, x â int â»(y). Also y â â»(y). By Exercise 1.218, ðŒx + (1 â ðŒ)y â int â»(y) for every 0 < ðŒ < 1, which implies ðŒx + (1 â ðŒ)y â» y for every 0 < ðŒ < 1 64
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
1.248 For every x â ð, there exists some z such that z â» x (Nonsatiation). For any ð, choose some ðŒ â (0, ð/ â¥x â zâ¥) and let y = ðŒz + (1 â ðŒ)x. Then â¥x â y⥠= ðŒ â¥x â z⥠< ð Moreover, since â¿ is strictly convex, y = ðŒz + (1 â ðŒ)x â» x Thus, â¿ is locally nonsatiated. We have previously shown that local nonsatiation implies nonsatiation (Exercise 1.233). Consequently, these two properties are equivalent for strictly convex preferences. 1.249 Assume that x is not strongly Pareto eï¬cient. That is, there exist allocation y such that y â¿ð x for all ð and some individual ð for which y â»ð x. Take 1 â ð¡ percent of ðâs consumption and distribute it equally to the other participants. By continuity, 1âð¡ there exists some ð¡ such that ð¡y â»ð x. The other agents receive yð + ðâ1 yð which, by monotonicity, they strictly prefer to xð . 1.250 Assume that (pâ , xâ ) is a competitive equilibrium of an exchange economy, but that xâ does not belong to the core of the corresponding market game. Then there exists some coalition ð and allocation y ââ ð (ð) such that yð â»ð xâð for every ð â ð. â Since y â ð (ð), we must have ðâð yð = ðâð wð . Since xâ is a competitive equilibrium and yð â»ð xâð for every ð â ð, y must be unaffordable, that is ð â
ðð ðŠðð >
ð=1
ð â
ðð wðð for every ð â ð
ð=1
and therefore ð ââ ðâð ð=1
ðð ðŠðð >
ð ââ
ðð wðð
ðâð ð=1
which contradicts the assumption that ðŠ â ð (ð). 1.251 Combining the previous exercise with Exercise 1.64 xâ â core â Pareto
65
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 2: Functions 2.1 In general, the birthday mapping is not one-to-one since two individuals may have the same birthday. It is not onto since some days may be no oneâs birthday. 2.2 The origin 0 is ï¬xed point for every ð. Furthermore, when ð = 0, ð is an identity function and every point is a ï¬xed point. 2.3 For every ð¥ â ð, there exists some ðŠ â ð such that ð (ð¥) = ðŠ, whence ð¥ â ð â1 (ðŠ). Therefore, every ð¥ belongs to some contour. To show that distinct contours are disjoint, assume ð¥ â ð â1 (ðŠ1 ) â© ð â1 (ðŠ2 ). Then ð (ð¥) = ðŠ1 and also ð (ð¥) = ðŠ2 . Since ð is a function, this implies that ðŠ1 = ðŠ2 . 2.4 Assume ð is one-to-one and onto. Then, for every ðŠ â ð , there exists ð¥ â ð such that ð (ð¥) = ðŠ. That is, ð â1 (ðŠ) â= â
for every ðŠ â ð . If ð is one to one, ð (ð¥) = ðŠ = ð (ð¥â² ) implies ð¥ = ð¥â² . Therefore, ð â1 (ðŠ) consists of a single element. Therefore, the inverse function ð â1 exists. Conversely, assume that ð : ð â ð has an inverse ð â1 . As ð â1 is a function mapping ð to ð, it must be deï¬ned for every ðŠ â ð . Therefore ð is onto. Assume there exists ð¥, ð¥â² â ð and ðŠ â ð such that ð (ð¥) = ðŠ = ð (ð¥â² ). Then ð â1 (ðŠ) = ð¥ and also ð â1 (ðŠ) = ð¥â² . Since ð â1 is a function, this implies that ð¥ = ð¥â² . Therefore ð is one-to-one. 2.5 Choose any ð¥ â ð and let ðŠ = ð (ð¥). Since ð is one-to-one, ð¥ = ð â1 (ðŠ) = ð â1 (ð (ð¥)). The second identity is proved similarly. 2.6 (2.2) implies for every ð¥ â â ðð¥ ðâð¥ = ð0 = 1 and therefore ðâð¥ =
1 ðð¥
For every ð¥ ⥠0 ðð¥ = 1 +
ð¥3 ð¥ ð¥2 + + + â
â
â
> 0 1 2 6
and therefore by (2.28) ðð¥ > 0 for every ð¥ â â. For every ð¥ ⥠1 ðð¥ = 1 +
ð¥3 ð¥ ð¥2 + + + â
â
â
⥠1 + ð¥ â â as ð¥ â â 1 2 6
and therefore ðð¥ â â as ð¥ â â. By (2.28) ðð¥ â 0 as ð¥ â ââ. 2.7 ðð¥/2 ðð¥/2 ðð¥ = ð¥ 2 ð¥2 ( ð¥/2 ) 1 ð = ðð¥/2 â â as ð¥ â â 2 ð¥/2 66
(2.28)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
since the term in brackets is strictly greater than 1 for any ð¥ > 0. Similarly ðð¥ (ðð¥/(ð+1) )ð ðð¥/(ð+1) = ð¥ ð ð¥ (ð + 1)ð ( ð+1 ) ( ð¥/(ð+1) )ð 1 ð = ðð¥/(ð+1) â â (ð + 1)ð ð¥/(ð + 1) 2.8 Assume that ð â â is compact. Then ð is bounded (Proposition 1.1), and there exists ð such that â£ð¥â£ †ð for every ð¥ â ð. For all ð ⥠ð ⥠2ð ð â ð¥ð ð¥ð+1 ðâð â ( ð¥ )ð â£ðð (ð¥) â ðð (ð¥)⣠= †ð! (ð + 1)! ð ð=ð+1 ð=0 ð ð+1 ðâð â ( ð )ð †(ð + 1)! ð ð=0 ( ( )ðâð ) ð ð+1 1 1 1 †1 + + + â
â
â
+ (ð + 1)! 2 4 2 ( ( )ð ) ð+1 ð 1 ð ð+1 â€2 â€2 †(ð + 1)! ð+1 2 by Exercise 1.206. Therefore ðð converges to ð for all ð¥ â ð. 2.9 This is a special case of Example 2.8. For any ð, ð â ð¹ (ð), deï¬ne (ð + ð) = ð (ð¥) + ð(ð¥) (ðŒð )(ð¥) = ðŒð (ð¥) With these deï¬nitions ð + ð and ðŒð also map ð to â. Hence ð¹ (ð) is closed under addition and scalar multiplication. It is straightforward but tedious to verify that ð¹ (ð) satisï¬es the other requirements of a linear space. 2.10 The zero element in ð¹ (ð) is the constant function ð (ð¥) = 0 for every ð¥ â ð. 2.11
1. From the deï¬nition of â¥ð ⥠it is clear that â â¥ð ⥠⥠0. â â¥ð ⥠= 0 of and only ð is the zero functional. â â¥ðŒð ⥠= â£ðŒâ£ â¥ð ⥠since supð¥âð â£ðŒð (ð¥)⣠= â£ðŒâ£ supð¥âð â£ð (ð¥)⣠It remains to verify the triangle inequality, namely â¥ð + ð⥠= sup â£(ð + ð)(ð¥)⣠ð¥âð
= sup â£ð (ð¥) + ð(ð¥)⣠ð¥âð { } †sup â£ð (ð¥)⣠+ â£ð(ð¥)⣠ð¥âð
†sup â£(ð (ð¥)⣠+ sup â£ð(ð¥)⣠ð¥âð
ð¥âð
= â¥ð ⥠+ â¥ð⥠2. Consequently, for any ð â ðµ(ð), ðŒð (ð¥) †â£ðŒâ£ â¥ð ⥠for every ð¥ â ð and therefore ðŒð â ðµ(ð). Similarly, for any ð, ð â ðµ(ð), (ð + ð)(ð¥) †â¥ð ⥠+ â¥ð⥠for every 67
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
ð¥ â ð and therefore ð + ð â ðµ(ð). Hence, ðµ(ð) is closed under addition and scalar multiplication; it is a subspace of the linear space ð¹ (ð). We conclude that ðµ(ð) is a normed linear space. 3. To show that ðµ(ð) is complete, assume that (ð ð ) is a Cauchy sequence in ðµ(ð). For every ð¥ â ð â£ð ð (ð¥) â ð ð (ð¥)⣠†â¥ð ð â ð ð ⥠â 0 Therefore, for ð¥ â ð, ð ð (ð¥) is a Cauchy sequence of real numbers. Since â is complete, this sequence converges. Deï¬ne the function ð (ð¥) = lim ð ð (ð¥) ðââ
We need to show â â¥ð ð â ð ⥠â 0 and â ð â ðµ(ð) ð
(ð ) is a Cauchy sequence. For given ð > 0, choose ð such that â¥ð ð â ð ð ⥠< ð/2 for very ð, ð ⥠ð . For any ð¥ â ð and ð ⥠ð , â£ð ð (ð¥) â ð (ð¥)⣠†â£ð ð (ð¥) â ð ð (ð¥)⣠+ â£ð ð (ð¥) â ð (ð¥)⣠†â¥ð ð â ð ð ⥠+ â£ð ð (ð¥) â ð (ð¥)⣠By suitable choice of ð (which may depend upon ð¥), each term on the right can be made smaller than ð/2 and therefore â£ð ð (ð¥) â ð (ð¥)⣠< ð for every ð¥ â ð and ð ⥠ð . â¥ð ð â ð ⥠= sup â£ð ð (ð¥) â ð (ð¥)⣠†ð ð¥âð
Finally, this implies â¥ð ⥠= limðââ â¥ð ð â¥. Therefore ð â ðµ(ð). 2.12 If the die is fair, the probability of the elementary outcomes is ð ({1}) = ð ({2}) = ð ({3}) = ð ({4}) = ð ({5}) = ð ({6}) = 1/6 By Condition 3 ð ({2, 4, 6}) = ð ({2}) + ð ({4}) + ð ({6}) = 1/2 2.13 The proï¬t maximization problem of a competitive single-output ï¬rm is to choose the combination of inputs x â âð+ and scale of production ðŠ to maximize net proï¬t. This is summarized in the constrained maximization problem max ððŠ â x,ðŠ
âð
ð â
ð€ð ð¥ð
ð=1
subject to x â ð (ðŠ)
where ððŠ is total revenue and ð=1 ð€ð ð¥ð total cost. The proï¬t function, which depends upon both ð and w, is deï¬ned by Î (ð, w) =
max ððŠ â
ðŠ,xâð (ðŠ)
68
ð â ð=1
ð€ð ð¥ð
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
For analysis, it is convenient to represent the technology ð (ðŠ) by a production function (Example 2.24). The ï¬rmâs optimization can then be expressed as maxð ðð (x) â
xââ+
ð â
ð€ð ð¥ð
ð=1
and the proï¬t function as Î (ð, w) = maxð ðð (x) â xââ+
2.14
ð â
ð€ð ð¥ð
ð=1
1. Assume that production is proï¬table at p. That is, there exists some y â ð such that ð (y, p) > 0. Since the technology exhibits constant returns to scale, ð is a cone (Example 1.101). Therefore ðŒy â ð for every ðŒ > 0 and â â ðð (ðŒðŠð ) = ðŒ ðð ðŠð = ðŒð (y, p) ð (ðŒy, p) = ð
ð
Therefore { ð (ðŒy, p) : ðŒ > 0 } is unbounded and Î (p) = sup ð (y, p) ⥠sup ð (ðŒy, p) = +â ðŒ>0
yâð
2. Assume to the contrary that there exists p â âð+ with Î (p) = ð â / { 0, +â, ââ }. There are two possible cases. (a) 0 < ð < +â. Since ð = supðŠâð ð (y, p) > 0, there exists y â ð such that ð (y, p) > 0 The previous part implies Î (p) = +â. (b) ââ < ð < 0. Then there exists y â ð such that ð (y, p) < 0 By a similar argument to the previous part, this implies Î (p) = ââ. 2.15 Assume xâ is a solution to (2.4). ð (xâ , ðœ) ⥠ð (x, ðœ) for every x â ðº(ðœ) and therefore ð (xâ , ðœ) ⥠sup ð (x, ðœ) = ð£(ðœ) xâðº(ðœ)
On the other hand xâ â ðº(ðœ) and therefore ð£(ðœ) = sup ð (x, ðœ) ⥠ð (xâ , ðœ) xâðº(ðœ)
Therefore, xâ satisï¬es (2.5). Conversely, assume xâ â ðº(ðœ) satisï¬es (2.5). Then ð (xâ , ðœ) = ð£(ðœ) = sup ð (x, ðœ) ⥠ð (x, ðœ) for every x â ðº(ðœ) xâðº(ðœ)
xâ solve (2.4). 2.16 The assumption that ðº(ð¥) â= â
for every ð¥ â ð implies Î(ð¥0 ) â= â
for every ð¥0 â ð. There always exist feasible plans from any starting point. Since ð¢ is bounded, there exists ð such that â£ð (ð¥ð¡ , ð¥ð¡+1 )⣠†ð for every x â Î(ð¥0 ). Consequently, for every x â Î(ð¥0 ), ð (x) â â and â â â â â â ð â£ð (x)⣠= ðœ ð¡ ð (ð¥ð¡ , ð¥ð¡+1 ) †ðœ ð¡ â£ð (ð¥ð¡ , ð¥ð¡+1 )⣠†ðœð¡ð = 1âðœ ð¡=0 ð¡=0 ð¡=0 69
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
using the formula for a geometric series (Exercise 1.108). Therefore ð£(ð¥0 ) =
sup ð (x) â€
xâÎ(ð¥0 )
ð 1âðœ
and ð£ â ðµ(ð). Next, we note that for every feasible plan x â Î(ð¥0 ) ð (x) =
â â
ðœ ð¡ ð (ð¥ð¡ , ð¥ð¡+1 )
ð¡=0
= ð (ð¥0 , ð¥1 ) + ðœ
â â
ðœ ð¡ ð (ð¥ð¡+1 , ð¥ð¡+2 )
ð¡=0
= ð (ð¥0 , ð¥1 ) + ðœð (xâ² )
(2.29)
where xâ² = (ð¥1 , ð¥2 , . . . ) is the continuation of the plan x starting at ð¥1 . For any ð¥0 â ð and ð > 0, there exists a feasible plan x â Î(ð¥0 ) such that ð (x) ⥠ð£(ð¥0 ) â ð Let xâ² = (ð¥1 , ð¥2 , . . . ) be the continuation of the plan x starting at ð¥1 . Using (2.29) and the fact that ð (xâ² ) †ð£(ð¥1 ), we conclude that ð£(ð¥0 ) â ð †ð (x) = ð (ð¥0 , ð¥1 ) + ðœð (xâ² ) †ð (ð¥0 , ð¥1 ) + ðœð£(ð¥1 ) †sup { ð (ð¥0 , ðŠ) + ðœð£(ðŠ) } ðŠâðº(ð¥)
Since this is true for every ð > 0, we must have ð£(ð¥0 ) †sup { ð (ð¥0 , ðŠ) + ðœð£(ðŠ) } ðŠâðº(ð¥)
On the other hand, choose any ð¥1 â ðº(ð¥0 ) â ð. Since ð£(ð¥1 ) =
sup ð (x)
xâÎ(ð¥1 )
there exists a feasible plan xâ² = (ð¥1 , ð¥2 , . . . ) starting at ð¥1 such that ð (xâ² ) ⥠ð£(ð¥1 ) â ð Moreover, the plan x = (ð¥0 , ð¥1 , ð¥2 , . . . ) is feasible from ð¥0 and ð£(ð¥0 ) ⥠ð (x) = ð (ð¥0 , ð¥1 ) + ðœð (xâ² ) ⥠ð (ð¥0 , ð¥1 ) + ðœð£(ð¥1 ) â ðœð Since this is true for every ð > 0 and ð¥1 â ðº(ð¥0 ), we conclude that ð£(ð¥0 ) ⥠sup { ð (ð¥0 , ðŠ) + ðœð£(ðŠ) } ðŠâðº(ð¥)
Together with (2.30) this establishes the required result, namely ð£(ð¥0 ) = sup { ð (ð¥0 , ðŠ) + ðœð£(ðŠ) } ðŠâðº(ð¥)
70
(2.30)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.17 Assume x is optimal, so that ð (xâ ) ⥠ð (x) for every x â Î(ð¥0 ) This implies (using (2.39)) ð (ð¥0 , ð¥â1 ) + ðœð (xâ â² ) ⥠ð (ð¥0 , ð¥1 ) + ðœð (xâ² ) where xâ² = (ð¥1 , ð¥2 , . . . ) is the continuation of the plan x starting at ð¥1 and xâ â² = (ð¥â1 , ð¥â2 , . . . ) is the continuation of the plan xâ . In particular, this is true for every plan x â Î(ð¥0 ) with ð¥1 = ð¥â1 and therefore ð (ð¥0 , ð¥â1 ) + ðœð (xâ â² ) ⥠ð (ð¥0 , ð¥â1 ) + ðœð (xâ² ) for every xâ² â Î(ð¥â1 ) which implies that ð (xâ â² ) ⥠ð (xâ² ) for every xâ² â Î(ð¥â1 ) That is, xâ â² is optimal starting at ð¥â1 and therefore ð (xâ â² ) = ð£(ð¥â1 ) (Exercise 2.15). Consequently ð£(ð¥0 ) = ð (xâ ) = ð (ð¥0 , ð¥â1 ) + ðœð (xâ â² ) = ð (ð¥0 , ð¥â1 ) + ðœð£(ð¥â1 ) This veriï¬es (2.13) for ð¡ = 0. A similar argument veriï¬es (2.13) for any period ð¡. To show the converse, assume that xâ = (ð¥0 , ð¥â1 , ð¥â2 , . . . ) â Î(ð¥0 ) satisï¬es (2.13). Successively using (2.13) ð£(ð¥0 ) = ð (ð¥0 , ð¥â1 ) + ðœð£(ð¥â1 ) = ð (ð¥0 , ð¥â1 ) + ðœð (ð¥â1 , ð¥â2 ) + ðœ 2 ð£(ð¥â1 ) =
1 â
ðœ ð¡ ð (ð¥âð¡ , ð¥âð¡+1 ) + ðœ 2 ð£(ð¥â2 )
ð¡=0
=
2 â
ðœ ð¡ ð (ð¥âð¡ , ð¥âð¡+1 ) + ðœ 3 ð£(ð¥â3 )
ð¡=0
=
ð â1 â
ðœ ð¡ ð (ð¥ð¡ , ð¥ð¡ + 1) + ðœ ð ð£(ð¥âð )
ð¡=0
for any ð = 1, 2, . . . . Since ð£ is bounded (Exercise 2.16), ðœ ð ð£(ð¥âð ) â 0 as ð â â and therefore ð£(ð¥0 ) =
â â
ðœ ð¡ ð (ð¥ð¡ , ð¥ð¡+1 ) = ð (xâ )
ð¡=0
Again using Exercise 2.15, xâ is optimal. 2.18 We have to show that â for any ð£ â ðµ(ð), ð ð£ is a functional on ð. â ð ð£ is bounded. Since ð¹ â ðµ(ð à ð), there exists ð1 < â such that â£ð (ð¥, ðŠ)⣠†ð1 for every (ð¥, ðŠ) â ð à ð. Similarly, for any ð£ â ðµ(ð), there exists ð2 < â such that â£ð£(ð¥)⣠†ð2 for every ð¥ â ð. Consequently for every (ð¥, ðŠ) â ð à ð and ð£ â ðµ(ð) â£ð (ð¥, ðŠ) + ðœð£(ðŠ)⣠†â£ð (ð¥, ðŠ)⣠+ ðœ â£ð£(ðŠ)⣠†ð1 + ðœð2 < â 71
(2.31)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics For each ð¥ â ð, the set ðð¥ =
{
ð (ð¥, ðŠ) + ðœð£(ðŠ) : ðŠ â ðº(ð¥)
}
is a nonempty bounded subset of â, which has least upper bound. Therefore (ð ð£)(ð¥) = sup ðð¥ = sup ð (ð¥, ðŠ) + ðœð£(ðŠ) ðŠâðº(ð¥)
deï¬nes a functional on ð. Moreover by (2.31) â£(ð ð£)(ð)⣠†ð1 + ðœð2 < â Therefore ð ð£ â ðµ(ð). 2.19 Let ð = {1, 2, 3}. Any individual is powerless so that ð€({ð}) = 0
ð = 1, 2, 3
Any two players can allocate the $1 to between themselves, leaving the other player out. Therefore ð€({ð, ð}) = 1
ð, ð â ð, ð â= ð
The best that the three players can achieve is to allocate the $1 amongst themselves, so that ð€(ð ) = 1 2.20 If the consumers preferences are continuous and strictly convex, she has a unique optimal choice xâ for every set of prices p and income ð in ð (Example 1.116). Therefore, the demand correspondence is single valued. 2.21 Assume ð âð â ðµ(sâ ) for every ð â ð . Then for every player ð â ð (ð ð , sâð ) â¿ð (ð â²ð , sâð ) for every ð â²ð â ðð sâ = (ð â1 , ð â2 , . . . , ð âð ) is a Nash equilibrium. Conversely, assume sâ = (ð â1 , ð â2 , . . . , ð âð ) is a Nash equilibrium. Then for every player ð â ð (ð ð , sâð ) â¿ð (ð â²ð , sâð ) for every ð â²ð â ðð which implies that ð âð â ðµ(sâ ) for every ð â ð 2.22 For any nonempty compact set ð â ð, ðµ(ð ) is nonempty and compact provided â¿ð is continuous (Proposition 1.5) and ðµ(ð ) â ð . Therefore ðµð1 â ðµð2 â ðµð3 . . . is a nested sequence ofâ©nonempty compact sets. By the nested intersection theorem â (Exercise 1.117), ð
ð = ð=0 ðµðð â= â
. 2.23 If sâ is a Nash equilibrium, ð ð â ðµðð for every ð.
72
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.24 For any ðœ, let xâ â ð(ðœ). Then ð (xâ , ðœ) ⥠ð (x, ðœ)
for every x â ðº(ðœ)
Therefore ð (xâ , ðœ) ⥠ð£(ðœ) = sup ð (x, ðœ) xâðº(ðœ)
Conversely ð£(ðœ) = sup ð (x, ðœ) ⥠sup ð (x, ðœ) ⥠ð (xâ , ðœ) for every xâ â ð(ðœ) xâðº(ðœ)
xâðº(ðœ)
Consequently ð£(ðœ) = ð (xâ , ðœ) for any xâ â ð(ðœ) 2.25 The graph of ð is graph(ð ) = { (ðŠ, x) â â+ à âð+ : x â ð (ðŠ) } while the production possibility set ð is ð = { (ðŠ, âx) â â+ à âð+ : ð¥ â ð (ðŠ) } Assume that ð is convex and let (ðŠ ð , xð ) â graph(ð ), ð = 1, 2. This means that (ðŠ 1 , âx1 ) â ð and (ðŠ 2 , âx2 ) â ð Let ¯ = ðŒx1 + (1 â ðŒ)x2 ðŠÂ¯ = ðŒðŠ 1 + (1 â ðŒ)ðŠ 2 and x for some 0 †ðŒ †1. Since ð is convex (¯ ðŠ , ⯠x) = ðŒ(ðŠ 1 , âx1 ) + (1 â ðŒ)(ðŠ 2 , âx2 ) â ð ¯ â ð (¯ ¯ ) â graph(ð ). That is graph(ð ) is convex. and therefore x ðŠ ) so that (¯ ðŠ, x Conversely, assuming graph(ð ) is convex, if (ðŠ ð , âxð ) â ð , ð = 1, 2, then (ðŠ ð , xð ) â graph(ð ) and therefore ¯ ) â graph(ð ) =â x ¯ â ð (¯ (¯ ðŠ, x ðŠ ) =â (¯ ðŠ, ⯠x) â ð so that ð is convex. 2.26 The graph of ð is graph(ðº) = { (ðœ, x) â Î Ã ð : x â ðº(ðœ) } Assume that (ðœð , xð ) â graph(ðº), ð = 1, 2. This means that xð â ðº(ðœ) and therefore ð ð (x, ðœ) †ðð for every ð and ð = 1, 2. Since ð ð is convex ð(ðŒx1 + (1 â ðŒ)x2 , ðŒðœ1 + (1 â ðŒ)ðœ 2 ) ⥠ðŒð(x1 , ðœ 1 ) + (1 â ðŒ)ð(x2 , ðœ2 ) ⥠ðð Therefore ðŒx1 +(1âðŒ)x2 â ðº(ðŒðœ 1 +(1âðŒ)ðœ2 ) and (ðŒðœ 1 +(1âðŒ)ðœ 2 , ðŒx1 +(1âðŒ)x2 ) â graph(ðº). ðº is convex. 73
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.27 The identity function ðŒð : ð â ð is deï¬ned by ðŒð (ð¥) = ð¥ for every ð¥ â ð. Therefore ð¥2 â»ð ð¥1 =â ðŒð (ð¥2 ) = ð¥2 â»ð ð¥1 = ðŒð (ð¥1 ) 2.28 Assume that ð and ð are increasing. Then, for every ð¥1 , ð¥2 â ð with ð¥2 â¿ð ð¥1 ð (ð¥2 ) â¿ð ð (ð¥1 ) =â ð(ð (ð¥2 )) â¿ð ð(ð (ð¥1 )) ð â ð is also increasing. Similarly, if ð and ð are strictly increasing, ð¥2 â»ð ð¥1 =â ð (ð¥2 ) â»ð ð (ð¥1 ) =â ð(ð (ð¥2 )) â»ð ð(ð (ð¥1 )) 2.29 For every ðŠ â ð (ð), there exists a unique ð¥ â ð such that ð (ð¥) = y. (For if ð¥1 , ð¥2 are such that ð (ð¥1 ) = ð (ð¥2 ), then ð¥1 = ð¥2 .) Therefore, ð is one-to-one and onto ð (ð), and so has an inverse (Exercise 2.4). Further ð¥2 > ð¥2 ââ ð (ð¥2 ) > ð (ð¥1 ) Therefore ð â1 is strictly increasing. 2.30 Assume ð : ð â â is increasing. Then, for every ð¥2 â¿ ð¥1 , ð (ð¥2 ) ⥠ð (ð¥1 ) which implies that âð (ð¥2 ) †âð (ð¥1 ). âð is decreasing. 2.31 For every ð¥2 â¿ ð¥1 . ð (ð¥2 ) ⥠ð (ð¥1 ) ð(ð¥2 ) ⥠ð(ð¥1 ) Adding (ð + ð)(ð¥2 ) = ð (ð¥2 ) + ð(ð¥2 ) ⥠ð (ð¥1 ) + ð (ð¥1 ) = (ð + ð)(ð¥1 ) That is, ð + ð is increasing. Similarly for every ðŒ ⥠0 ðŒð (ð¥2 ) ⥠ðŒð (ð¥1 ) and therefore ðŒð is increasing. By Exercise 1.186, the set of all increasing functionals is a convex cone in ð¹ (ð). If ð is strictly increasing, then for every ð¥2 â» ð¥1 , ð (ð¥2 ) > ð (ð¥1 ) ð(ð¥2 ) ⥠ð(ð¥1 ) Adding (ð + ð)(ð¥2 ) = ð (ð¥2 ) + ð(ð¥2 ) > ð (ð¥1 ) + ð(ð¥1 ) = (ð + ð)(ð¥1 ) ð + ð is strictly increasing. Similarly for every ðŒ > 0 ðŒð (ð¥2 ) > ðŒð (ð¥1 ) ðŒð is strictly increasing. 2.32 For every ð¥2 â» ð¥1 . ð (ð¥2 ) > ð (ð¥1 ) > 0 ð(ð¥2 ) > ð(ð¥1 ) > 0 and therefore (ð ð)(ð¥2 ) = ð (ð¥2 )ð(ð¥2 ) > ð (ð¥2 )ð(ð¥1 ) > ð (ð¥1 )ð(ð¥1 ) = (ð ð)(ð¥1 ) using Exercise 2.31. 74
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.33 By Exercise 2.31 and Example 2.53, each ðð is strictly increasing on â+ . That is ð¥1 < ð¥2 =â ðð (ð¥1 ) < ðð (ð¥2 ) for every ð
(2.32)
and therefore lim ðð (ð¥1 ) †lim ðð (ð¥2 )
ðââ
ðââ
This suï¬ces to show that ð(ð¥) = limðââ ðð (ð¥) is increasing (not strictly increasing). However, 1 + ð¥ is strictly increasing, and therefore by Exercise 2.31 ðð¥ = 1 + ð¥ + ð(ð¥) is strictly increasing on â+ . While it is the case that ð = lim ðð is strictly increasing on â+ , (2.32) does not suï¬ce to show this. 2.34 For every ð > 0, ð log ð¥ is strictly increasing (Exercise 2.32) and therefore ðð log ð¥ is strictly increasing (Exercise 2.28). For every ð < 0, âð log ð¥ is strictly increasing and therefore (Exercise 2.30 ð log ð¥ is strictly decreasing. Therefore ðð log ð¥ is strictly decreasing (Exercise 2.28). 2.35 Apply Exercises 2.31 and 2.28 to Example 2.56. 2.36 ð¢ is (strictly) increasing so that ð¥2 â¿ ð¥1 =â ð¢(ð¥2 ) ⥠ð¢(ð¥1 ) To show the converse, assume that ð¥1 , ð¥2 â ð with ð¢(ð¥2 ) ⥠ð¢(ð¥1 ). Since â¿ is complete, either ð¥2 â¿ ð¥1 or ð¥1 â» ð¥2 . However, the second possibility cannot occur since ð¢ is strictly increasing and therefore ð¥1 â» ð¥2 =â ð¢(ð¥1 ) > ð¢(ð¥2 ) contradicting the hypothesis that ð¢(ð¥2 ) ⥠ð¢(ð¥1 ). We conclude that ð¢(ð¥2 ) ⥠ð¢(ð¥1 ) =â ð¥2 â¿ ð¥1 2.37 Assume ð¢ represents the preference ordering â¿ on ð and let ð : â â â be strictly increasing. Then composition ð â ð¢ : ð â â is strictly increasing (Exercise 2.28). Therefore ð â ð¢ is a utility function (Example 2.58). Since ð is strictly increasing ð(ð¢(ð¥2 )) ⥠ð(ð¢(ð¥1 )) ââ ð¢(ð¥2 ) ⥠ð¢(ð¥1 ) ââ ð¥2 â¿ ð¥1 for every ð¥1 , ð¥2 â ð Therefore, ð â ð¢ also represents â¿. 2.38
¯ = ð§Â¯1 â¿ x and therefore z¯ â ðx+ . Similarly, 1. (a) Let ð§Â¯ = maxðð=1 ð¥ð . Then z ð let ð§ = minð=1 ð¥ð . Then z = ð§1 â ðxâ . Therefore, ðx+ and ðxâ are both nonempty. By continuity, the upper and lower contour sets â¿(x) and âŸ(x) are closed. ð is a closed cone. Since ðx+ = â¿(x) â© ð and ðxâ = âŸ(x) â© ð ðx+ and ðxâ are closed. (b) By completeness, ðx+ ⪠ðxâ = ð. Since ð is connected, ðx+ â© ðxâ â= â
. (Otherwise, ð is the union of two disjoint closed sets and hence the union of two disjoint open sets.) (c) Let zx â ðx+ â© ðxâ . Then z â¿ x and also z ⟠x. That is, z ⌠x. 75
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
(d) Suppose x ⌠z1x and x ⌠z2x with z1x â= z2x . Then either z1x > z2x or z1x < z2x . Without loss of generality, assume z2x > z1x . Then monotonicity and transitivity imply x ⌠z2x â» z1x ⌠x which is a contradiction. Therefore zx is unique. Let ð§x denote the scale of zx , that is zx = ð§x 1. For every x â âð+ , there is a unique zx ⌠x and the function ð¢ : âð+ â â given by ð¢(x) = ð§x is well-deï¬ned. Moreover x2 â¿ x1 ââ zx2 â¿ zx1 ââ ð§x2 ⥠ð§x1 ââ ð¢(x2 ) ⥠ð¢(x1 ) ð¢ represents the preference order â¿. 2.39
1. For every ð¥1 â â, (ð¥1 , 2) â»ð¿ (ð¥1 , 1) in the lexicographic order. If ð¢ represents â¿ð¿ , ð¢ is strictly increasing and therefore ð¢(ð¥1 , 2) > ð¢(ð¥1 , 1). There exists a rational number ð(ð¥1 ) such that ð¢(ð¥1 , 2) > ð(ð¥1 ) > ð¢(ð¥1 , 1).
2. The preceding construction associates a rational number with every real number ð¥1 â â. Hence ð is a function from â to the set of rational numbers ð. For any ð¥11 , ð¥21 â â with ð¥21 > ð¥11 ð(ð¥21 ) > ð¢(ð¥21 , 1) > ð¢(ð¥11 , 2) > ð(ð¥11 ) Therefore ð¥21 > ð¥11 =â ð(ð¥21 ) > ð(ð¥11 ) ð is strictly increasing. 3. By Exercise 2.29, ð has an inverse. This implies that ð is one-to-one and onto, which is impossible since ð is countable and â is uncountable (Example 2.16). This contradiction establishes that â¿ð¿ has no such representation ð¢. 2.40 Let a1 , a2 â ðŽ with a1 â¿2 a2 . Since the game is strictly competitive, a2 â¿1 a1 . Since ð¢1 represents â¿1 , ð¢1 (a2 ) ⥠ð¢1 (a1 ) which implies that âð¢1 (a2 ) †âð¢1 (a1 ), that is ð¢2 (a1 ) ⥠ð¢2 (a2 ) where ð¢2 = âð¢1 . Similarly ð¢2 (a1 ) ⥠ð¢2 (a2 ) =â ð¢1 (a1 ) †ð¢1 (a2 ) ââ a1 âŸ1 a2 =â a1 â¿2 a2 Therefore ð¢2 = âð¢1 represents â¿2 and ð¢1 (a) + ð¢2 (a) = 0 for every a â ðŽ 2.41 Assume ð â« ð . By superadditivity ð€(ð ) ⥠ð€(ð) + ð€(ð â ð) ⥠ð€(ð) 2.42 Assume ð£, ð€ â ðµ(ð) with ð€(ðŠ) ⥠ð£(ðŠ) for every ðŠ â ð. Then for any ð¥ â ð ð (ð¥, ðŠ) + ðœð€(ðŠ) ⥠ð (ð¥, ðŠ) + ðœð£(ðŠ) for every ðŠ â ð and therefore (ð ð€)(ð¥) = sup {ð (ð¥, ðŠ) + ðœð€(ðŠ)} ⥠sup {ð (ð¥, ðŠ) + ðœð£(ðŠ)} = (ð ð£)(ð¥) ðŠâðº(ð¥)
ðŠâðº(ð¥)
T is increasing. 76
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.43 For every ð2 ⥠ð1 â Î, if ð¥1 â ðº(ð1 ) and ð¥2 â ðº(ð2 ), then ð¥1 ⧠ð¥2 †ð¥1 and therefore ð¥1 ⧠ð¥2 â ðº(ð1 ). If ð¥1 ⥠ð¥2 , then ð¥1 âš ð¥2 = ð¥1 †ð(ð1 ) †ð(ð2 ) and therefore ð¥1 âš ð¥2 â ðº(ð2 ). On the other hand, if ð¥1 †ð¥2 , then ð¥1 âš ð¥2 = ð¥2 â ðº(ð2 ). 2.44 Assume ð is increasing, and let ð¥1 , ð¥2 â ð with ð¥2 â¿ ð¥1 . Let ðŠ1 â ð(ð¥1 ). Choose any ðŠ â² â ð(ð¥2 ). Since ð is increasing, ð(ð¥2 ) â¿ð ð(ð¥1 ) and therefore ðŠ2 = ðŠ1 âš ðŠ â² â ð(ð¥2 ). ðŠ2 â¿ ðŠ1 as required. Similarly, for every ðŠ2 â ð(ð¥2 ), there exists some ðŠ â² â ð(ð¥2 ) such that ðŠ1 = ðŠ Ⲡ⧠ðŠ2 â ð(ð¥1 ) with ðŠ2 â¿ ðŠ1 . 2.45 Since ð(ð¥) is a sublattice, sup ð(ð¥) â ð(ð¥) for every ð¥. Therefore, the function ð (ð¥) = sup ð(ð¥) is a selection. Similarly ð(ð¥) = inf ð(ð¥) is a selection. Both ð and ð are increasing (Exercise 1.50).
â 2.46 Let ð¥1 , ð¥2 belong to ðâ with ð¥2 â¿ ð¥1 . Choose y1 = (ðŠ11 , ðŠ21 , . . . , ðŠð1 ) â ð ðð (ð¥1 ) and y2 = (ðŠ12 , ðŠ22 , . . . , ðŠð2 ) â ð ðð (ð¥2 ). Then, for each ð = 1, 2, . . . , ð, ðŠð1 â ðð (ð¥1 ) and (ð¥1 ) and ðŠð1 âš ðŠð2 ââðð (ð¥2 ) for ðŠð2 â ðð (ð¥2 ). Since each ðð isâincreasing, ðŠð1 ⧠ðŠð2 â ððâ 1 2 1 1 2 each ð. Therefore y ⧠y â ð ðð (ð¥ ) and y âš y â ð ðð (ð¥2 ). ð(ð¥) = ð ðð (ð¥) is increasing. â© â© 2.47 Let ð¥1 , ð¥2 belong to ð with ð¥2 â¿ ð¥1 . Choose ðŠ 1 â ð ðð (ð¥1 ) and ðŠ 2 â ð ðð (ð¥2 ). Then ðŠ 1 â ðð (ð¥1 ) and ðŠ 2 â ðð (ð¥2 ) for each ð = 1, 2, . . . , ð. Since each ððâ©is increasing, ðð (ð¥1 ) and ðŠ 1 âšâ©ðŠ 2 â ðð (ð¥2 ) for each ð. Therefore ðŠ 1 ⧠ðŠ 2 â ð ðð (ð¥1 ) and ðŠ1 ⧠ðŠ2 â â© 1 2 ðŠ âš ðŠ â ð ðð (ð¥2 ). ð = ð ð is increasing. 2.48 Let ð be a selection from an always increasing correspondence ð : ð â ð . For every ð¥1 , ð¥2 â ð, ð (ð¥1 ) â ð(ð¥1 ) and ð (ð¥2 ) â ð(ð¥2 ). Since ð is always increasing ð¥1 â¿ð ð¥2 =â ð (ð¥1 ) â¿ð ð (ð¥2 ) ð is increasing. Conversely, assume every selection ð â ð is increasing. Choose any ð¥1 , ð¥2 â ð with ð¥1 â¿ ð¥2 . For every ðŠ1 â ð(ð¥1 ) and ðŠ2 â ð(ð¥2 ), there exists a selection ð with ðŠð = ð(ð¥ð ), ð = 1, 2. Since ð is increasing, ð¥1 â¿ð ð¥2 =â ðŠ1 â¿ð ðŠ2 ð is increasing. 2.49 Let ð¥1 , ð¥2 â ð. If ð is a chain, either ð¥1 â¿ ð¥2 or ð¥2 â¿ ð¥1 . Without loss of generality , assume ð¥2 â¿ ð¥1 . Then ð¥1 âš ð¥2 = ð¥2 and ð¥1 ⧠ð¥2 = ð¥1 and (2.17) is satisï¬ed as an identity. 2.50 (ð + ð)(ð¥1 âš ð¥2 ) + (ð + ð)(ð¥1 ⧠ð¥2 ) = ð (ð¥1 âš ð¥2 ) + ð(ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) + ð(ð¥1 ⧠ð¥2 ) = ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) + ð(ð¥1 âš ð¥2 ) + ð(ð¥1 ⧠ð¥2 ) ⥠ð (ð¥1 ) + ð (ð¥2 ) + ð(ð¥1 ) + ð(ð¥2 ) = (ð + ð)(ð¥1 ) + (ð + ð)(ð¥2 ) Similarly ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) ⥠ð (ð¥1 ) + ð (ð¥2 ) 77
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
implies ðŒð (ð¥1 âš ð¥2 ) + ðŒð (ð¥1 ⧠ð¥2 ) ⥠ðŒð (ð¥1 ) + ðŒð (ð¥2 ) for all ðŒ ⥠0. By Exercise 1.186, the set of all supermodular functions is a convex cone in ð¹ (ð). 2.51 Since ð is supermodular and ð is nonnegative deï¬nite, ( ) ð (ð¥1 âš ð¥2 )ð(ð¥1 âš ð¥2 ) ⥠ð (ð¥1 ) + ð (ð¥2 ) â ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 âš ð¥2 ) ( ) = ð (ð¥2 )ð(ð¥1 âš ð¥2 ) + ð (ð¥1 ) â ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 âš ð¥2 ) for any ð¥1 , ð¥2 â ð. Since ð and ð are increasing, this implies ( ) ð (ð¥1 âš ð¥2 )ð(ð¥1 âš ð¥2 ) ⥠ð (ð¥2 )ð(ð¥1 âš ð¥2 ) + ð (ð¥1 ) â ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 )
(2.33)
Similarly, since ð is nonnegative deï¬nite, ð supermodular, and ð and ð increasing ( ) ð (ð¥2 )ð(ð¥1 âš ð¥2 ) ⥠ð (ð¥2 ) ð(ð¥1 ) + ð(ð¥2 ) â ð(ð¥1 ⧠ð¥2 ) ( ) = ð (ð¥2 )ð(ð¥2 ) + ð (ð¥2 ) ð(ð¥1 ) â ð(ð¥1 ⧠ð¥2 ) ( ) ⥠ð (ð¥2 )ð(ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 ) â ð(ð¥1 ⧠ð¥2 ) Combining this inequality with (2.33) gives ( ) ð (ð¥1 âš ð¥2 )ð(ð¥1 âš ð¥2 ) ⥠ð (ð¥2 )ð(ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 ) â ð(ð¥1 ⧠ð¥2 ) ( ) + ð (ð¥1 ) â ð (ð¥1 ⧠ð¥2 ) ð(ð¥1 ) = ð (ð¥2 )ð(ð¥2 ) + ð (ð¥1 ⧠ð¥2 )ð(ð¥1 ) â ð (ð¥1 ⧠ð¥2 )ð(ð¥1 ⧠ð¥2 ) + ð (ð¥1 )ð(ð¥1 ) â ð (ð¥1 ⧠ð¥2 )ð(ð¥1 ) = ð (ð¥2 )ð(ð¥2 ) â ð (ð¥1 ⧠ð¥2 )ð(ð¥1 ⧠ð¥2) + ð (ð¥1 )ð(ð¥1 ) or ð ð(ð¥1 âš ð¥2 ) + ð ð(ð¥1 ⧠ð¥2 ) ⥠ð ð(ð¥1 ) + ð ð(ð¥2 ) ð ð is supermodular. (I acknowledge the help of Don Topkis in formulating this proof.) 2.52 Exercises 2.49 and 2.50. 2.53 For simplicity, assume that the ï¬rm produces two products. For every production plan y = (ðŠ1 , ðŠ2 ), y = (ðŠ1 , 0) âš (0, ðŠ2 ) 0 = (ðŠ1 , 0) ⧠(0, ðŠ2 ) If ð is strictly submodular ð(w, y) + ð(w, 0) < ð(w, (ðŠ1 , 0)) + ð(w, (0, ðŠ2 )) Since ð(w, 0) = 0 ð(w, y) < ð(w, (ðŠ1 , 0)) + ð(w, (0, ðŠ2 )) The technology displays economies of scope.
78
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.54 Assume (ð, ð€) is convex, that is ð€(ð ⪠ð ) + ð€(ð â© ð ) ⥠ð€(ð) + ð€(ð ) for every ð, ð â ð For all disjoint coalitions ð â© ð = â
ð€(ð ⪠ð ) ⥠ð€(ð) + ð€(ð ) ð€ is superadditive. 2.55 Rewriting (2.18), this implies ð€(ð ⪠ð ) â ð€(ð ) ⥠ð€(ð) â ð€(ð â© ð ) for every ð, ð â ð
(2.34)
Let ð â ð â ð â {ð} and let ð â² = ð ⪠{ð}. Substituting in (2.34) ð€(ð Ⲡ⪠ð ) â ð€(ð ) ⥠ð€(ð â² ) â ð€(ð â² â© ð ) Since ð â ð ð Ⲡ⪠ð = (ð ⪠{ð}) ⪠ð = ð ⪠{ð} ð â² â© ð = (ð ⪠{ð}) â© ð = ð Substituting in the previous equation gives the required result, namely ð€(ð ⪠{ð}) â ð€(ð ) ⥠ð€(ð ⪠{ð}) â ð€(ð) Conversely, assume that ð€(ð ⪠{ð}) â ð€(ð ) ⥠ð€(ð ⪠{ð}) â ð€(ð)
(2.35)
for every ð â ð â ð â {ð}. Let ð and ð be arbitrary coalitions. Assume ð â© ð â ð and ð â© ð â ð (otherwise (2.18) is trivially satisï¬ed). This implies that ð â ð â= â
. Assume these players are labelled 1, 2, . . . , ð, that is ð â ð = {1, 2, . . . , ð}. By (2.35) ð€(ð ⪠{1}) â ð€(ð) ⥠ð€((ð â© ð ) ⪠{1}) â ð€(ð â© ð )
(2.36)
Successively adding the remaining players in ð â ð ð€(ð ⪠{1, 2}) â ð€(ð ⪠{1}) ⥠ð€((ð â© ð ) ⪠{1, 2}) â ð€((ð â© ð ) ⪠{1}) .. . ( ) ð€(ð ⪠(ð â ð)) â ð€(ð ⪠{1, 2, . . . , ð â 1}) ⥠ð€ ð â© ð ) ⪠(ð â ð) â ð€((ð â© ð ) ⪠{1, 2, . . . , ð â 1}) Adding these inequalities to (2.36), we get ( ) ð€(ð ⪠(ð â ð)) â ð€(ð) ⥠ð€ ð â© ð ) ⪠(ð â ð) â ð€(ð â© ð ) This simpliï¬es to ð€(ð ⪠ð ) â ð€(ð) ⥠ð(ð ) â ð€(ð â© ð ) which can be arranged to give (2.18).
79
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.56 The cost allocation game is not convex. Let ð = {ðŽð, ðŸð }, ð = {ðŸð, ð ð }. Then ð ⪠ð = {ðŽð, ðŸð, ð ð } = ð and ð â© ð = {ðŸð } and ð€(ð ⪠ð ) + ð€(ð â© ð ) = 1530 < 1940 = 770 + 1170 = ð€(ð) + ð€(ð ) Alternatively, observe that TNâs marginal contribution to coalition {ðŸð, ð ð } is 1170, which is greater than its marginal contribution to the grand coalition {ðŽð, ðŸð, ð ð } (1530 â 770 = 760). 2.57 ð is supermodular if ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) ⥠ð (ð¥1 ) + ð (ð¥2 ) which can be rearranged to give ð (ð¥1 âš ð¥2 ) â ð (ð¥2 ) ⥠ð (ð¥1 ) â ð (ð¥1 ⧠ð¥2 ) If the right hand side of this inequality is nonnegative, then so a fortiori is the left hand side, that is ð (ð¥1 ) ⥠ð (ð¥1 ⧠ð¥2 ) =â ð (ð¥1 âš ð¥2 ) ⥠ð (ð¥2 ) If the right hand side is strictly positive, so must be the left hand side ð (ð¥1 ) > ð (ð¥1 ⧠ð¥2 ) =â ð (ð¥1 âš ð¥2 ) > ð (ð¥2 ) 2.58 Assume ð¥2 â¿ ð¥1 â ð and ðŠ2 â¿ð ðŠ2 â ð . Assume that ð displays increasing diï¬erences in (ð¥, ðŠ), that is ð (ð¥2 , ðŠ2 ) â ð (ð¥1 , ðŠ2 ) ⥠ð (ð¥2 , ðŠ1 ) â ð (ð¥1 , ðŠ1 )
(2.37)
ð (ð¥2 , ðŠ2 ) â ð (ð¥2 , ðŠ1 ) ⥠ð (ð¥1 , ðŠ2 ) â ð (ð¥1 , ðŠ1 )
(2.38)
Rearranging
Conversely, (2.38) implies (2.37) . 2.59 We showed in the text that supermodularity implies increasing diï¬erences. To show that reverse, assume that ð : ð à ð â â displays increasing diï¬erences in (ð¥, ðŠ). Choose any (ð¥1 , ðŠ1 ), (ð¥2 , ðŠ2 ) â ð à ð . If (ð¥1 , ðŠ1 ), (ð¥2 , ðŠ2 ) are comparable, so that either (ð¥1 , ðŠ1 ) â¿ (ð¥2 , ðŠ2 ) or (ð¥1 , ðŠ1 ) ⟠(ð¥2 , ðŠ2 ), then (2.17) holds has an equality. Therefore assume that (ð¥1 , ðŠ1 ), (ð¥2 , ðŠ2 ) are incomparable. Without loss of generality, assume that ð¥1 ⟠ð¥2 while ðŠ1 â¿ ðŠ2 . (This is where we require that ð and ð be chains). This implies (ð¥1 , ðŠ1 ) ⧠(ð¥2 , ðŠ2 ) = (ð¥1 , ðŠ2 ) and (ð¥1 , ðŠ1 ) âš (ð¥2 , ðŠ2 ) = (ð¥2 , ðŠ1 ) Increasing diï¬erences implies that ð (ð¥2 , ðŠ1 ) â ð (ð¥1 , ðŠ1 ) ⥠ð (ð¥2 , ðŠ2 ) â ð (ð¥1 , ðŠ2 ) which can be rewritten as ð (ð¥2 , ðŠ1 ) + ð (ð¥1 , ðŠ2 ) ⥠ð (ð¥1 , ðŠ1 ) + ð (ð¥2 , ðŠ2 ) Substituting (2.39) ( ) ( ) ð (ð¥1 , ðŠ1 ) âš (ð¥2 , ðŠ2 ) + ð (ð¥1 , ðŠ1 ) ⧠(ð¥2 , ðŠ2 ) ⥠ð (ð¥1 , ðŠ1 ) + ð (ð¥2 , ðŠ2 ) which establishes the supermodularity of ð on ð à ð (2.17). 80
(2.39)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.60 In the standard Bertrand model of oligopoly â the strategy space of each ï¬rm is â+ , a lattice. â ð¢ð (ðð , pâð ) is supermodular in ðð (Exercise 2.51). â If the other ï¬rmâs increase their prices from p1âð to p2âð , the eï¬ect on the demand for ï¬rm ðâs product is â ð (ðð , p2âð ) â ð (ðð , p1âð ) = ððð (ð2ð â ð1ð ) ðâ=ð
If the goods are gross substitutes, demand for ï¬rm ð increases and the amount of the increase is independent of ðð . Consequently, the eï¬ect on proï¬t will be increasing in ðð . That is the payoï¬ function (net revenue) has increasing diï¬erences in (ðð , pâð ). Speciï¬cally, â ð¢(ðð , p2âð ) â ð¢(ðð , p1âð ) = ððð (ðð â ð¯ð )(ð2ð â ð1ð ) ðâ=ð
For any price increase p2âð â© p1âð , the change in proï¬t ð¢(ðð , p2âð ) â ð¢(ðð , p1âð ) is increasing in ðð . Hence, the Bertrand oligopoly model is a supermodular game. 2.61 Suppose ð displays increasing diï¬erences so that for all ð¥2 â¿ ð¥1 and ðŠ2 â¿ ðŠ1 ð (ð¥2 , ðŠ2 ) â ð (ð¥1 , ðŠ2 ) ⥠ð (ð¥2 , ðŠ1 ) â ð (ð¥1 , ðŠ1 ) Then ð (ð¥2 , ðŠ1 ) â ð (ð¥1 , ðŠ1 ) ⥠0 =â ð (ð¥2 , ðŠ2 ) â ð (ð¥1 , ðŠ2 ) ⥠0 and ð (ð¥2 , ðŠ1 ) â ð (ð¥1 , ðŠ1 ) > 0 =â ð (ð¥2 , ðŠ2 ) â ð (ð¥1 , ðŠ2 ) > 0 2.62 For any ðœ â Îâ , let x1 , x2 â ð(ðœ). Supermodularity implies ð (x1 âš x2 , ðœ) + ð (x1 ⧠x2 , ðœ) ⥠ð (x1 , ðœ) + ð (x2 , ðœ) which can be rearranged to give ð (x1 âš x2 , ðœ) â ð (x2 , ðœ) ⥠ð (x1 , ðœ) â ð (x1 ⧠x2 , ðœ)
(2.40)
However x1 and x2 are both maximal in ðº(ðœ). ð (x2 , ðœ) ⥠ð (x1 âš x2 , ðœ) =â ð (x1 âš x2 , ðœ) â ð (x2 , ðœ) †0 ð (x1 , ðœ) ⥠ð (x1 ⧠x2 , ðœ) =â ð (x1 , ðœ) â ð (x1 ⧠x2 , ðœ) ⥠0 Substituting in (2.40), we conclude 0 ⥠ð (x1 âš x2 , ðœ) â ð (x2 , ðœ) ⥠ð (x1 , ðœ) â ð (x1 ⧠x2 , ðœ) ⥠0 This inequality must be satisï¬ed as an equality with ð (x1 âš x2 , ðœ) = ð (x2 , ðœ) ð (x1 ⧠x2 , ðœ) = ð (x1 , ðœ) That is x1 âš x2 â ð(ðœ) and x1 ⧠x2 â ð(ðœ). By Exercise 2.45, ð has an increasing selection. 81
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.63 As in the proof of the theorem, let ðœ1 , ðœ 2 belong to Î with ðœ 2 â¿ ðœ1 . Choose any optimal solutions x1 â ð(ðœ 1 ) and x2 â ð(ðœ2 ). We claim that x2 â¿ð x1 . Assume otherwise, that is assume x2 ââ¿ð x1 . This implies (Exercise 1.44) that x1 ⧠x2 â= x1 . Since x1 â¿ x1 ⧠x2 , we must have x1 â» x1 ⧠x2 . Strictly increasing diï¬erences implies ð (x1 , ðœ 2 ) â ð (x1 , ðœ1 ) > ð (x1 ⧠x2 , ðœ2 ) â ð (x1 ⧠x2 , ðœ 1 ) which can be rearranged to give ð (x1 , ðœ 2 ) â ð (x1 ⧠x2 , ðœ2 ) > ð (x1 , ðœ1 ) â ð (x1 ⧠x2 , ðœ 1 )
(2.41)
Supermodularity implies ð (x1 âš x2 , ðœ2 ) + ð (x1 ⧠x2 , ðœ2 ) ⥠ð (x1 , ðœ2 ) + ð (x2 , ðœ 2 ) which can be rearranged to give ð (x1 âš x2 , ðœ2 ) â ð (x2 , ðœ2 ) ⥠ð (x1 , ðœ2 ) â ð (x1 ⧠x2 , ðœ 2 ) Combining this inequality with (2.41) gives ð (x1 âš x2 , ðœ2 ) â ð (x2 , ðœ2 ) > ð (x1 , ðœ1 ) â ð (x1 ⧠x2 , ðœ 1 )
(2.42)
However x1 and x2 are optimal for their respective parameter values, that is ð (x2 , ðœ2 ) ⥠ð (x1 âš x2 , ðœ 2 ) =â ð (x1 âš x2 , ðœ 2 ) â ð (x2 , ðœ2 ) †0 ð (x1 , ðœ1 ) ⥠ð (x1 ⧠x2 , ðœ 1 ) =â ð (x1 , ðœ 1 ) â ð (x1 ⧠x2 , ðœ1 ) ⥠0 Substituting in (2.42), we conclude 0 ⥠ð (x1 âš x2 , ðœ2 ) â ð (x2 , ðœ2 ) > ð (x1 , ðœ1 ) â ð (x1 ⧠x2 , ðœ 1 ) ⥠0 This contradiction implies that our assumption that x2 ââ¿ð x1 is false. x2 â¿ð x1 as required. ð is always increasing. 2.64 The budget correspondence is descending in p and therefore ascending in âp. Consequently, the indirect utility function ð£(p, ð) =
sup
xâð(p,ð)
ð¢(x)
is increasing in âp, that is decreasing in p. 2.65 â= Let ðœ2 â¿ ðœ 1 and ðº2 â¿ð ðº1 . Select x1 â ð(ðœ 1 , ðº1 ) and x2 â ð(ðœ 2 , ðº2 ). Since ðº2 â¿ð ðº1 , x1 ⧠x2 â ðº1 . Since x1 is optimal (x1 â ð(ðœ1 , ðº1 )), ð (x1 , ðœ1 ) ⥠ð (x1 ⧠x2 , ðœ1 ). Quasisupermodularity implies ð (x1 âš x2 , ðœ 1 ) ⥠ð (x2 , ðœ1 ). By the single crossing condition ð (x1 âš x2 , ðœ2 ) ⥠ð (x2 , ðœ 2 ). Therefore x1 âš x2 â ð(ðœ2 , ðº2 ). Similarly, since ðº2 â¿ð ðº1 , x1 âš x2 â ðº(ðœ 2 ). But x2 is optimal, which implies that ð (x2 , ðœ2 ) ⥠ð (x1 âš x2 , ðœ2 ) or ð (x1 âš x2 , ðœ 2 ) †ð (x2 , ðœ2 ). The single crossing condition implies that a similar inequality holds at ðœ 1 , that is ð (x1 âš x2 , ðœ1 ) †ð (x2 , ðœ1 ). Quasisupermodularity implies that ð (x1 , ðœ 1 ) †ð (x1 â§x2 , ðœ 1 ). Therefore x1 â§x2 â ð(ðœ 1 , ðº1 ). Since x1 âš x2 â ð(ðœ 2 , ðº2 ) and x1 ⧠x2 â ð(ðœ 1 , ðº1 ), ð is increasing in (ðœ, ðº). =â To show that ð is quasisupermodular, suppose that ðœ is ï¬xed. Choose any x1 , x2 â ð. Let ðº1 = {x1 , x1 ⧠x2 } and ðº2 = {x2 , x1 âš x2 }. Then ðº2 â¿ð ðº1 . Assume that ð (x1 , ðœ) ⥠ð (x1 â§x2 , ðœ). Then x1 â ð(ðœ, ðº1 ) which implies that x1 âšx2 â ð(ðœ, ðº2 ). (If x2 â ð(ðœ, ðº2 ), then also x1 âš x2 â ð(ðœ, ðº2 ) since ð is increasing in (ðœ, ðº)). But this implies that ð (x1 âš x2 , ðœ) ⥠ð (x2 , ðœ). ð is quasisupermodular in ð. 82
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
To show that ð satisï¬es the single crossing condition, choose any x2 â¿ x1 and let ðº = {x1 , x2 }. Assume that ð (x2 , ðœ1 ) ⥠ð (x1 , ðœ 1 ). Then x2 â ð(ðœ1 , ðº) which implies that x2 â ð(ðœ 2 , ðº) for any ðœ 2 â¿ ðœ1 . (If x1 â ð(ðœ 2 , ðº), then also x1 âšx2 = x2 â ð(ðœ 2 , ðº) since ð is increasing in (ðœ, ðº).) But this implies that ð (x2 , ðœ2 ) ⥠ð (x1 , ðœ2 ). ð satisï¬es the single crossing condition. 2.66 First, assume that ð is continuous. Let ð be an open subset in ð and ð = ð â1 (ð ). If ð = â
, it is open. Otherwise, choose ð¥0 â ð and let ðŠ0 = ð (ð¥0 ) â ð . Since ð is open, there exists a neighborhood ð (ðŠ0 ) â ð . Since ð is continuous, there exists a corresponding neighborhood ð (ð¥0 ) with ð (ð (ð¥0 )) â ð (ð (ð¥0 )). Since ð (ð (ð¥0 )) â ð , ð (ð¥0 ) â ð. This establishes that for every ð¥0 â ð there exist a neighborhood ð (ð¥0 ) contained in ð. That is, ð is open in ð. Conversely, assume that the inverse image of every open set in ð is open in ð. Choose some ð¥0 â ð and let ðŠ0 = ð (ð¥0 ). Let ð â ð be a neighborhood of ðŠ0 . ð contains an open ball ðµð (ðŠ0 ) about ðŠ0 . By hypothesis, the inverse image ð = ð â1 (ðµð (ðŠ0 )) is open in ð. Therefore, there exists a neighborhood ð (ð¥0 ) â ð â1 (ðµð (ðŠ0 )). Since ðµð (ðŠ0 ) â ð , ð (ð (ð¥0 )) â ð . Since the choice of ð¥0 was arbitrary, we conclude that ð is continuous. 2.67 Assume ð is continuous. Let ð be a closed set in ð and let ð = ð â1 (ð ). Then, ð ð is open. By the previous exercise, ð â1 (ð ð ) = ð ð is open and therefore ð is closed. Conversely, for every open set ð â ð , ð ð is closed. By hypothesis, ð ð = ð â1 (ð ð ) is closed and therefore ð = ð â1 (ð ) is open. ð is continuous by the previous exercise. 2.68 Assume ð is continuous. Let ð¥ð be a sequence converging to ð¥ Let ð be a neighborhood of ð (ð¥). Since ð is continuous, there exists a neighborhood ð â ð¥ such that ð (ð) â ð . Since ð¥ð converges to ð¥, there exists some ð such that ð¥ð â ð for all ð ⥠ð . Consequently ð (ð¥ð ) â ð for every ð ⥠ð . This establishes that ð (ð¥ð ) â ð (ð¥). Conversely, assume that for every sequence ð¥ð â ð¥, ð (ð¥ð ) â ð (ð¥). We show that if ð were not continuous, it would be possible to construct a sequence which violates this hypothesis. Suppose then that ð is not continuous. Then there exists a neighborhood / ð . In ð of ð (ð¥) such that for every neighborhood ð of ð¥, there is ð¥â² â ð with ð (ð¥â² ) â particular, consider the sequence of open balls ðµ1/ð (ð¥). For every ð, choose a point / ð . Then ð¥ð â ð¥ but ð (ð¥ð ) does not converge to ð (ð¥). ð¥ð â ðµ1/ð (ð¥) with ð (ð¥ð ) â This contradicts the assumption. We conclude that ð must be continuous. 2.69 Since ð is one-to-one and onto, it has an inverse ð = ð â1 which maps ð onto ð. Let ð be an open set in ð. Since ð is open, ð = ð â1 (ð) = ð (ð) is open in ð . Therefore ð = ð â1 is continuous. 2.70 Assume ð is continuous. Let (ð¥ð , ðŠ ð ) be a sequence of points in graph(ð ) converging to (ð¥, ðŠ). Then ðŠ ð = ð (ð¥ð ) and ð¥ð â ð¥. Since ð is continuous, ðŠ = ð (ð¥) = limðââ ð (ð¥ð ) = limðââ ðŠ ð . Therefore (ð¥, ðŠ) â graph(ð ) which is therefore closed. 2.71 By the previous exercise, ð continuous implies graph(ð ) closed. Conversely, suppose graph(ð ) is closed and let ð¥ð be a sequence converging to ð¥. Then (ð¥ð , ð (ð¥ð ))) is a sequence in graph(ð ). Since ð is compact, ð (ð¥ð ) contains a subsequence which converges ðŠ. Since graph(ð ) is closed, (ð¥, ðŠ) â graph(ð ) and therefore ðŠ = ð (ð¥) and ð (ð¥ð ) â ð (ð¥). 2.72 Let ð be an open set in ð. Since ð and ð are continuous, ð â1 (ð ) is open in ð and ð â1 (ð â1 (ð )) is open in ð. But ð â1 (ð â1 (ð )) = (ð â ð)â1 (ð ). Therefore ð â ð is continuous. 2.73 Exercises 1.201 and 2.68.
83
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.74 Let ð¢ be deï¬ned as in Exercise 2.38. Let (xð ) be a sequence converging to x. Let ð§ ð = ð¢(xð ) and ð§ = ð¢(x). We need to show that ð§ ð â ð§. (ð§ ð ) has a convergent subsequence. Let ð§Â¯ = maxð ð¥ð and ð§ = minð ð¥ð . Then ð§ â [ð§, ð§Â¯]. Fix some ð > 0. Since xð â x, there exists some ð such that â¥xð â xâ¥â < ð for every ð ⥠ð . Consequently, for all ð ⥠ð , the terms of the sequence (ð§ ð ) lie in the compact set [ð§ â ð, ð§Â¯ + ð]. Hence, (ð§ ð ) has a convergent subsequence (ð§ ð ). Every convergent subsequence (ð§ ð ) converges to ð§. Suppose not. That is, suppose there exists a convergent subsequence which converges to ð§ â² . Without loss of generality, assume ð§ â² > ð§. Let ð§Ë = 12 (ð§ + ð§ â² ) and let z = ð§1, zâ² = ð§ â² 1, Ëz = ð§Ë1 be the corresponding commodity bundles (see Exercise 2.38). Since ð§ ð â ð§ â² > ð§Ë, there exists some ð such that ð§ ð > ð§Ë for every ð ⥠ð . This implies that xð ⌠zð â» zË for every ð ⥠ð by monotonicity. Now xð â x and continuity of preferences implies that x â¿ Ëz. However x ⌠z which implies that z â¿ zË which contradicts monotonicity, since Ë z > z. Consequently, every convergent subsequence (ð§ ð ) converges to ð§. 2.75 Assume ð is compact. Let ðŠ ð be a sequence in ð (ð). There exists a sequence ð¥ð in ð with ðŠ ð = ð (ð¥ð ). Since ð is compact, it contains a convergent subsequence ð¥ð â ð¥. If ð is continuous, the subsequence ðŠ ð = ð (ð¥ð ) converges in ð (ð) (Exercise 2.68). Therefore ð (ð) is compact. Assume ð is connected but ⪠ð (ð) is not. This means â© there exists open subsets ðº and ð» in ð such that ð (ð) â ðº ð» and (ðº â© ð (ð)) (ð» â© ð (ð)) = â
. This implies that ð = ð â1 (ðº) ⪠ð â1 (ð») is a disconnection of ð, which contradicts the connectedness of ð. 2.76 Let ð be any open set in ð. Its complement ð ð is closed and therefore compact. Consequently, ð (ð ð ) is compact (Exercise 2.3) and hence closed. Since ð is one-to-one and onto, ð (ð) is the complement of ð (ð ð ), and thus open in ð . Therefore, ð is an open mapping. By Exercise 2.69, ð â1 is continuous and ð is a homeomorphism. 2.77 Assume ð continuous. The sets {ð (ð¥) ⥠ð} and {ð (ð¥) †ð} are closed subsets of the â and hence â¿(ð) = ð â1 {ð (ð¥) ⥠ð} and âŸ(ð) = ð â1 {ð (ð¥) †ð} are closed subsets of ð (Exercise 2.67). Conversely, assume that all upper â¿(ð) and lower âŸ(ð) contour sets are closed. This implies that the sets â»(ð) and âº(ð) are open. Let ðŽ be an open set in â. Then for every ð â ðŽ, there exists an open ball ðµðð (ð) â ðŽ ⪠ðŽ= ðµðð (ð) ðâðŽ
For every ð â ðŽ, ðµðð (ð) = (ð â ðð , ð + ðð ) and ð â1 (ðµðð (ð)) = â»(ð â ðð ) â© âº(ð + ðð ) which is open. Consequently ⪠⪠( ) â»(ð â ðð ) â© âº(ð + ðð ) ð â1 (ðµðð (ð)) = ð â1 (ðŽ) = ðâðŽ
ðâðŽ
is open. ð is continuous by Exercise 2.66. 84
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.78 Choose any ð¥0 â ð and ð > 0. Since ð is continuous, there exists ð¿1 such that ð(ð¥, ð¥0 ) < ð¿1 =â â£ð (ð¥) â ð (ð¥0 )⣠< ð/2 Similarly, there exists ð¿2 such that ð(ð¥, ð¥0 ) < ð¿2 =â â£ð(ð¥) â ð(ð¥0 )⣠< ð/2 Let ð¿ = min{ð¿1 , ð¿2 }. Then, provided ð(ð¥, ð¥0 ) < ð¿ â£(ð + ð)(ð¥) â (ð + ð)(ð¥0 )⣠= â£ð (ð¥) + ð(ð¥) â ð (ð¥0 ) â ð(ð¥0 )⣠†â£ð (ð¥) â ð (ð¥0 )⣠+ â£ð(ð¥) â ð(ð¥0 )⣠0 such that â£ð (ð¥) â ð (ð¥0 )⣠< ð and â£ð(ð¥) â ð(ð¥0 )⣠< ð whenever ð(ð¥, ð¥0 ) < ð¿. Consequently, while ð(ð¥, ð¥0 ) < ð¿ â£ð (ð¥)⣠†â£ð (ð¥) â ð (ð¥0 )⣠+ â£ð (ð¥0 )⣠< ð + â£ð (ð¥0 )⣠†1 + â£ð (ð¥0 )⣠and â£(ð ð)(ð¥) â (ð ð)(ð¥0 )⣠= â£ð (ð¥)ð(ð¥) â ð (ð¥0 )ð(ð¥0 )⣠= â£ð (ð¥)(ð(ð¥) â ð(ð¥0 )) + ð(ð¥0 )(ð (ð¥) â ð (ð¥0 ))⣠†â£ð (ð¥)⣠â£ð(ð¥) â ð(ð¥0 )⣠+ â£ð(ð¥0 )⣠â£ð (ð¥) â ð (ð¥0 )⣠< ð(1 + â£ð (ð¥0 )⣠+ â£ð(ð¥0 )â£) Given ð > 0, let ð = min{1, ð/(1 + â£ð (ð¥0 )⣠+ â£ð(ð¥0 )â£)}. Then, we have shown that there exists ð¿ > 0 such that ð(ð¥, ð¥0 ) < ð¿ =â â£(ð ð)(ð¥) â (ð ð)(ð¥0 )⣠< ð Therefore, ð ð is continuous at ð¥0 . 2.80 Apply Exercises 2.78 and 2.72. 2.81 For any ð â â, the upper and lower contour sets of ð âš ð, namely { ð¥ : max{ð (ð¥), ð(ð¥)} ⥠ð} = {ð¥ : ð (ð¥) ⥠ð } ⪠{ ð¥ : ð(ð¥) ⥠ð } { ð¥ : max{ð (ð¥), ð(ð¥)} †ð} = {ð¥ : ð (ð¥) †ð } â© { ð¥ : ð(ð¥) †ð } are closed. Therefore ð âš ð is continuous (Exercise 2.77). Similarly for ð ⧠ð. 2.82 The set ð = ð (ð) is compact (Proposition 2.3). We want to show that ð has both largest and smallest elements. Assume otherwise, that is assume that ð has no largest element. Then, the set of intervals {(ââ, ð¡) : ð¡ â ð } forms an open covering of ð . Since ð is compact, there exists a ï¬nite subcollection of intervals {(ââ, ð¡1 ), (ââ, ð¡2 ), . . . , (ââ, ð¡ð )} which covers ð . Let ð¡â be the largest of these ð¡ð . Then ð¡â does not belong to any of the intervals {(ââ, ð¡1 ), (ââ, ð¡2 ), . . . , (ââ, ð¡ð )}, contrary to the fact that they cover ð . This contradiction shows that, contrary to our assumption, there must exist a largest element ð¡â â ð , that is ð¡â ⥠ð¡ for all ð¡ â ð . Let ð¥â â ð â1 (ð¡â ). Then ð¡â = ð (ð¥â ) ⥠ð (ð¥) for all ð¥ â ð. The existence of a smallest element is proved analogously. 85
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.83 By Proposition 2.3, ð (ð) is connected and hence an interval (Exercise 1.95). 2.84 The range ð (ð) is a compact subset of â (Proposition 2.3). Therefore ð is bounded (Proposition 1.1). Ë 2.85 Let ð¶(ð) denote the set of all continuous (not necessarily bounded) functionals on ð. Then Ë ð¶(ð) = ðµ(ð) â© ð¶(ð) Ë ðµ(ð), ð¶(ð) are a linear subspaces of the set of all functionals ð¹ (ð) (Exercises 2.11, Ë 2.78 respectively). Therefore ð¶(ð) = ðµ(ð) â© ð¶(ð) is a subspace of ð¹ (ð) (Exercise 1.130). Clearly ð¶(ð) â ðµ(ð). Therefore ð¶(ð) is a linear subspace of ðµ(ð). Let ð be a bounded function in the closure of ð¶(ð), that is ð â ð¶(ð). We show that ð is continuous. For any ð > 0, there exists ð0 â ð¶(ð) such that â¥ð â ð0 ⥠< ð/3. Therefore â£ð (ð¥) â ð0 (ð¥)⣠< ð/3 for every ð¥ â ð. Choose some ð¥0 â ð. Since ð0 is continuous, there exists ð¿ > 0 such that ð(ð¥, ð¥0 ) < ð¿ =â â£ð0 (ð¥) â ð0 (ð¥0 )⣠< ð/3 Therefore, for every ð¥ â ð such that ð(ð¥, ð¥0 ) < ð¿ â£ð (ð¥) â ð (ð¥0 )⣠= â£ð (ð¥) â ð0 (ð¥) + ð0 (ð¥) â ð0 (ð¥0 ) + ð0 (ð¥0 ) â ð (ð¥0 )⣠†â£ð (ð¥) â ð0 (ð¥)⣠+ â£ð0 (ð¥) â ð0 (ð¥0 )⣠+ â£ð0 (ð¥0 ) â ð (ð¥0 )⣠< ð/3 + ð/3 + ð/3 = ð Therefore ð is continuous at ð¥0 . Since ð¥0 was arbitrary, we conclude that is continuous everywhere, that is ð â ð¶(ð). Therefore ð¶(ð) = ð¶(ð) and ð¶(ð) is closed in ðµ(ð). Since ðµ(ð) is complete (Exercise 2.11), we conclude that ð¶(ð) is complete (Exercise 1.107). Therefore ð¶(ð) is a Banach space. 2.86 For every ðŒ â â, { ð¥ : ð (ð¥) ⥠ðŒ } = {ð¥ : âð (ð¥) †âðŒ } and therefore { ð¥ : ð (ð¥) ⥠ðŒ } is closed ââ {ð¥ : âð (ð¥) †âðŒ } is closed 2.87 Exercise 2.77. 2.88 1 implies 2 Suppose ð is upper semi-continuous. Let ð¥ð be a sequence converging to ð¥0 . Assume ð (ð¥ð ) â ð. For every ðŒ < ð, there exists some ð such that ð (ð¥ð ) > ðŒ for every ð ⥠ð . Hence ð¥0 â { ð¥ : ð (ð¥) ⥠ðŒ } = { ð¥ : ð (ð¥) ⥠ðŒ } since ð is upper semi-continuous. That is, ð (ð¥0 ) ⥠ðŒ for every ðŒ < ð. Hence ð (ð¥0 ) ⥠ð = limðââ ð (ð¥ð ). 2 implies 3 Let (ð¥ð , ðŠ ð ) be a sequence in hypo ð which converges to (ð¥, ðŠ). That is, ð¥ð â ð¥, ðŠ ð â ðŠ and ðŠ ð †ð (ð¥ð ). Condition 2 implies that ð (ð¥) ⥠ðŠ. Hence, (ð¥, ðŠ) â hypo ð . Therefore hypo ð is closed. 3 implies 1 For ï¬xed ðŒ â â, let ð¥ð be a sequence in { ð¥ : ð (ð¥) ⥠ðŒ }. Suppose ð¥ð â ð¥0 . Then, the sequence (ð¥ð , ðŒ) converges to (ð¥0 , ðŒ) â hypo ð . Hence ð (ð¥0 ) ⥠ðŒ and ð¥0 â { ð¥ : ð (ð¥) ⥠ðŒ }, which is therefore closed (Exercise 1.106). 86
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 2.89 Let ð = supð¥âð ð (ð¥), so that ð (ð¥) †ð for every ð¥ â ð
(2.43)
There exists a sequence ð¥ð in ð with ð (ð¥ð ) â ð . Since ð is compact, there exists a convergent subsequence ð¥ð â ð¥â and ð (ð¥ð ) â ð . However, since ð is upper semi-continuous, ð (ð¥â ) ⥠lim ð (ð¥ð ) = ð . Combined with (2.43), we conclude that ð (ð¥â ) = ð . 2.90 Choose some ð > 0. Since ð is uniformly continuous, there exists some ð¿ > 0 such that ð(ð (ð¥ð ), ð (ð¥ð )) < ð for every ð¥ð , ð¥ð â ð such that ð(ð¥ð , ð¥ð ) < ð¿. Let (ð¥ð ) be a Cauchy sequence in ð. There exists some ð such that ð(ð¥ð , ð¥ð ) < ð¿ for every ð, ð ⥠ð . Uniform continuity implies that ð(ð (ð¥ð ), ð (ð¥ð )) < ð for every ð, ð ⥠ð . (ð (ð¥ð )) is a Cauchy sequence. 2.91 Suppose not. That is, suppose ð is continuous but not uniformly continuous. Then there exists some ð > 0 such that for ð = 1, 2, . . . , there exist points ð¥1ð , ð¥2ð such that ð(ð¥1ð , ð¥2ð ) < 1/ð but ð(ð (ð¥1ð ), ð (ð¥2ð )) ⥠ð
(2.44)
Since ð is compact, (ð¥1ð ) has a subsequence (ð¥1ð ) converging to some ð¥ â ð. By construction (ð(ð¥1ð , ð¥2ð ) < 1/ð), the sequence (ð¥2ð ) also converges to ð¥ and by continuity lim ð (ð¥1ð ) = lim ð (ð¥2ð )
ðââ
ðââ
which contradicts (2.44). 2.92 Assume ð is Lipschitz with constant ðœ. For any ð > 0, let ð¿ = ð/2ðœ. Then, provided ð(ð¥, ð¥0 ) †ð¿ ð(ð (ð¥), ð (ð¥0 )) †ðœð(ð¥, ð¥0 ) = ðœð¿ = ðœ
ð ð = 0. ð¹ is totally bounded (Exercise 1.113), so that there exist ï¬nite set of functions {ð1 , ð2 , . . . , ðð } in F such that ð
min â¥ð â ðð ⥠†ð/3 ð=1
Each ðð is uniformly continuous (Exercise 2.91), so that there exists ð¿ð > 0 such that ð(ð¥, ð¥0 ) †ð¿ =â ð(ðð (ð¥), ðð (ð¥0 ) < ð/3 Let ð¿ = min{ð¿1 , ð¿2 , . . . , ð¿ð }. Given any ð â ð¹ , let ð be such that â¥ð â ðð ⥠< ð/3. Then for any ð¥, ð¥0 â ð, ð(ð¥, ð¥0 ) †ð¿ implies ð(ð (ð¥), ð (ð¥0 ) †ð(ð (ð¥), ðð (ð¥)) + ð(ðð (ð¥), ðð (ð¥0 )) + ð(ðð (ð¥0 ), ð (ð¥0 )) <
ð ð ð + + =ð 3 3 3
for every ð â ð¹ . Therefore, ð¹ is equicontinuous. Conversely, assume that ð¹ â ð¶(ð) is closed, bounded and equicontinuous. Let (ðð ) be a bounded equicontinuous sequence of functions in ð¹ . We show that (ðð ) has a convergent subsequence. 1. First, we show that for any ð > 0, there is exists a subsequence (ðð ) such that â¥ðð â ððⲠ⥠< ð for every ðð , ððâ² in the subsequence. Since the functions are equicontinuous, there exists ð¿ > 0 such that ð(ðð (ð¥) â ðð (ð¥0 ) <
ð 3
for every ð¥, ð¥0 in ð with ð(ð¥, ð¥0 ) †ð¿. Since ð is compact, it is totally bounded (Exercise 1.113). That is, there exist a ï¬nite number of open balls ðµð¿ (ð¥ð ), ð = 1, 2 . . . , ð which cover ð. The sequence (ðð (ð¥1 ), ðð (ð¥2 , . . . , ðð (ð¥ð )) is a bounded sequence in âð . By the Bolzano-Weierstrass theorem (Exercise 1.119), this sequence has a convergent subsequence (ðð (ð¥1 ), ðð (ð¥2 ), . . . , ðð (ð¥ð )) such that ðð (ð¥ð ) â ððâ² (ð¥ð ) < ð/3 for ð and every ðð , ððâ² in the subsequence. Consequently, for any ð¥ â ð, there exists ð such that ð(ðð (ð¥), ððâ² (ð¥) †ð(ðð (ð¥), ðð (ð¥ð )) + ð(ðð (ð¥ð ), ððâ² (ð¥ð )) + ð(ððâ² (ð¥ð ), ððâ² (ð¥)) ð ð ð < + + =ð 3 3 3 That is, â¥ðð â ððⲠ⥠< ð for every ðð , ððâ² in the subsequence. 2. Choose a ball ðµ1 of radius 1 in ð¶(ð) which contains inï¬nitely many elements of (ðð ). Applying step 1, there exists a ball ðµ2 of radius 1/2 containing inï¬nitely many elements of (ðð ). Proceeding in this fashion, we obtain a nested sequence ðµ1 â ðµ2 â . . . of balls in ð¶(ð) such that (a) ð(ðµð ) â 0 and (b) each ðµð contains inï¬nitely many terms of (ðð ). Choosing ððð â ðµð gives a convergent subsequence. 88
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
2.96 Let ð â ð¹ . Then for every ð > 0 there exists ð¿ > 0 and ð â ð¹ such that â¥ð â ð⥠< ð/3 and ð(ð¥, ð¥0 ) †ð¿ =â ð(ð (ð¥), ð (ð¥0 ) < ð/3 so that if ð(ð¥, ð¥0 ) †ð¿ â¥ð(ð¥) â ð(ð¥0 )⥠†â¥ð (ð¥) â ð(ð¥)⥠+ â¥ð (ð¥) â ð (ð¥0 )⥠+ â¥ð (ð¥0 ) â ð(ð¥0 )⥠<
ð ð ð + + =ð 3 3 3
2.97 For every ð â ð ðâ (ð ð ) = { ð¥ â ð : ð(ð¥) â© ð ð â= â
} ð+ (ð ) = { ð¥ â ð : ð(ð¥) â ð } For every x â ð either ð(ð¥) â ð or ð(ð¥) â© ð ð â= â
but not both. Therefore ð+ (ð ) ⪠ðâ (ð ð ) = ð ð+ (ð ) â© ðâ (ð ð ) = â
That is
( )ð ð+ (ð ) = ðâ (ð ð )
2.98 Assume ð¥ â ð(ð )â1 . Then ð(ð¥) = ð , ð(ð¥) â ð and ð¥ â ð+ (ð ). Now assume ð¥ â ð+ (ð ) so that ð(ð¥) â ð . Consequently, ð(ð¥) â© ð = ð(ð¥) â= â
and ð¥ â ðâ (ð ). 2.99 The respective inverses are: {ð¡1 } {ð¡2 } {ð¡1 , ð¡2 } {ð¡2 , ð¡3 } {ð¡1 , ð¡2 , ð¡3 }
ðâ1 2 â
â
{ð 1 } {ð 2 } â
ð+ 2 â
â
{ð 1 } {ð 2 } {ð 1 , ð 2 }
ðâ 2 {ð 1 } {ð 1 , ð 2 } {ð 1 , ð 2 } {ð 1 , ð 2 } {ð 1 , ð 2 }
2.100 Let ð be an open interval meeting ð(1), that is ð(1) â© ð â= â
. Since ð(1) = {1}, we must have 1 â ð and therefore ð(ð¥) â© ð â= â
for every ð¥ â ð. Therefore ð is lhc at ð¥ = 1. On the other hand, the open interval ð = (1/2, 3/2) contains ð(1) but it does not contain ð(ð¥) for any ð¥ > 1. Therefore, ð is not uhc at ð¥ = 1. 2.101 Choose any open set ð â ð and ð¥ â ð. Since ð(ð¥) = ðŸ = ð(ð¥â² ) for every ð¥, ð¥â² â ð â ð(ð¥) â ð if and only if ð(ð¥â² ) â ð for every ð¥, ð¥â² â ð â ð(ð¥) â© ð â= â
if and only if ð(ð¥â² ) â© ð â= â
for every ð¥, ð¥â² â ð. Consequently, ð is both uhc and lhc at all ð¥ â ð. 2.102 First assume that the ð is uhc. Let ð be any open subset in ð and ð = ð+ (ð ). If ð = â
, it is open. Otherwise, choose ð¥0 â ð so that ð(ð¥0 ) â ð . Since ð is uhc, there exists a neighborhood ð(ð¥0 ) such that ð(ð¥) â ð for every ð¥ â ð(ð¥0 ). That is, ð(ð¥0 ) â ð+ (ð ) = ð. This establishes that for every ð¥0 â ð there exist a neighborhood ð(ð¥0 ) contained in ð. That is, ð is open in ð. Conversely, assume that the upper inverse of every open set in ð is open in ð. Choose some ð¥0 â ð and let ð be an open set containing ð(ð¥0 ). Let ð = ð+ (ð ). ð is an open set containing ð¥0 . That is, ð is a neighborhood of ð¥0 with ð(ð¥) â ð for every ð¥ â ð. Since the choice of ð¥0 was arbitrary, we conclude that ð is uhc. The lhc case is analogous. 89
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.103 Assume ð is uhc and ð be any closed set in ð . By Exercise 2.97 [ ] ðâ (ð ) = ð+ (ð ð ) ð ð is open. By the previous exercise, ð+ (ð ð ) is open which implies that ðâ (ð ) is closed. Conversely, assume ðâ (ð ) is closed for every closed set ð . Let ð be an open subset of ð so that ð ð is closed. Again by Exercise 2.97, [ ] ð+ (ð ) = ðâ (ð ð ) By assumption ðâ (ð ð ) is closed and therefore ð+ (ð ) is open. By the previous exercise, ð is uhc. The lhc case is analogous. 2.104 Assume that ð is uhc at ð¥0 . We ï¬rst show that (ðŠ ð ) is bounded and hence has a convergent subsequence. Since ð(ð¥0 ) is compact, there exists a bounded open set ð containing ð(ð¥0 ). Since ð is uhc, there exists a neighborhood ð of ð¥0 such that ð(ð¥) â ð for ð¥ â ð. Since ð¥ð â ð¥0 , there exists some ð such that ð¥ð â ð for every ð ⥠ð . Consequently, ð(ð¥ð ) â ð for every ð ⥠ð and therefore ðŠ ð â ð for every ð ⥠ð . The sequence ðŠ ð is bounded and hence has a convergent subsequence ðŠ ð â ðŠ0 . To complete the proof, we have to show that ðŠ0 â ð(ð¥0 ). Assume not, assume that / ð(ð¥0 ). Then, there exists an open set ð containing ð(ð¥0 ) such that ðŠ0 â / ð ðŠ0 â (Exercise 1.93). Since ð is uhc, there exists ð such that ð(ð¥ð ) â ð for every ð ⥠ð . This implies that ðŠ ð â ð for every ð ⥠ð . Since ðŠ ð â ðŠ0 , we conclude that ðŠ0 â ð , contradicting the speciï¬cation of ð . Conversely, suppose that for every sequence ð¥ð â ð¥0 , ðŠ ð â ð(ð¥ð ), there is a subsequence of ðŠ ð â ðŠ0 â ð(ð¥0 ). Suppose that ð is not uhc at ð¥0 . That is, there exists an open set ð â ð(ð¥0 ) such that every neighborhood contains some ð¥ with ð(ð¥) ââ ð . From the sequence of neighborhoods ðµ1/ð (ð¥0 ), we can construct a sequence ð¥ð â ð¥ and ðŠ ð â ð(ð¥ð ) but ðŠ ð â / ð . Such a sequence cannot have a subsequence which converges to ðŠ 0 â ð(ð¥) , contradicting the hypothesis. We conclude that ð must be uhc at ð¥0 . 2.105 Assume that ð is lhc. Let ð¥ð be a sequence converging to ð¥0 and ðŠ0 â ð(ð¥0 ). Consider the sequence of open balls ðµ1/ð (ðŠ0 ), ð = 1, 2, . . . . Note that every ðµ1/ð (ðŠ0 ) meets ð(ð¥0 ). Since ð is lhc, there exists a sequence (ð ð ) of neighborhoods of ð¥0 such that ð(ð¥) â© ðµ1/ð â= â
for every ð¥ â ð ð . Since ð¥ð â ð¥, for every ð, there exists some ðð such that ð¥ð â ðð for every ð ⥠ðð . Without loss of generality, we can assume that ð1 < ð2 < ð3 . . . . We can now construct the desired sequence ðŠ ð . For each ð = 1, 2, . . . , choose ðŠ ð in the set ð(ð¥ð ) â© ðµ 1/m where ðð †ð †ðð+1 since ð ⥠ðð =â ð¥ð â ðð =â ð(ð¥ð ) â© ðµ1/ð â= â
Since ðŠ ð â ðµ 1/m (ðŠ0 ), the sequence (ðŠ ð ) converges to ðŠ0 and ð â â. Conversely, assume that ð is not lhc at ð¥0 , that is there exists an open set ð with ð â© ð(ð¥0 ) â= â
such that every neighborhood ð â ð¥0 contains some ð¥ with ð(ð¥) â© ð = â
. Therefore, there exists a sequence ð¥ð â ð¥ with ð(ð¥)â©ð = â
. Choose any ðŠ0 â ð(ð¥0 )â©ð . By assumption, there exists a sequence ðŠ ð â ðŠ with ðŠ ð â ð(ð¥ð ). Since ð is open and ðŠ0 â ð , there exists some ð such that ðŠ ð â ð for all ð ⥠ð , for which ð(ðŠ ð ) â© ð â= â
. This contradiction establishes that ð is lhc at ð¥0 . 90
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 2.106
1. Assume ð is closed. For any ð¥ â ð, let (ðŠ ð ) be a sequence in ð(ð¥). Since ð is closed, ðŠ ð â ðŠ â ð(ð¥). Therefore ð(ð¥) is closed.
2. Assume ð is closed-valued and uhc. Choose any (ð¥, ðŠ) â / graph(ð). Since ð(ð¥) is closed, there exist disjoint open sets ð1 and ð2 in ð such that ðŠ â ð1 and ð(ð¥) â ð2 (Exercise 1.93). Since ð is uhc, ð+ (ð2 ) is a neighborhood of ð¥. Therefore ð+ (ð2 ) à ð1 is a neighborhood of (ð¥, ðŠ) disjoint from graph(ð). Therefore the complement of graph(ð) is open, which implies that graph(ð) is closed. 3. Since ð is closed and ð compact, ð is compact-valued. Let (ð¥ð ) â ð¥ be a sequence in ð and (ðŠ ð ) a sequence in ð with ðŠ ð â ð(ð¥ð ). Since ð is compact, there exists a subsequence ðŠ ð â ðŠ. Since ð is closed, ðŠ â ð(ð¥). Therefore, by Exercise 2.104, ð is uhc. 2.107 Assume ð is closed-valued and uhc. Then ð is closed (Exercise 2.106). Conversely, if ð is closed, then ð(ð¥) is closed for every ð¥ (Exercise 2.106). If ð is compact, then ð is compact-valued (Exercise 1.110). By Exercise 2.104, ð is uhc. 2.108 ð1 is closed-valued (Exercise 2.106). Similarly, ð2 is closed-valued (Proposition 1.1). Therefore, for every ð¥ â ð, ð(ð¥) = ð1 (ð¥) â© ð2 (ð¥) is closed (Exercise 1.85) and hence compact (Exercise 1.110). Hence ð is compact-valued. Now, for any ð¥0 â ð, let ð be an open neighborhood of ð(ð¥0 ). We need to show that there is a neighborhood ð of ð¥0 such that ð(ð) â ð . Case 1 ð â ð2 (ð¥0 ): Since ð2 is uhc, there exists a neighborhood such that ð â ð¥0 such that ð2 (ð) â ð which implies that ð(ð) â ð2 (ð) â ð Case 2 ð ââ ð2 (ð¥0 ): Let ðŸ = ð2 (ð¥0 ) â ð â= â
. For every ðŠ â ðŸ, there exist neighborhoods ððŠ (ð¥0 ) and ð (ðŠ) such that ð1 (ððŠ (ð¥0 )) â© ð (ðŠ) = â
(Exercise 1.93). The sets ð (ðŠ) constitute an open covering of ðŸ. Since ðŸ is compact, there exists a ï¬nite subcover, that is there exists a ï¬nite number of elements ðŠ1 , ðŠ2 , . . . ðŠð such that ð âª
ðŸâ
ð (ðŠð )
ð=1
âªð Let ð (ðŸ) denote ð=1 ð (ðŠð ). Note that ð âªð (ðŸ) is an open set containing ð2 (ð¥0 ). Since ð2 is uhc, there exists a neighborhood ð â² (ð¥0 ) such that ð2 (ð â² (ð¥0 )) â ð ⪠ð (ðŸ). Let ð(ð¥0 ) =
ð â©
ððŠð (ð¥0 ) â© ð â² (ð¥0 )
ð=1
ð(ð¥0 ) is an open neighborhood of ð¥0 for which ð1 (ð(ð¥0 )) â© ð (ðŸ) = â
and ð2 (ð(ð¥0 )) â ð ⪠ð (ðŸ) from which we conclude that ð(ð(ð¥0 )) = ð1 (ð(ð¥0 )) â© ð2 (ð(ð¥0 )) â ð âð 2.109 1. Let x â ð(p, ð) â© ð . Then x â ð(p, ð) and ð=1 ðð ð¥ð †ð. Since ð is Ë = ðŒx â ð and open, there exists ðŒ < 1 such that x ð â ð=1
ðð ð¥Ëð = ðŒ
ð â
ðð ð¥ð <
ð=1
91
ð â ð=1
ðð ð¥ð †ð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. (a) Suppose that ð(p, ð) is not lhc. Then for every neighborhood ð of (p, ð), there exists (pâ² , ðâ² ) â ð such that ð(pâ² , ðâ² ) â© ð = â
. In particular, for every open ball ðµð (p, ð), there exists a point (pð , ðð ) â ðµð (p, ð) such that ð(pð , ðð ) â© ð = â
. ((pð , ðð )) is the required sequence. (b) By construction, â¥pð â p⥠< 1/ð â 0 which implies that ððð â ðð for every ð. Therefore (Exercise 1.202) â â Ëð â ðð ð¥Ëð < ð and ðð â ð ððð ð¥ and therefore there exists ð such that â Ë ð < ðð ðð ð ð¥ which implies that Ë â ð(pð , ðð ) x (c) Also by construction ð(pð , ðð ) â© ð = â
which implies ð(pð , ðð ) â ð ð and therefore Ë â ð(pð , ðð ) =â x Ëâ x /ð Ë â The assumption that ð(p, ð) is not lhc at (p, ð) implies that x / ð , contraË â ð. dicting the conclusion in part 1 that x 3. This contradiction establishes that (p, ð) is lhc at (p, ð). Since the choice of (p, ð) was arbitrary, we conclude that the budget correspondence ð(p, ð) is lhc for all (p, ð) â ð (assuming ð = âð+ ). 4. In the previous example (Example 2.89), we have shown that ð(p, ð) is uhc. Hence, â the budget correspondence is continuous for all (p, ð) such that ð > ð inf xâð ð=1 ðð ð¥ð . 2.110 We give two alternative proofs. Proof 1 Let ð = {ð} be an open cover of ð(ðŸ). For every ð¥ â ðŸ, ð(ð¥) â ð(ðŸ) is compact and hence can be covered by a ï¬nite number of the sets ð â ð. Let ðð¥ denote the union of the ï¬nite cover of ð(ð¥). Since ð is uhc, every ð+ (ðð¥ ) is open in ð. Therefore { ð+ (ðð¥ ) : ð¥ â ðŸ } is an open covering of ðŸ. If ðŸ is compact, it contains an ï¬nite covering { ð+ (ðð¥1 ), ð+ (ðð¥2 ), . . . , ð+ (ðð¥ð ) }. The sets ðð¥1 , ðð¥2 , . . . , ðð¥ð are a ï¬nite subcovering of ð(ðŸ). Proof 2 Let (ðŠ ð ) be a sequence in ð(ðŸ). We have to show that (ðŠ ð ) has a convergent subsequence with a limit in ð(ðŸ). For every ðŠ ð , there is an ð¥ð with ðŠ ð â ð(ð¥ð ). Since ðŸ is compact, the sequence (ð¥ð ) has a convergent subsequence ð¥ð â ð¥ â ðŸ. Since ð is uhc, the sequence (ðŠ ð ) has a subsequence (ðŠ ð ) which converges to ðŠ â ð(ð¥) â ð(ðŸ). Hence the original sequence (ðŠ ð ) has a convergent subsequence. 2.111 The sets ð, ð(ð), ð2 (ð), . . . form a sequence of nonempty compact sets. Since ð(ð) â ð, ð2 (ð) â ð(ð) and so on, the sequence of sets ðð ð is decreasing. Let ðŸ=
â â©
ðð (ð)
ð=1
By the nested intersection theorem (Exercise 1.117), ðŸ â= â
. Since ðŸ â ððâ1 (ð), ð(ðŸ) â ðð (ð) for every ð, which implies that ð(ðŸ) â ðŸ. 92
Solutions for Foundations of Mathematical Economics To show that ðŸ that ðŠ â ð(ð¥ð ). ð¥ð â ðð (ð) for (Exercise 2.107),
c 2001 Michael Carter â All rights reserved
â ð(ðŸ), let ðŠ â ðŸ. For every ð there exists an ð¥ð â ðð (ð) such Since ð is compact, there exists a subsequence ð¥ð â ð¥0 . Since every ð, ð¥0 â ðŸ. The sequence (ð¥ð , ðŠ) â (ð¥0 , ðŠ). Since ð is closed ðŠ â ð(ð¥0 ). Therefore ðŠ â ð(ðŸ) which implies that ðŸ â ð(ðŸ).
2.112 ð(ð¥) is compact for every ð¥ â ð by Tychonoï¬âs theorem (Proposition 1.2). Let ð¥ð â ð¥ be a sequence in ð and let ðŠ ð = (ðŠ1ð , ðŠ2ð , . . . , ðŠðð ) with ðŠðð â ð(ð¥ð ) be a corresponding sequence of points in ð . For each ðŠðð , ð = 1, 2, . . . , ð, there exists a â² subsequence ðŠðð â ðŠð with ðŠð â ðð (ð¥) (Exercise 2.104). Therefore ðŠ = (ðŠ1 , ðŠ2 , . . . , ðŠð ) â ð(ð¥) which implies that ð is uhc. 2.113 Let ð£ â ð¶(ð). For every x â ð, the maximand ð (ð¥, ðŠ) + ðœð£(ðŠ) is a continuous function on a compact set ðº(ð¥). Therefore the supremum is attained, and max can replace sup in the deï¬nition of the operator ð (Theorem 2.2). ð ð£ is the value function for the constrained optimization problem max { ð (ð¥, ðŠ) + ðœð£(ðŠ) }
ðŠâðº(ð¥)
satisfying the requirements of the continuous maximum theorem (Theorem 2.3), which ensures that ð ð£ is continuous on ð. We have previously shown that ð ð£ is bounded (Exercise 2.18). Therefore ð ð£ â ð¶(ð). 2.114
1. ð has a least upper bound since ð is a complete lattice. Let ð â = sup ð. Then ð â = â¿(ð â ) is a complete sublattice of ð (Exercise 1.48).
2. For every ð â ð, ð ⟠ð â and since ð is increasing and ð is a ï¬xed point ð = ð (ð ) ⟠ð (ð â ) Therefore ð (ð â ) â ð â . (ð (ð â ) is an upper bound of ð). Again, since ð is increasing, this implies that ð (ð¥) â¿ ð (ð â ) for every ð¥ â ð â . Therefore ð (ð â ) â ð â . 3. Let ð be the restriction of ð to the sublattice ð â . Since ð (ð â ) â ð â , ð is an increasing function on a complete lattice. Applying Theorem 2.4, ð has a smallest ï¬xed point ð¥ Ë. 4. ð¥ Ë is a ï¬xed point of ð , that is ð¥ Ë â ðž. Furthermore, ð¥Ë â ð â . Therefore ð¥Ë is an upper bound for ð in ðž. Moreover, ð¥ Ë is the smallest ï¬xed point of ð in ð â . Therefore, ð¥ Ë is the least upper bound of ð in ðž. 5. By Exercise 1.47, this implies that ðž is a complete lattice. In Example 2.91, if ð = {(2, 1), (1, 2)}, ð â = {(2, 2), (3, 2), (2, 3), (3, 3)} and ð¥Ë = (3, 3). 2.115
1. For every ð¥ â ð , there exists some ðŠð¥ â ð(ð¥) such that ðŠð¥ ⟠ð¥. Moreover, ð¥) such that since ð is increasing and ð¥ Ë âŸ ð¥, there exists some ð§ð¥ â ð(Ë ð§ð¥ ⟠ðŠð¥ ⟠ð¥ for every ð¥ â ð
2. Let ð§Ë = inf{ð§ð¥ } Ë. (a) Since ð§ð¥ ⟠ð¥ for every ð¥ â ð , ð§Ë = inf{ð§ð¥ } ⟠inf{ð¥} = ð¥ (b) Since ð(Ë ð¥) is a complete sublattice of ð, ð§Ë = inf{ð§ð¥ } â ð(Ë ð¥). 3. Therefore, ð¥ Ë â ð. 4. Since ð§Ë ⟠ð¥Ë and ð is increasing, there exists some ðŠ â ð(Ë ð§ ) such that ðŠ ⟠ð§Ë â ð(Ë ð¥) Hence ð§Ë â ð . 93
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5. This implies that ð¥ Ë âŸ ð§Ë. Therefore ð¥ Ë = ð§Ë â ð(Ë ð¥) ð¥ Ë is a ï¬xed point of ð. 6. Since ðž â ð , ð¥Ë = inf ð is the least ï¬xed point of ð. 2.116
1. Let ð â ðž and ð â = sup ð. For every ð¥ â ð, ð¥ â ð(ð¥). Since ð is increasing, there exists some ð§ð¥ â ð(ð â ) such that ð§ð¥ â¿ ð¥.
2. Let ð§ â = sup ð§ð¥ . Then (a) Since ð§ð¥ â¿ ð¥ for every ð¥ â ð, ð§ â = sup ð§ð¥ â¿ sup ð¥ = ð â (b) ð§ â â ð(ð â ) since ð(ð â ) is a complete sublattice. 3. Deï¬ne ð â = { ð¥ â ð : ð¥ â¿ ð for every ð â ð } ð â is the set of all upper bounds of ð in ð. Then ð â is a complete lattice, since ð â = â¿(ð â ) 4. Let ð : ð â â ð â be the correspondence ð(ð¥) = ð(ð¥) â© ð(ð¥) where ð : ð â â ð â is the constant correspondence deï¬ned by ð(ð¥) = ð â for every ð¥ â ð â . Then (a) Since ð is increasing, for every ð¥ â¿ ð â , there exists some ðŠð¥ â ð(ð¥) such that ðŠð¥ â¿ ð â . Therefore ð(ð¥) â= â
for every ð¥ â ð â . (b) Both ð(ð¥) and ð(ð¥) are complete sublattices for every ð¥ â ð â . Therefore ð(ð¥) is a complete sublattice for every ð¥ â ð â . (c) Since both ð and ð are increasing on ð â , ð is increasing on ð â (Exercise 2.47). 5. By the previous exercise, ð has a least ï¬xed point ð¥Ë. 6. ð¥ Ë â ð â is an upper bound of ð. Therefore ð¥ Ë is the least upper bound of ð in ðž. 7. By the previous exercise, ðž has a least element. Since we have shown every subset ð â ðž has a least upper bound, this establishes that ðž is complete lattice (Exercise 1.47). 2.117 For any ð, let a1âð , a2âð â ðŽâð with a2âð â¿ a1âð . Let ð ¯1ð = ð (a1âð ) and ð ¯2ð = ð (a2âð ). 2 1 1 1 We want to show that ð ¯ð â¿ ð ¯ð . Since ð ¯ð â ðµ(aâð ) and ðµ(aâð ) is increasing, there ¯1ð . (Exercise 2.44). Therefore exists some ðð â ðµ(a2âð ) such that ðð â¿ ð sup ðµ(aâð ) = ð ¯2ð â¿ ðð â¿ ð ¯1ð ð¯ð is increasing. 2.118 For any player ð, their best response correspondence ðµð (aâð ) is 1. increasing by the monotone maximum theorem (Theorem 2.1). 2. a complete sublattice of ðŽð for every aâð â ðŽâð (Corollary 2.1.1). 94
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics The joint best response correspondence
ðµ(a) = ðµ1 (aâ1 ) Ã ðµ2 (aâ2 ) Ã â
â
â
à ðµð (aâð ) is also 1. increasing (Exercise 2.46) 2. a complete sublattice of ðŽ for every a â ðŽ Therefore, the best response correspondence ðµ(a) satisï¬es the conditions of Zhouâs theorem, which implies that the set ðž of ï¬xed points of ðµ is a nonempty complete lattice. ðž is precisely the set of Nash equilibria of the game. 2.119 In proving the theorem, we showed that ð(ð¥ð , ð¥ð+ð ) â€
ðœð ð(ð¥0 , ð¥1 ) 1âðœ
for every ð, ð ⥠0. Letting ð â â, ð¥ð+ð â ð¥ and therefore ð(ð¥ð , ð¥) â€
ðœð ð(ð¥0 , ð¥1 ) 1âðœ
Similarly, for every ð, ð ⥠0 ð(ð¥ð , ð¥ð+ð ) †ð(ð¥ð , ð¥ð+1 ) + ð(ð¥ð+1 , ð¥ð+2 ) + â
â
â
+ ð(ð¥ð+ðâ1 , ð¥ð+ð ) †(ðœ + ðœ 2 + â
â
â
+ ðœ ð )ð(ð¥ðâ1 , ð¥ð ) â€
ðœ(1 â ðœ ð ) ð(ð¥ðâ1 , ð¥ð ) 1âðœ
Letting ð â â, ð¥ð+ð â ð¥ and ðœ ð â 0 so that ð(ð¥ð , ð¥) â€
ðœ ð(ð¥ðâ1 , ð¥ð ) 1âðœ
2.120 First observe that ð (ð¥) ⥠1 for every ð¥ ⥠1. Therefore ð : ð â ð. For any ð¥, ð§ â ð ð¥ â ðŠ + ð¥2 â ð (ð¥) â ð (ðŠ) = ð¥âðŠ 2(ð¥ â ðŠ) Since
1 ð¥ðŠ
=
1 1 â 2 ð¥ðŠ
†1 for all ð¥, ðŠ â ð â
so that
2 ðŠ
ð (ð¥) â ð (ðŠ) 1 1 ††2 ð¥âðŠ 2
ð (ð¥) â ð (ðŠ) â£ð (ð¥) â ð (ðŠ)⣠1 = †ð¥âðŠ â£ð¥ â ðŠâ£ 2
or â£ð (ð¥) â ð (ðŠ)⣠†ð is a contraction on ð with modulus 1/2. 95
1 â£ð¥ â ðŠâ£ 2
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ð is closed and hence complete (Exercise 1.107). Therefore, ð has a ï¬xed point. That is, there exists ð¥0 â ð such that ð¥0 = ð (ð¥0 ) =
1 2 (ð¥0 + ) 2 ð¥0
Rearranging
so that ð¥0 =
2ð¥20 = ð¥20 + 2 =â ð¥20 = 2
â 2.
Letting ð¥0 = 2 ð¥1 =
3 1 (2 + 1) = 2 2
Using the error bounds in Corollary 2.5.1, â ðœð ð(ð¥ð , 2) †ð(ð¥0 , ð¥1 ) 1âðœ (1/2)ð = 1/2 1/2 1 = ð 2 1 < 0.001 = 1024 when ð = 10. Therefore, we conclude that 10 iterations are ample to reduce the error below 0.001. Actually, with experience, we can reï¬ne this a priori estimate. In Example 1.64, we calculated the ï¬rst ï¬ve terms of the sequence to be (2, 1.5, 1.416666666666667, 1.41421568627451, 1.41421356237469) We observe that ð(ð¥3 , ð¥4 ) = 1.41421568627451 â 1.41421356237469) = 0.0000212389982 so that using the second inequality of Corollary 2.5.1 ð(ð¥4 ,
â
2) â€
1/2 0.0000212389982 < 0.001 1/2
ð¥4 = 1.41421356237469 is the desired approximation. 2.121 Choose any ð¥0 â ð. Deï¬ne the sequence ð¥ð = ð (ð¥ð ) = ð ð (ð¥0 ). Then (ð¥ð ) is a Cauchy sequence in ð converging to ð¥. Since ð is closed, ð¥ â ð. 2.122 By the Banach ï¬xed point theorem, ð ð has a unique ï¬xed point ð¥. Let ðœ be the Lipschitz constant of ð ð . We have to show ð¥ is a ï¬xed point of ð ð(ð (ð¥), ð¥) = ð(ð (ð ð (ð¥), ð ð (ð¥)) = ð(ð ð (ð (ð¥), ð ð (ð¥)) †ðœð(ð (ð¥), ð¥) Since ðœ < 1, this implies that ð(ð (ð¥), ð¥) = 0 or ð (ð¥) = ð¥. ð¥ is the only ï¬xed point of ð Suppose ð§ = ð (ð§) is another ï¬xed point of ð . Then ð§ is a ï¬xed point of ð ð and ð(ð¥, ð§) = ð(ð ð (ð¥), ð ð (ð§)) †ðœð(ð¥, ð§) which implies that ð¥ = ð§. 96
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.123 By the Banach ï¬xed point theorem, for every ð â Î, there exists ð¥ð â ð such that ðð (ð¥ð ) = ð¥ð . Choose any ð0 â Î. ð(ð¥ð , ð¥ð0 ) = ð(ðð (ð¥ð ), ðð0 (ð¥ð0 )) †ð(ðð (ð¥ð ), ðð (ð¥ð0 )) + ð(ðð (ð¥ð0 ), ðð0 (ð¥ð0 )) †ðœð(ð¥ð , ð¥ð0 ) + ð(ðð (ð¥ð0 ), ðð0 (ð¥ð0 )) (1 â ðœ)ð(ð¥ð , ð¥ð0 ) †ð(ðð (ð¥ð0 ), ðð0 (ð¥ð0 )) ð(ð¥ð , ð¥ð0 ) â€
ð(ðð (ð¥ð0 ), ðð0 (ð¥ð0 )) â0 (1 â ðœ)
as ð â ð0 . Therefore ð¥ð â ð¥ð0 . 2.124
1. Let x be a ï¬xed point of ð . Then x satisï¬es x = (ðŒ â ðŽ)x + c = x â ðŽx + ð which implies that ðŽx = ð.
2. For any x1 , x2 â ð ð (x1 ) â ð (x2 ) = (ðŒ â ðŽ)(x1 â x2 ) †â¥ðŒ â ðŽâ¥ x1 â x2 Since ððð = 1, the norm of ðŒ â ðŽ is â¥ðŒ â ðŽâ¥ = max ð
â
â£ððð ⣠= ð
ðâ=ð
and ð (x1 ) â ð (x2 ) †ð x1 â x2 By the assumption of strict diagonal dominance, ð < 1. Therefore ð is a contraction and has a unique ï¬xed point x. 2.125
1. ð(ð¥) = { ðŠ â â ðº(ð¥) : ð (ð¥, ðŠ â ) + ðœð£(ðŠ â ) = ð£(ð¥) } = {ðŠ â â ðº(ð¥) : ð (ð¥, ðŠ â ) + ðœð£(ðŠ â ) = sup {ð (ð¥, ðŠ) + ðœð£(ðŠ)}} ðŠâðº(ð¥)
â
â
â
= {ðŠ â ðº(ð¥) : ð (ð¥, ðŠ ) + ðœð£(ðŠ ) ⥠ð (ð¥, ðŠ) + ðœð£(ðŠ) for every ðŠ â ðº(ð¥)} = arg max {ð (ð¥, ðŠ) + ðœð£(ðŠ)} ðŠâðº(ð¥)
2. ð(ð¥) is the solution correspondence of a standard constrained maximization problem, with ð¥ as parameter and ðŠ the decision variable. By assumption the maximand ð (ð¥, ðŠ) = ð (ð¥, ðŠ) + ðœð£(ðŠ) is continuous and the constraint correspondence ðº(ð¥) is continuous and compact-valued. Applying the continuous maximum theorem (Theorem 2.3), ð is nonempty, compact-valued and uhc. 3. We have just shown that ð(ð¥) is nonempty for every ð¥ â ð. Starting at ð¥0 , choose some ð¥â1 â ð(ð¥0 ). Then choose ð¥â2 â ð(ð¥â1 ). Proceeding in this way, we can construct a plan xâ = ð¥0 , ð¥â1 , ð¥â2 , . . . such that ð¥âð¡+1 â ð(ð¥âð¡ ) for every ð¡ = 0, 1, 2, . . . .
97
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
4. Since ð¥âð¡+1 â ð(ð¥âð¡ ) for every ð¡, x satisï¬es Bellmanâs equation, that is ð£(ð¥âð¡ ) = ð (ð¥âð¡ , ð¥âð¡+1 ) + ðœð£(ð¥âð¡+1 ),
ð¡ = 0, 1, 2, . . .
Therefore x is optimal (Exercise 2.17). 2.126
1. In the previous exercise (Exercise 2.125) we showed that the set ð of solutions to Bellmanâs equation (Exercise 2.17) is the solution correspondence of the constrained maximization problem ð(ð¥) = arg max { ð (ð¥, ðŠ) + ðœð£(ðŠ) } ðŠâðº(ð¥)
This problem satisï¬es the requirements of the monotone maximum theorem (Theorem 2.1), since the objective function ð (ð¥, ðŠ) + ðœð£(ðŠ) â supermodular in ðŠ â displays strictly increasing diï¬erences in (ð¥, ðŠ) since for every ð¥2 ⥠ð¥1 ð (ð¥2 , ðŠ) + ðœð£(ðŠ) â ð (ð¥1 , ðŠ) + ðœð£(ðŠ) = ð (ð¥2 , ðŠ) â ð (ð¥1 , ðŠ) â ðº(ð¥) is increasing. By Corollary 2.1.2, ð(ð¥) is always increasing. 2. Let xâ = (ð¥0 , ð¥â1 , ð¥â2 , . . . ) be an optimal plan. Then (Exercise 2.17) ð¥âð¡+1 â ð(ð¥âð¡ ),
ð¡ = 0, 1, 2, . . .
Since ð is always increasing ð¥âð¡ ⥠ð¥âð¡â1 =â ð¥âð¡+1 ⥠ð¥âð¡ for every ð¡ = 1, 2, . . . . Similarly ð¥âð¡ †ð¥âð¡â1 =â ð¥âð¡+1 †ð¥âð¡ xâ = (ð¥0 , ð¥â1 , ð¥â2 , . . . ) is a monotone sequence. 2.127 Let ð(ð¥) = ð (ð¥) â ð¥. ð is continuous (Exercise 2.78) with ð(0) ⥠0 and ð(1) †0 By the intermediate value theorem (Exercise 2.83), there exists some point ð¥ â [0, 1] with ð(ð¥) = 0 which implies that ð (ð¥) = ð¥. 2.128
1. To show that a label min{ ð : ðœð †ðŒð â= 0 } exists for every x â ð, assume to the contrary that, for some x â ð, ðœð > ðŒð for every ð = 0, 1, . . . , ð. This implies ð â
ðœð >
ð=0
ð â
ðŒð = 1
ð=0
contradicting the requirement that ð â
ðœð = 1 for every ð (x) â ð
ð=0
98
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. The barycentric coordinates of vertex xð are ðŒð = 1 with ðŒð = 0 for every ð â= ð. Therefore the rule assigns vertex xð the label ð. 3. Similarly, if x belongs to a proper face of ð, it coordinates relative to the vertices not in that face are 0, and it cannot be assigned a label corresponding to a vertex not in the face. To be concrete, suppose that x â conv {x1 , x2 , x4 }. Then x = ðŒ1 x1 + ðŒ2 x2 + ðŒ4 x4 ,
ðŒ1 + ðŒ2 + ðŒ4 = 1
/ {1, 2, 4}. Therefore and ðŒð = 0 for ð â x +ââ min{ ð : ðœð †ðŒð â= 0 } â {1, 2, 4} 2.129
1. Since ð is compact, it is bounded (Proposition 1.1) and therefore it is contained in a suï¬ciently large simplex ð .
2. By Exercise 3.74, there exists a continuous retraction ð : ð â ð. The composition ð â ð : ð â ð â ð . Furthermore as the composition of continuous functions, ð â ð is continuous (Exercise 2.72). Therefore ð â ð has a ï¬xed point xâ â ð , that is ð â ð(xâ ) = xâ . 3. Since ð â ð(x) â ð for every x â ð , we must have ð â ð(xâ ) = xâ â ð. Therefore, ð(xâ ) = xâ which implies that ð (xâ ) = xâ . That is, xâ is a ï¬xed point of ð . 2.130 Convexity of ð is required to ensure that there is a continuous retraction of the simplex onto ð. 2.131
1. ð (ð¥) = ð¥2 on ð = (0, 1) or ð (ð¥) = ð¥ + 1 on ð = â+ .
2. ð (ð¥) = 1 â ð¥ on ð = [0, 1/3] ⪠[2/3, 1]. 3. Let ð = [0, 1] and deï¬ne
{ ð (ð¥) =
1 0
0 †ð¥ < 1/2 otherwise
2.132 Suppose such a function exists. Deï¬ne ð (x) = âð(x). Then ð : ðµ â ðµ continously, and has no ï¬xed point since for â x â ð, ð (x) = âð(x) = âx â= x â x â ðµ â ð, ð (x) â / ðµ â ð and thereforeð (x) â= x Therefore ð has no ï¬xed point contradicting Brouwerâs theorem. 2.133 Suppose to the contrary that ð has no ï¬xed point. For every x â ðµ, let ð(z) denote the point where the line segment from ð (x) through x intersects the boundary ð of ðµ. Since ð is continuous and ð (x) â= x, ð is a continuous function from ðµ to its boundary, that is a retraction, contradicting Exercise 2.132. We conclude that ð must have a ï¬xed point. 2.134 No-retraction =â Brouwer Note ï¬rst that the no-retraction theorem (Exercise 2.132) generalizes immediately to a closed ball about 0 of arbitrary radius. Assume that ð is a continuous operator on a compact, convex set ð in a ï¬nite dimensional normed linear space. There exists a closed ball ðµ containing ð (Proposition 1.1). Deï¬ne ð : ðµ â ð by ð(y) = { x â ð : x is closest to y } As in Exercise 2.129, ð is well-deï¬ned, continuous and ð(x) = x for every x â ð. ð â ð : ðµ â ð â ðµ and has a ï¬xed point xâ = ð (ð(xâ )) by Exercise 2.133. Since 99
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
ð â ð(x) â ð for every x â ðµ, we must have ð â ð(xâ ) = xâ â ð. Therefore, ð(xâ ) = xâ which implies that ð (xâ ) = xâ . That is, xâ is a ï¬xed point of ð . Brouwer =â no-retraction Exercise 2.132. 2.135 Let Îð , ð = 1, 2, . . . be a sequence of simplicial partitions of ð in which the maximum diameter of the subsimplices tend to zero as ð â â. By Spernerâs lemma (Proposition 1.3), every partition Îð has a completely labeled subsimplex with vertices xð0 , xð1 , . . . , xðð . By construction of an admissible labeling, each xðð belongs to a face containing xð , that is xðð â conv {xð , . . . } and therefore xðð â ðŽð ,
ð = 0, 1, . . . , ð â²
Since ð is compact, each sequence xðð has a convergent subsequence xðð . Moreover, since the diameters of the subsimplices converge to zero, these subsequences must converge to the same point, say xâ . That is, â²
lim xðð = xâ ,
ð = 0, 1, . . . , ð
ðâ² ââ
Since the sets ðŽð are closed, xâ â ðŽð for every ð and therefore ð â©
xâ â
ðŽð â= â
ð=0
2.136
=â Let ð : ð â ð be a continuous operator on an ð-dimensional simplex ð with vertices x0 , x1 , . . . , xð . For ð = 0, 1, . . . , ð, let ðŽð = { x â ð : ðœð †ðŒð } where ðŒ0 , ðŒ1 , . . . , ðŒð and ðœ0 , ðœ1 , . . . , ðœð are the barycentric coordinates of x and ð (x) respectively. Then â ð continuous =â ðŽð closed for every ð = 0, 1, . . . , ð (Exercise 1.106) â Let x â conv { xð : ð â ðŒ } for some ðŒ â { 0, 1, . . . , ð }. Then â
ðŒð = 1 =
ð â
ðœð
ð=0
ðâðŒ
which implies that ðœð †ðŒð for some ð â ðŒ, so that x â ðŽð . Therefore ⪠ðŽð conv { xð : ð â ðŒ } â ðâðŒ
Therefore the collection ðŽ0 , ðŽ1 , . . . , ðŽð satisï¬es the hypotheses of the K-K-M theorem and their intersection is nonempty. That is, there exists xâ â
ð â©
ðŽð â= â
with ðœðâ †ðŒâð ,
ð = 0, 1, . . . , ð
ð=0
where â ðŒâ and â ðœ â are the barycentric coordinates of xâ and ð (xâ ) respectively. â Since ðœð = ðŒâð = 1, this implies that ðœðâ = ðŒâð
ð = 0, 1, . . . , ð
In other words, ð (xâ ) = xâ . 100
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics â=
Let ðŽ0 , ðŽ1 , . . . , ðŽð be closed subsets of an ð dimensional simplex ð with vertices x0 , x1 , . . . , xð such that ⪠conv { xð : ð â ðŒ } â ðŽð ðâðŒ
for every ðŒ â { 0, 1, . . . , ð }. For ð = 0, 1, . . . , ð, let ðð (x) = ð(x, ðŽð ) For any x â ð with barycentric coordinates ðŒ0 , ðŒ1 , . . . , ðŒð , deï¬ne ð (x) = ðœ0 x0 + ðœ1 x1 + â
â
â
+ ðœð xð where ðœð =
ðŒð + ðð (x) â 1 + ðð=0 ðð (x)
ð = 0, 1, . . . , ð
(2.45)
â By construction ðœð ⥠0 and ðð=0 ðœð = 1. Therefore ð (x) â ð. That is, ð : ð â ð. Furthermore ð is continuous. By Brouwerâs theorem, there exists a ï¬xed point ð¥â with ð (xâ ) = xâ . That is ðœðâ = ðŒâð for ð = 0, 1, . . . , ð. Now, since the collection ðŽ0 , ðŽ1 , . . . , ðŽð covers ð, there exists some ð for which ð(xâ , ðŽð ) = 0. Substituting ðœðâ = ðŒâð in (2.45) we have ðŒâð =
1+
ðŒâ âð ð
ð=0
ðð (xâ )
which implies that ðð (xâ ) = 0 for every ð. Since the ðŽð are closed, xâ â ðŽð for every ð and therefore xâ â
ð â©
ðŽð â= â
ð=0
( ) 2.137 To simplify the notation, let ð§ð+ (p) = max 0, zð (p) . Assume pâ is a ï¬xed point of ð. Then for every ð = 1, 2, . . . , ð ðâð =
ðð + ð§ð+ (pâ ) âð 1 + ð=1 ð§ð+ (pâ )
Cross-multiplying ðâð + ðâð
ð â ð=1
ð§ð+ (p) = ðâð + ð§ð+ (pâ )
or ðâð
ð â ð=1
ð§ð+ (p) = ð§ð+ (pâ )
ð = 1, 2, . . . ð
Multiplying each equation by ð§ð (p) we get ðâð ð§ð (pâ )
ð â ð=1
ð§ð+ (p) = ð§ð (pâ )ð§ð+ (pâ ) 101
ð = 1, 2, . . . ð
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Summing over ð ð â
Since
âð
ð=1
ðâð ð§ð (pâ )
ð â ð=1
ð=1
ð§ð+ (p) =
ð â ð=1
ð§ð (pâ )ð§ð+ (pâ )
ðâð ð§ð (pâ ) = 0 this implies that ð â ð=1
ð§ð (pâ )ð§ð+ (pâ ) = 0
( )2 Each term of this sum is nonnegative, since it is either 0 or ð§ð (pâ ) . Consequently, every term must be zero which implies that ð§ð (pâ) †0 for every ð = 1, 2, . . . , ð. In other words, z(pâ ) †0. 2.138 Every individual demand function xð (p, ð) is continuous (Example 2.90) in p and ð. For given endowment ð ð ðð =
ð â
ðð ð ðð
ð=1
is continuous in p (Exercise 2.78). Therefore the excess demand function zð (p) = xð (p, ð) â ð ð is continuous for every consumer ð and hence the aggregate excess demand function is continuous. Similarly, the consumerâs demand function xð (p, ð) is homogeneous of degree 0 in p and ð. For given endowment ð ð , the consumerâs wealth is homogeneous of degree 1 in p and therefore the consumerâs excess demand function zð (p) is homogeneous of degree 0. So therefore is the aggregate excess demand function z(p). 2.139 z(p) = =
ð â ð=1 ð â
zð (p) ( ) xð (p, ð) â ð ð
ð=1
and therefore pð z(p) =
ð â
pð xð (p, ð) â
ð=1
ð â
pð ð ð
ð=1
Since preferences are nonsatiated and strictly convex, they are locally nonsatiated (Exercise 1.248) which implies (Exercise 1.235) that every consumer must satisfy his budget constraint pð xð (p, ð) = pð ðð for every ð = 1, 2, . . . , ð Therefore in aggregate pð z(p) =
ð â
pð xð (p, ð) â
ð=1
ð â ð=1
for every p. 102
pð ð ð = 0
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 2.140 Assume there exists pâ such that z(pâ ) †0. That is z(pâ ) =
ð â
zð (p) =
ð=1
ð ð ð â â ( ) â xð (p, ð) â ðð = xð (p, ð) â ðð †0 ð=1
or
ð=1
â ðâð
xð â€
â
ð=1
ðð
ðâð
Aggregate demand is less or equal to available supply. âð Let ðâð = ð=1 ðâð ð ðð denote the wealth of consumer ð when the price system is pâ and let xâð = x(pâ , ðâ ) be his chosen consumption bundle. Then xâð â¿ xð for every xð â ð(pâ , ðð ) Let xâ = (xâ1 , xâ2 , . . . , xâð ) be the allocation comprising these optimal bundles. The pair (pâ , xâ ) is a competitive equilibrium. 2.141 For each xð , let ð ð denote the subsimplex of Îð which contains xð and let xð0 , xð1 , . . . , xðð denote the vertices of ð ð . Let ðŒð0 , ðŒð1 , . . . , ðŒðð denote the barycentric coordinates (Exercise 1.159) of x with respect to the vertices of ð ð and let yðð = ð ð (xðð ), ð = 0, 1, . . . , ð, denote the images of the vertices. Since ð is compact, there exists â² â² â² subsequences xðð , yðð and ðŒð such that xðð â xâð
yðð â yðâ and ðŒðð â ðŒâð
ð = 0, 1, . . . , ð
Furthermore, ðŒâð ⥠0 and ðŒâ0 +ðŒâ1 +â
â
â
+ðŒâð = 1. Since the diameters of the subsimplices converge to zero, their vertices must converge to the same point. That is, we must have xâ0 = xâ1 = â
â
â
= xâð = xâ By deï¬nition of ð ð ð ð (xð ) = ðŒð0 ð (xð0 ) + ðŒð1 ð (xð1 ) + â
â
â
+ ðŒðð ð (xðð ) Substituting yðð = ð ð (xðð ), ð = 0, 1, . . . , ð and recognizing that xð is a ï¬xed point of ð ð , we have ð¥ð = ð ð (xð ) = ðŒð0 y0ð + ðŒð1 y1ð + â
â
â
+ ðŒðð yðð Taking limits xâ = ðŒâ0 y0â + ðŒâ1 y1â + â
â
â
+ ðŒâð yðâ
(2.46)
For each coordinate ð, (xðð , yðð ) â graph(ð) for every ð = 0, 1, . . . . Since ð is closed, (xâð , yðâ ) â graph(ð). That is, yðâ â ð(xâð ) = ð(xâ ) for every ð = 0, 1, . . . , ð. Therefore, (2.46) implies xâ â conv ð(xâ ) Since ð is convex valued, xâ â ð(xâ )
103
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2.142 Analogous to Exercise 2.129, there exists a simplex ð containing ð and a retraction of ð onto ð, that is a continuous function ð : ð â ð with ð(x) = x for every x â ð. Then ð â ð : ð â ð â ð is closed-valued (Exercise 2.106) and uhc (Exercise 2.103). By the argument in the proof, there exists a point xâ â ð such that xâ â ð â ð(xâ ). However, since ð â ð(xâ ) â ð, we must have xâ â ð and therefore ð(xâ ) = xâ . This implies xâ â ð(xâ ). That is, xâ is a ï¬xed point of ð. 2.143 ðµ = ðµ1 à ðµ2 à . . . à ðµð is the Cartesian product of uhc, compact- and convexvalued correspondences. Therefore ðµ is also compact-valued and uhc (Exercise 2.112 and also convex-valued (Exercise 1.165). By Exercise 2.106, ðµ is closed. 2.144 Strict quasiconcavity ensures that the best response correspondence is in fact a function ðµ : ð â ð. Since the hypotheses of Example 2.96 apply, there exists at least one equilibrium. Suppose that there are two Nash equilibria s and sâ² . Since ðµ is a contraction, ð(ðµ(s), ðµ(sâ² ) †ðœð(s, sâ² ) for some ðœ < 1. However ðµ(s) = s and ðµ(sâ² ) = sâ² and (2.46) implies that ð(s, sâ² ) †ðœð(s, sâ² ) which is possible if and only if s = sâ² . This implies that the equilibrium must be unique. 2.145 Since ðŸ is compact, it is totally bounded (Exercise 1.112). There exists a ï¬nite set of points x1 , x2 , . . . , xð such that ð â©
ðŸâ
ðµð (xð )
ð=1
Let ð = conv {x1 , x2 , . . . , xð }. For ð = 1, 2, . . . , ð and x â ð, deï¬ne ðŒð (x) = max{0, ð â â¥x â xð â¥} Then for every x â ðŸ, 0 †ðŒð (x) †ð,
ð = 1, 2, . . . , ð
and ðŒð (x) > 0 ââ x â ðµð (xð ) Note that ðŒð (x) > 0 for some ð. Deï¬ne
â ðŒð (x)xð â(x) = â ðŒð (x)
Then â(x) â ð and therefore â : ðŸ â ð. Furthermore, â is continuous and â ðŒð (x)xð â â x â¥â(x) â x⥠= ðŒð (x) â ðŒð (x)(xð â x) â = ðŒð (x) â ðŒð (x) â¥xð â x⥠â = ðŒð (x) â ðŒð (x)ð †â =ð ðŒð (x) since ðŒð (x) > 0 ââ â¥xð â x⥠†ð. 104
Solutions for Foundations of Mathematical Economics 2.146
c 2001 Michael Carter â All rights reserved
( ) 1. For every x â ð ð , ð (x) â ð and therefore ð ð (x) = âð ð (x) â ð ð .
2. For any x â ð ð , let y = ð (x) â ð (ð) and therefore ð â (y) â y < 1 ð which implies ð ð (x) â ð (x) †1 for every x â ð ð ð 2.147 By the Triangle inequality ð x â ð (x) †ð ð (xð ) â ð (xð ) + ð (xð ) â ð (x) As shown in the previous exercise ð ð ð (x ) â ð (xð ) †1 â 0 ð as ð â â. Also since ð is continuous ð (xð ) â ð (x) â 0 Therefore ð x â ð (x) â 0 =â x = ð (x) x is a ï¬xed point of ð . 2.148 ð (ð¹ ) is bounded and equicontinuous and so therefore is ð (ð¹ ) (Exercise 2.96). By Ascoliâs theorem (Exercise 2.95), ð (ð¹ ) is compact. Therefore ð is a compact operator. Applying Corollary 2.8.1, ð has a ï¬xed point.
105
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 3: Linear Functions 3.1 Let x1 , x2 â ð and ðŒ1 , ðŒ2 â â. Homogeneity implies that ð (ðŒ1 x1 ) = ðŒ1 ð (ð¥1 ) ð (ðŒ2 x2 ) = ðŒ2 ð (ð¥2 ) and additivity implies ð (ðŒ1 x1 + ðŒ2 x2 ) = ðŒ1 ð (x1 ) + ðŒ2 ð (x2 ) Conversely, assume ð (ðŒ1 x1 + ðŒ2 x2 ) = ðŒ1 ð (x1 ) + ðŒ2 ð (x2 ) for all x1 , x2 â ð and ðŒ1 , ðŒ2 â â. Letting ðŒ1 = ðŒ2 = 1 implies ð (x1 + x2 ) = ð (x1 ) + ð (x2 ) while setting x2 = 0 implies ð (ðŒ1 x1 ) = ðŒ1 ð (x1 ) 3.2 Assume ð1 , ð2 â ð¿(ð, ð ). Deï¬ne the mapping ð1 + ð2 : ð â ð by (ð1 + ð2 )(x) = ð1 (x) + ð2 (x) We have to conï¬rm that ð1 + ð2 is linear, that is (ð1 + ð2 )(x1 + x2 ) = ð1 (x1 + x2 ) + ð2 (x1 + x2 ) = ð1 (x1 ) + ð1 (x2 ) + ð2 (x1 ) + ð2 (x2 ) = ð1 (x1 ) + ð2 (x1 ) + ð1 (x1 ) + ð2 (x2 ) = (ð1 + ð2 )(x1 ) + (ð1 + ð2 )(x2 ) and (ð1 + ð2 )(ðŒx) = ð1 (ðŒx) + ð2 (ðŒx) = ðŒ(ð1 (x) + ð2 (x)) = ðŒ(ð1 + ð2 )(x) Similarly let ð â ð¿(ð, ð ) and deï¬ne ðŒð : ð â ð by (ðŒð )(x) = ðŒð (x) ðŒð is also linear, since (ðŒð )(ðœx) = ðŒð (ðœx) = ðŒðœð (x) = ðœðŒð (x) = ðœ(ðŒð )(x) 106
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.3 Let x, x1 , x2 â â2 . Then ð (x1 + x2 ) = ð (ð¥11 + ð¥21 , ð¥12 + ð¥22 ) ) ( = (ð¥11 + ð¥21 ) cos ð â (ð¥12 + ð¥22 ) sin ð, (ð¥11 + ð¥21 ) sin ð â (ð¥12 + ð¥22 ) cos ð ) ( = (ð¥11 cos ð â ð¥12 sin ð) + (ð¥21 cos ð â ð¥22 sin ð), (ð¥11 sin ð + ð¥12 cos ð) + (ð¥21 sin ð â ð¥22 cos ð) ( ) = (ð¥11 cos ð â ð¥12 sin ð, ð¥11 sin ð + ð¥12 cos ð) + (ð¥21 cos ð â ð¥22 sin ð, ð¥21 sin ð â ð¥22 cos ð = ð (ð¥11 , ð¥12 ) + ð (ð¥21 , ð¥22 ) = ð (x1 ) + ð (x2 ) and ð (ðŒx) = ð (ðŒð¥1 , ðŒð¥2 ) = (ðŒð¥1 cos ð â ðŒð¥2 sin ð, ðŒð¥1 sin ð + ðŒð¥2 cos ð) = ðŒ (ð¥1 cos ð â ð¥1 sin ð, ð¥1 sin ð + ð¥2 cos ð) = ðŒð (ð¥1 , ð¥2 ) = ðŒð (x) 3.4 Let x, x1 , x2 â â3 . ð (x1 + x2 ) = ð (ð¥11 + ð¥2 , ð¥12 + ð¥22 , ð¥13 + ð¥23 ) = (ð¥11 + ð¥21 , ð¥12 + ð¥22 , 0) = (ð¥11 , ð¥12 , 0) + (ð¥21 , ð¥22 , 0) = ð (ð¥11 , ð¥12 , ð¥13 ) + ð (ð¥21 , ð¥22 , ð¥23 ) = ð (x1 ) + ð (x2 ) and ð (ðŒx) = ð (ðŒð¥1 , ðŒð¥2 , ðŒð¥3 ) = (ðŒð¥1 , ðŒð¥2 , 0) = ðŒ(ð¥1 , ð¥2 , 0) = ðŒð (ð¥1 , ð¥2 , ð¥3 ) = ðŒð (x) This mapping is the projection of 3-dimensional space onto the (2-dimensional) plane. 3.5 Applying the deï¬nition
( )( ) 0 1 ð¥1 ð (ð¥1 , ð¥2 ) = 1 0 ð¥2 = (ð¥2 , ð¥1 )
This function interchanges the two coordinates of any point in the plane â2 . Its action corresponds to reï¬ection about the line ð¥1 = ð¥2 ( 45 degree diagonal). 3.6 Assume (ð, ð€) and (ð, ð€â² ) are two games in ð¢ ð . For any coalition ð â ð (ð€ + ð€â² )(ð) â (ð€ + ð€â² )(ð â {ð}) = ð€(ð) + ð€(ð â² ) â ð€(ð â {ð}) â ð€â² (ð â {ð}) = (ð€(ð) â ð€(ð â {ð})) + (ð€â² (ð) â ð€â² (ð â {ð})) = ðð (ð€) + ðð (ð€â² ) 107
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.7 The characteristic function of cost allocation game is ð€(ðŽð ) = 0 ð€(ð ð ) = 0
ð€(ðŽð, ð ð ) = 210 ð€(ðŽð, ðŸð ) = 770
ð€(ðŸð ) = 0
ð€(ðŸð, ð ð ) = 1170
ð€(ð ) = 1530
The following table details the computation of the Shapley value for player ðŽð . ð ðŽð ðŽð, ð ð ðŽð, ðŸð ðŽð, ð ð, ðŸð ðð (ð€)
ðŸð 1/3 1/6 1/6 1/3
ð€(ð) 0 210 770 1530
ð€(ð â {ð}) 0 0 0 1170
ðŸð (ð€(ð) â ð€(ð â {ð})) 0 35 128 1/3 120 283 1/3
Thus ððŽð ð€ = 283 1/3. Similarly, we can calculate that ðð ð ð€ = 483 1/3 and ððŸð ð€ = 763 1/3. 3.8 â
ðð ð€ =
ðâð
=
â
( â
ðâð
ðâð
â
(
ðâð
=
â
) ðŸð (ð€(ð) â ð€(ð â {ð})) ) ðŸð (ð€(ð) â ð€(ð â {ð}))
ðâð
ââ
ðŸð ð€(ð) â
ðâð ðâð
=
â â
ðŸð ð€(ð â {ð})
ðâð ðâð
ð à ðŸð ð€(ð) â
ðâð
=
ââ
(
â
ðŸð
ðâð
ð à ðŸð ð€(ð) â
â
â
) ð€(ð â {ð})
ðâð
ð à ðŸð ð€(ð)
ðâð
ðâð
= ð à ðŸð ð€(ð ) = ð€(ð ) 3.9 If ð, ð â ð ð€(ð â {ð}) = ð€(ð â {ð, ð} ⪠{ð}) = ð€(ð â {ð, ð} ⪠{ð}) = ð€(ð â {ð}) ðð (ð€) =
â
ðŸð (ð€(ð) â ð€(ð â {ð}))
ðâð
=
â
ðŸð (ð€(ð) â ð€(ð â {ð})) +
ðâð,ð
=
â
â
ðŸð (ð€(ð) â ð€(ð â {ð})) +
=
â
ðŸð (ð€(ð ⪠{ð}) â ð€(ð))
ðââð,ð
ðŸð (ð€(ð) â ð€(ð â {ð})) +
â
ðŸð â² (ð€(ð Ⲡ⪠{ð}) â ð€(ð â² ))
ð â² ââð,ð
ðâð,ð
â
ðŸð (ð€(ð) â ð€(ð â {ð}))
ðâð,ðââð
ðâð,ð
=
â
ðŸð (ð€(ð) â ð€(ð â {ð})) +
ðâð,ð
â
ðââð,ðâð
= ðð (ð€) 108
ðŸð (ð€(ð) â ð€(ð â {ð}))
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.10 For any null player ð€(ð) â ð€(ð â {ð}) = 0 for every ð â ð . Consequently
â
ðð (ð€) =
ðŸð (ð€(ð) â ð€(ð â {ð})) = 0
ðâð
3.11 Every ð â / ð is a null player, so that ðð (ð¢ð ) = 0 Feasibility requires that
â
for every ð â /ð â
ðð (ð¢ð ) =
ðâð
ðð (ð¢ð ) = 1
ðâð
Further, any two players in ð are substitutes, so that symmetry requires that ðð (ð¢ð ) = ðð (ð¢ð )
for every ð, ð â ð
Together, these conditions require that ðð (ð¢ð ) =
1 ð¡
for every ð â ð
The Shapley value of the a T-unanimity game is { 1 ðâð ðð (ð¢ð ) = ð¡ 0 ðâ /ð where ð¡ = â£ð â£. 3.12 Any coalitional game can be represented as a linear combination of unanimity games ð¢ð (Example 1.75) â ð€= ðŒð ð¢ð ð
By linearity, the Shapley value is
â
â
ðð€ = ð â =
â
â ðŒð ð¢ð â
ð âð
ðŒð ðð¢ð
ð âð
and therefore for player ð ðð ð€ =
â
ðŒð ðð ð¢ð
ð âð
=
â 1 ðŒð ð¡
ð âð ð âð
â 1 â 1 ðŒð â ðŒð = ð¡ ð¡ ð âð
ð âð ðâð /
= ð (ð, ð€) â ð (ð â {ð}, ð€) 109
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Using Exercise 3.8 ð€(ð ) =
â
ðð ð€
ðâð
=
â(
) ð (ð, ð€) â ð (ð â {ð}, ð£)
ðâð
= ðð (ð, ð€) â
â
ð (ð â {ð}, ð£)
ðâð
which implies that 1 ð (ð, ð€) = ð
( ð€(ð ) â
â
) ð (ð â {ð}, ð£)
ðâð
3.13 Choose any x â= 0 â ð. 0ð = x â x and by additivity ð (0ð ) = ð (x â x) = ð (x) â ð (x) = 0ð 3.14 Let x1 , x2 belong to ð. Then ð â ð (x1 + x2 ) = ð â ð (x1 + x2 ) ) ( = ð ð (x1 ) + ð (x2 ) ) ( ) ( = ð ð (x1 ) + ð ð (x2 ) = ð â ð (x1 ) + ð â ð (x2 ) and ð â ð (ðŒx) = ð (ð (ðŒx)) = ð (ðŒð (x)) = ðŒð (ð (x)) = ðŒð â ð (x) Therefore ð â ð is linear. 3.15 Let ð be a subspace of ð and let y1 , y2 belong to ð (ð). Choose any x1 â ð â1 (y1 ) and x2 â ð â1 (y2 ). Then for ðŒ1 , ðŒ2 â â ðŒ1 x1 + ðŒ2 x2 â ð Since ð is linear (Exercise 3.1) ðŒ1 y1 + ðŒ2 y2 = ðŒ1 ð (x1 ) + ðŒ2 ð (x2 ) = ð (ðŒ1 x1 + ðŒ2 x2 ) â ð (ð) ð (ð) is a subspace. Let ð be a subspace of ð and let x1 , x2 belong to ð â1 (ð ). Let y1 = ð (x1 ) and y2 = ð (x2 ). Then y1 , y2 â ð . For every ðŒ1 , ðŒ2 â â ðŒ1 y1 + ðŒ2 y2 = ðŒ1 ð (x1 ) + ðŒ2 ð (x2 ) â ð 110
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Since ð is linear, this implies that ð (ðŒ1 x1 + ðŒ2 x2 ) = ðŒ1 ð (x1 ) + ðŒ2 ð (x2 ) â ð Therefore ðŒ1 x1 + ðŒ2 x2 â ð â1 (ð ) We conclude that ð â1 (ð ) is a subspace. 3.16 ð (ð) is a subspace of ð . rank ð (ð) = rank ð implies that ð (ð) = ð . ð is onto. 3.17 This is a special case of the previous exercise, since {0ð } is a subspace of ð . 3.18 Assume not. That is, assume that there exist two distinct elements x1 and x2 with ð (x1 ) = ð (x2 ). Then x1 â x2 â= 0ð but ð (x1 â x2 ) = ð (x1 ) â ð (x2 ) = 0ð so that x1 â x2 â kernel ð which contradicts the assumption that kernel ð = {0}. 3.19 If ð has an inverse, then it is one-to-one and onto (Exercise 2.4), that is ð â1 (0) = 0 and ð (ð) = ð . Conversely, if kernel ð = {0} then ð is one-to-one by the previous exercise. If furthermore ð (ð) = ð , then ð is one-to-one and onto, and therefore has an inverse (Exercise 2.4). 3.20 Let ð be a nonsingular linear function from ð to ð with inverse ð â1 . Choose y1 , y2 â ð and let x1 = ð â1 (y1 ) x2 = ð â1 (y2 ) so that y1 = ð (x1 ) y2 = ð (x2 ) Since ð is linear ð (x1 + x2 ) = ð (x1 ) + ð (x2 ) = y1 + y2 which implies that ð â1 (y1 + y2 ) = x1 + x2 = ð â1 (y1 ) + ð â1 (y2 ) The homogeneity of ð â1 can be demonstrated similarly. 3.21 Assume that ð : ð â ð and ð : ð â ð are nonsingular. Then (Exercise 3.19) â ð (ð) = ð and ð(ð ) = ð â kernel ð = {0ð } and kernel ð = {0ð } We have previously shown (Exercise 3.14) that â = ð â ð : ð â ð is linear. To show that â is nonsingular, we note that â â(ð) = ð â ð (ð) = ð(ð ) = ð
111
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
â If x â kernel (â) then â(x) = ð (ð (x)) = 0 and ð (x) â kernel ð = {0ð }. Therefore ð (x) = 0ð which implies that x = 0ð . Thus kernel â = {0ð }. We conclude that â is nonsingular. Finally, let z be any point in ð and let x1 = ââ1 (z) = (ð â ð )â1 (z) y = ð â1 (z) x2 = ð â1 (y) Then z = â(x1 ) = ð â ð (x1 ) z = ð(y) = ð â ð (x2 ) which implies that x1 = x2 . 3.22 Suppose ð were one-to-one. Then kernel ð = {0} â kernel â and ð = â â ð â1 is a well-deï¬ned linear function mapping ð (ð) to ð with ( ) ð â ð = â â ð â1 â ð = â We need to show that this still holds if ð is not one-to-one. In this case, for arbitrary y â ð (ð), ð â1 (y) may contain more than one element. Suppose x1 and x2 are distinct elements in ð â1 (y). Then ð (x1 â x2 ) = ð (x1 ) â ð (x2 ) = y â y = 0 so that x1 â x2 â kernel ð â kernel â (by assumption). Therefore â(x1 ) â â(x2 ) = â(x1 â x2 ) = 0 which implies that â(x1 ) = â(x2 ) for all x1 , x2 â ð â1 (y). Thus ð = ââ ð â1 : ð (ð) â ð is well deï¬ned even if ð is many-to-one. To show that ð is linear, choose y1 , y2 in ð (ð) and let x1 â ð â1 (y1 ) x2 â ð â1 (y2 ) Since ð (x1 + x2 ) = ð (x1 ) + ð (x2 ) = y1 + y2 x1 + x2 â ð â1 (y1 + y2 ) and ð(y1 + y2 ) = â(x1 + x2 ) Therefore ð(y1 ) + ð(y2 ) = â(x1 ) + â(x2 ) = â(x1 + x2 ) = ð(y1 + y2 ) 112
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Similarly ðŒx1 â ð â1 (ðŒy1 ) and ð(ðŒy1 ) = â(ðŒx1 ) = ðŒâ(x1 ) = ðŒð(y1 ) We conclude that ð = â â ð â1 is a linear function mapping ð (ð) to ð with â = ð â ð . 3.23 Let y be an arbitrary element of ð (ð) with x â ð â1 (y). Since B is a basis for ð, x can be represented as a linear combination of elements of ðµ, that is there exists x1 , x2 , .., xð â ðµ and ðŒ1 , ..., ðŒð â ð
such that x=
ð â
ðŒð xð
ð=1
y = ð (x) ) ( â ðŒð xð =ð =
â
ð
ðŒð ð (xð )
ð
Since ð (xð ) â ð (ðµ), we have shown that y can be written as a linear combination of elements of ð (ðµ), that is y â lin ðµ Since the choice of y was arbitrary, ð (ðµ) spans ð (ð), that is lin ðµ = ð (ð) 3.24 Let ð = dim ð and ð = dim kernel ð . Let x1 , . . . , xð be a basis for the kernel of ð . This can be extended (Exercise 1.142) to a basis ðµ for ð. Exercise 3.23 showed lin ðµ = ð (ð) Since x1 , x2 , . . . , xð â kernel ð , ð (xð ) = 0 for ð = 1, 2, . . . , ð. This implies that {ð (xð+1 ), . . . , ð (xð )} spans ð (ð), that is lin {(xð+1 ), ..., ð (xð )} = ð (ð) To show that dim ð (ð) = ð â ð, we have to show that {ð (ð¥ð+1 ), ð (ð¥ð+2 ), . . . , ð (ð¥ð )} is linearly independent. Assume not. That is, assume there exist ðŒð+1 , ðŒð+2 , ..., ðŒð â ð
such that ð â
ðŒð ð (xð ) = 0
ð=ð+1
This implies that
( ð
)
ð â
ðŒð xð
=0
ð=ð+1
or x=
ð â
ðŒð ð¥ð â kernel ð
ð=ð+1
113
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
This implies that x can also be expressed as a linear combination of elements in {x1 , ð¥2 , ..., xð }, that is there exist scalars ðŒ1 , ðŒ2 , . . . , ðŒð such that x=
ð â
ðŒð xð
ð=1
or x=
ð â
ð â
ðŒð xð =
ð=1
ðŒð xð
ð=ð+1
which contradicts the assumption that ðµ is a basis for ð. Therefore {ð (xð+1 ), . . . , ð (xð )} is a basis for ð (ð) and therefore dim ð (ð¥) = ð â ð. We conclude that dim kernel ð + dim ð (ð) = ð = dim ð 3.25 Equation (3.2) implies that nullity ð = 0, and therefore ð is one-to-one (Exercise 3.18). 3.26 Choose some x = (ð¥1 , ð¥2 , . . . , ð¥ð ) â ð. x has a unique representation in terms of the standard basis (Example 1.79) x=
ð â
ð¥ð eð
ð=1
Let y = ð (x). Since ð is linear
â
y = ð (x) = ð â
â
ð â
ð¥ð eð â =
ð=1
ð â
xð ð (eð )
ð=1
Each ð (eð ) has a unique representation of the form ð (eð ) =
ð â
ððð eð
ð=1
so that y = ð (x) =
ð â ð=1
=
ð â ð=1
(
ð¥ð
ð â
) ððð eð
ð=1
â â ð â â ððð ð¥ð â eð ð=1
â â âð ð1ð ð¥ð âð=1 ð â ð=1 ð2ð ð¥ð â â â =â â .. â â âð . ð ð¥ ð=1 ðð ð = ðŽx where
â
â ð11 ð12 . . . ð1ð â ð21 ð22 . . . ð2ð â â ðŽ=â â . . . . . . . . . . . . . . . . . . . . .â ðð1 ðð2 . . . ððð 114
Solutions for Foundations of Mathematical Economics 3.27
( 1 0 0 1
c 2001 Michael Carter â All rights reserved
) 0 0
3.28 We must specify bases for each space. The most convenient basis for ðºð is the T-unanimity games. We adopt the standard basis for âð . With respect to these bases, the Shapley value ð is represented by the 2ðâ1 Ãð matrix where each row is the Shapley value of the corresponding T-unanimity game. For three player games (ð = 3), the matrix is â 1 0 â0 1 â â0 0 â1 1 â â 21 2 â â 2 01 â0 2 1 3
1 3
â 0 0â â 1â â 0â â 1â 2â 1â 2 1 3
3.29 Clearly, if ð is continuous, ð is continuous at 0. To show the converse, assume that ð : ð â ð is continuous at 0. Let (xð ) be a sequence which converges to x â ð. Then the sequence (xð â x) converges to 0ð and therefore ð (xð âx) â 0ð by continuity (Exercise 2.68). By linearity, ð (xð )âð (x) = ð (xð âx) â 0ð and therefore ð (xð ) converges to ð (x). We conclude that ð is continuous at x. 3.30 Assume that ð is bounded, that is â¥ð (x)⥠†ð â¥x⥠for every x â ð Then ð is Lipschitz at 0 (with Lipschitz constant ð ) and hence continuous (by the previous exercise). Conversely, assume ð is continuous but not bounded. Then, for every positive integer ð, there exists some xð â ð such that â¥ð (xð )⥠> ð â¥xð ⥠which implies that ( ) xð ð >1 ð â¥xð ⥠Deï¬ne yð =
xð ð â¥xð â¥
Then yð â 0 but ð (yð ) ââ 0. This implies that ð is not continuous at the origin, contradicting our hypothesis. 3.31 Let {x1 , x2 , . . . , xð } be a basis for ð. For every x â ð, there exists numbers ðŒ1 , ðŒ2 , . . . , ðŒð such that x=
ð â
ðŒð xð
ð=1
115
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics and ð (x) =
ð â
ðŒð ð (xð )
ð=1 ð â
â¥ð (x)⥠= â€
ð=1
ð â
ðŒð ð (xð )
â£ðŒð ⣠â¥ð (xð )â¥
ð=1
ð )â ( ð â£ðŒð ⣠†max â¥ð (xð )⥠ð=1
ð=1
By Lemma 1.1, there exists a constant ð such that ð ð 1 â 1 â â£ðŒð ⣠†ðŒð xð = â¥x⥠ð ð ð=1 ð=1 Combining these two inequalities â¥ð (x)⥠†ð â¥x⥠where ð = maxðð=1 â¥ð (xð )⥠/ð. 3.32 For any x â ð, let ð = â¥x⥠and deï¬ne y = x/ð. Linearity implies that â¥ð (x)⥠= sup â¥ð (x/ð)⥠= sup â¥ð (y)⥠ð xâ=0 xâ=0 â¥yâ¥=1
â¥ð ⥠= sup
3.33 â¥ð ⥠is a norm Let ð â ðµð¿(ð, ð ). Clearly â¥ð ⥠= sup â¥ð (x)⥠⥠0 â¥xâ¥=1
Further, for every ðŒ â â, â¥ðŒð ⥠= sup â¥ðŒð (x)⥠= â£ðŒâ£ â¥ð ⥠â¥xâ¥=1
Finally, for every ð â ðµð¿(ð, ð ), â¥ð + ð⥠= sup â¥ð (x) + ð(x)⥠†sup â¥ð (x)⥠+ sup â¥ð(x)⥠†â¥ð ⥠+ â¥ð⥠â¥xâ¥=1
â¥xâ¥=1
â¥xâ¥=1
verifying the triangle inequality. There â¥ð ⥠is a norm. ðµð¿(ð, ð ) is a linear space Let ð, ð â ðµð¿(ð, ð ). Since ðµð¿(ð, ð ) â ð¿(ð, ð ), ð + ð is linear, that is ð +ð â ð¿(ð, ð ) (Exercise 3.2). Similarly, ðŒð â ð¿(ð, ð ) for every ðŒ â â. Further, by the triangle inequality â¥ð + ð⥠†â¥ð ⥠+ â¥ð⥠and therefore for every x â ð â¥(ð + ð)(x)⥠†â¥ð + ð⥠â¥x⥠†(â¥ð ⥠+ â¥ðâ¥) â¥x⥠Therefore ð + ð â ðµð¿(ð, ð ). Similarly â¥(ðŒð )(x)⥠†(â£ðŒâ£ â¥ð â¥) â¥x⥠so that ðŒð â ðµð¿(ð, ð ). 116
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ðµð¿(ð, ð ) is complete with this norm Let (ð ð ) be a Cauchy sequence in ðµð¿(ð, ð ). For every x â ð â¥ð ð (x) â ð ð (x)⥠†â¥ð ð â ð ð ⥠â¥x⥠Therefore (ð ð (x)) is a Cauchy sequence in ð , which converges since ð is complete. Deï¬ne the function ð : ð â ð by ð (x) = limðââ ð ð (x). ð is linear since ð (x1 + x2 ) = lim ð ð (x1 + x2 ) = lim ð ð (x1 ) + lim ð ð (x2 ) = ð (x1 ) + ð (x2 ) and ð (ðŒx) = lim ð ð (ðŒx) = ðŒ lim ð ð (x) = ðŒð (x) To show that ð is bounded, we observe that â¥ð (x)⥠= lim ð ð (x) = lim â¥ð ð (x)⥠†sup â¥ð ð (x)⥠†sup â¥ð ð ⥠â¥x⥠ð
ð
ð
ð
Since (ð ð ) is a Cauchy sequence, (ð ð ) is bounded (Exercise 1.100), that is there exists ð such that â¥ð ð ⥠†ð . This implies â¥ð (x)⥠†sup â¥ð ð ⥠â¥x⥠†ð â¥x⥠ð
Thus, ð is bounded. To complete the proof, we must show ð ð â ð , that is â¥ð ð â ð ⥠â 0. Since (ð ð ) is a Cauchy sequence, for every ð > 0, there exists ð such that â¥ð ð â ð ð ⥠†ð for every ð, ð ⥠ð and consequently â¥ð ð (x) â ð ð (x)⥠= â¥(ð ð â ð ð )(x)⥠†ð â¥x⥠Letting ð go to inï¬nity, â¥ð ð (x) â ð (x)⥠= â¥(ð ð â ð )(x)⥠†ð â¥x⥠for every x â ð and ð ⥠ð and therefore â¥ð ð â ð ⥠= sup {ð ð â ð )(x)} †ð â¥xâ¥=1
for every ð ⥠ð . 3.34
1. Since ð is ï¬nite-dimensional, ð is compact (Proposition 1.4). Since ð is continuous, ð (ð) is a compact set in ð (Exercise 2.3). Since 0ð â / ð, 0ð = ð (0ð ) â / ð (ð). ( )ð 2. Consequently, ð (ð) is an open set containing 0ð . It contains an open ball ( )ð ð â ð (ð) around 0ð . 3. Let y â ð and choose any x â ð â1 (y) and consider y/ â¥xâ¥. Since ð is linear, ( ) x ð (x) y = =ð â ð (ð) â¥x⥠â¥x⥠â¥x⥠and therefore y/ â¥x⥠â / ð since ð â© ð (ð) = â
. 117
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Suppose that y â / ð (ðµ). Then â¥x⥠⥠1 and therefore y â ð =â
y âð â¥xâ¥
since ð is convex. This contradiction establishes that y â ð (ðµ) and therefore ð â ð (ðµ). We conclude that ð (ðµ) contains an open ball around 0ð . 4. Let ð be any open set in ð. We need to show that ð (ð) is open in ð . Choose any y â ð (ð) and x â ð â1 (y). Then x â ð and, since ð is open, there exists some ð > 0 such that ðµð (x) â ð. Now ðµð (x) = x + ððµ and ð (ðµð (x)) = y + ðð (ðµ) â ð (ð) by linearity. As we have just shown, there exists an open ball T about 0ð such that ð â ð (ðµ). Let ð (x) = y + ðð . ð (x) is an open ball about y. Since ð â ð (ðµ), ð (x) = y + ðð â ð (ðµð (x)) â ð (ð). This implies that ð (ð) is open. Since ð was an arbitrary open set, ð is an open map. 5. Exercise 2.69. 3.35 ð is linear ð (ðŒ + ðœ) =
ð â
(ðŒð + ðœð )xð =
ð=1
ð â
ðŒð xð +
ð=1
ð â
ðœð xð = ð (ðŒ) + ð (ðœ)
ð=1
Similarly for every ð¡ â â ð (ð¡ðŒ) = ð¡
ð â
ðŒð xð = ð¡ð (ð¡ðŒ)
ð=1
ð is one-to-one Exercise 1.137. ð is onto By deï¬nition of a basis lin {x1 , x2 , . . . , xð } = ð ð is continuous Exercise 3.31 ð is an open map Proposition 3.2 3.36 ð is bounded and therefore there exists ð such that â¥ð (x)⥠†ð â¥xâ¥. Similarly, ð â1 is bounded and therefore there exists ð such that for every x ð â1 (y) â€
1 â¥y⥠ð
where y = ð (x). This implies ð â¥x⥠†â¥ð (x)⥠and therefore for every x â ð. ð â¥x⥠†â¥ð (x)⥠†ð â¥x⥠By the linearity of ð , ð â¥x1 â x2 ⥠†â¥ð (x1 â x2 )⥠= â¥ð (x1 ) â ð (x2 )⥠†ð â¥x1 â x2 â¥
118
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.37 For any function, continuity implies closed graph (Exercise 2.70). To show the converse, assume that ðº = graph(ð ) is closed. ð Ãð with norm â¥(x, y)⥠= max{â¥x⥠, â¥yâ¥} is a Banach space (Exercise 1.209). Since ðº is closed, ðº is complete. Also, ðº is a subspace of ð à ð . Consequently, ðº is a Banach space in its own right. Consider the projection â : ðº â ð deï¬ned by â(x, ð (x)) = x. Clearly â is linear, one-to-one and onto with ââ1 (x) = (x, ð (x)) It is also bounded since â¥â(x, ð (x))⥠= â¥x⥠†â¥(x, ð (x)⥠By the open mapping theorem, ââ1 is bounded. For every x â ð â¥ð (x)⥠†â¥(x, ð (x))⥠= ââ1 (x) †ââ1 â¥x⥠We conclude that ð is bounded and hence continuous. 3.38 ð (1) = 5, ð (2) = 7 but ð (1 + 2) = ð (3) = 9 â= ð (1) + ð (2) Similarly ð (3 à 2) = ð (6) = 15 â= 3 à ð (2) 3.39 Assume ð is aï¬ne. Let y = ð (0) and deï¬ne ð(x) = ð (x) â y ð is homogeneous since for every ðŒ â â ð(ðŒx) = ð(ðŒx + (1 â ðŒ)0) = ð (ðŒx + (1 â ðŒ)0) â y = ðŒð (x) + (1 â ðŒ)ð (0) â y = ðŒð (x) + (1 â ðŒ)y â y = ðŒð (x) â ðŒy = ðŒ(ð (ð¥) â y) = ðŒð(x) Similarly for any x1 , x2 â ð ð(ðŒx1 + (1 â ðŒ)x2 ) = ð (ðŒx1 + (1 â ðŒ)x2 ) â ðŠ = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) â ðŠ Therefore, for ðŒ = 1/2 1 1 1 1 ð( x1 + x2 ) = ð (x1 ) + ð (x2 ) â ðŠ 2 2 2 2 1 1 = (ð (x1 ) â ðŠ) + (ð (x2 ) â ðŠ) 2 2 1 1 = ð(x1 ) + ð(x2 ) 2 2 119
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Since ð is homogeneous ð(x1 + x2 ) = ð(x1 ) + ð(x2 ) which shows that ð is additive and hence linear. Conversely if ð (x) = ð(x) + y with ð linear ð (ðŒx1 + (1 â ðŒ)x2 ) = ðŒð(x1 ) + (1 â ðŒ)ð(x2 ) + ðŠ = ðŒð(x1 ) + ðŠ + (1 â ðŒ)ð(x2 ) + ðŠ = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) 3.40 Let ð be an aï¬ne subset of ð and let y1 , y2 belong to ð (ð). Choose any x1 â ð â1 (y1 ) and x2 â ð â1 (y2 ). Then for any ðŒ â â ðŒx1 + (1 â ðŒ)x2 â ð Since ð is aï¬ne ðŒy1 + (1 â ðŒ)y2 = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) = ð (ðŒx1 + (1 â ðŒ)x2 ) â ð (ð) ð (ð) is an aï¬ne set. Let ð be an aï¬ne subset of ð and let x1 , x2 belong to ð â1 (ð ). Let y1 = ð (x1 ) and y2 = ð (x2 ). Then y1 , y2 â ð . For every ðŒ â â ðŒy1 + (1 â ðŒ)y2 = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) â ð Since ð is aï¬ne, this implies that ð (ðŒx1 + (1 â ðŒ)x2 ) = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) â ð Therefore ðŒx1 + (1 â ðŒ)x2 â ð â1 (ð ) We conclude that ð â1 (ð ) is an aï¬ne set. 3.41 For any y1 , y2 â ð (ð), choose x1 , x2 â ð such that yð = ð (xð ). Since ð is convex, ðŒx1 + (1 â ðŒ)x2 â ð and therefore ð (ðŒx1 + (1 â ðŒ)x2 ) = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) = ðŒy1 + (1 â ðŒ)y2 â ð (ð) Therefore ð (ð) is convex. 3.42 Suppose otherwise that y is not eï¬cient. Then there exists another production plan yâ² â ð such that yⲠ⥠y. Since p > 0, this implies that pyâ² > py, contradicting the assumption that y maximizes proï¬t. 3.43 The random variable ð can be represented as the sum â ð(ð )ð{ð } ð= ð âð
120
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
where ð{ð } is the indicator function of the set {ð }. Since ðž is linear ðž(ð) =
â
ð(ð )ðž(ð{ð } )
ð âð
=
â
ðð ð(ð )
ð âð
since ðž(ð{ð } = ð ({ð }) = ðð ⥠0. For the random variable ð = 1, ð(ð ) = 1 for every ð â ð and â ðð = 1 ðž(1) = ð âð
3.44 Let ð¥1 , ð¥2 â ð¶[0, 1]. Recall that addition in C[0,1] is deï¬ned by (ð¥1 + ð¥2 )(ð¡) = ð¥1 (ð¡) + ð¥2 (ð¡) Therefore ð (ð¥1 + ð¥2 ) = (ð¥1 + ð¥2 )(1/2) = ð¥1 (1/2) + ð¥2 (1/2) = ð (ð¥1 ) + ð (ð¥2 ) Similarly ð (ðŒð¥1 ) = (ðŒð¥1 )(1/2) = ðŒð¥1 (1/2) = ðŒð (ð¥1 ) 3.45 Assume that xâ = xâ1 + xâ2 + â
â
â
+ xâð maximizes ð over ð. Suppose to the contrary that there exists yð â ðð such that ð (yð ) > ð (xâð ). Then y = xâ1 + xâ2 + â
â
â
+ yð + â
â
â
+ xâð â ð and â â ð (y) = ð (xâð ) + ð (yð ) > ð (xâð ) = ð (xâ ) ð
ðâ=ð
contradicting the assumption at ð is maximized at xâ . Conversely, assume ð (xâð ) ⥠ð (xð ) for every xð â ðð for every ð = 1, 2, . . . , ð. Summing â â â â ð (xâ ) = ð ( xâð ) = ð (ð¥âð ) ⥠ð (xð ) = ð ( xð ) = ð (x) for every x â ð xâ = xâ1 + xâ2 + â
â
â
+ xâð maximizes ð over ð. 3.46
1. Assume (ð¥ð¡ ) is a sequence in ð1 with ð = the sequence of partial sums ð ð¡ =
ð¡ â
ââ
ð¡=1
â£ð¥ð ⣠< â. Let (ð ð¡ ) denote
â£ð¥ð â£
ð=1
Then (ð ð¡ ) is a bounded monotone sequence in âð which converges to ð . Consequently, (ð ð¡ ) is a Cauchy sequence. For every ð > 0 there exists an ð such that ð+ð â
â£ð¥ð¡ ⣠< ð
ð=ð
121
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
for every ð ⥠ð and ð ⥠0. Letting ð = 0 â£ð¥ð¡ ⣠< ð for every ð ⥠ð We conclude that ð¥ð¡ â 0 so that (ð¥ð¡ ) â ð0 . This establishes ð1 â ð0 . To see that the inclusion is strict, that is ð1 â ð0 , we observe that the sequence (1/ð) = (1, 1/2, 1/3, . . . ) converges to zero but that since â â 1 = 1 + 1 + 1+ = â ð 2 3 ð=1 (1/ð) â / ð1 . Every convergent sequence is bounded (Exercise 1.97). Therefore ð0 â ðâ . 2. Clearly, every sequence (ðð¡ ) â ð1 deï¬nes a linear functional ð â ðâ²0 given by ð (x) =
â â
ðð¡ ð¥ð¡
ð=1
for every x = (ð¥ð¡ ) â ð0 . To show that ð is bounded we observe that every (ð¥ð¡ ) â ð0 is bounded and consequently â£ð (x)⣠â€
â â ð=1
â£ðð¡ ⣠â£ð¥ð¡ ⣠†â¥(ð¥ð¡ )â¥â
â â ð=1
â£ðð¡ ⣠= â¥(ðð¡ )â¥1 â¥(ð¥ð¡ )â¥â
Therefore ð â ðâ0 . To show the converse, let eð¡ denote the unit sequences e1 = (1, 0, 0, . . . ) e2 = (0, 1, 0, . . . ) e3 = (0, 0, 1, . . . ) {e1 , e2 , e3 , . . . , } form a basis for ð0 . Then every sequence (ð¥ð¡ ) â ð0 has a unique representation (ð¥ð¡ ) =
â â
ð¥ð¡ eð¡
ð=1
Let ð â ðâ0 be a continuous linear functional on ð0 . By continuity and linearity ð (x) =
â â
ð¥ð¡ ð (eð¡ )
ð=1
Let ðð¡ = ð (eð¡ ) so that ð (x) =
â â
ðð¡ ð¥ð¡
ð=1
Every linear function is determined by its action on a basis (Exercise 3.23). 122
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
We need to show that the sequence (ðð¡ ) â ð1 . For any ð , consider the sequence xð¡ = (ð¥1 , ð¥2 , . . . , ð¥ð¡ , 0, 0, . . . ) where ⧠âš0 ðð¡ = 0 or ð ⥠ð ð¥ð¡ = â£ðð¡ ⣠⩠otherwise ðð¡ Then (xð¡ ) â ð0 , â¥xð¡ â¥â = 1 and ð (xð¡ ) =
ð¡ â
ðð¡ ð¥ð¡ =
ð=1
ð¡ â
â£ðð¡ â£
ð=1
Since ð â ðâ0 , ð is bounded and therefore ð (xð¡ ) †â¥ð ⥠â¥xð¡ ⥠= â¥ð ⥠< â and therefore ð¡ â
â£ðð¡ ⣠< â for every ð = 1, 2, . . .
ð=1
Consequently â â
â£ðð¡ ⣠= sup
ð¡ â
ð ð=1
ð=1
â£ðð¡ ⣠†â¥ð ⥠< â
We conclude that (ðð¡ ) â ð1 and therefore ðâ0 = ð1 3. Similarly, every sequence (ðð¡ ) â ðâ deï¬nes a linear functional ð on ð1 given by ð (x) =
â â
ðð¡ ð¥ð¡
ð=1
for every x = (ð¥ð¡ ) â ð1 . Moreover ð is bounded since â£ð (x)⣠â€
â â
â£ðð¡ ⣠â£ð¥ð¡ ⣠†â¥(ðð¡ )â¥
ð=1
â â
â£ð¥ð¡ ⣠< â
ð=1
for every x = (ð¥ð¡ ) â ð1 Again, given any linear functional ð â ð1â , let ðð¡ = ð (eð¡ ) where eð¡ is the ð unit sequence. Then ð has the representation ð (x) =
â â
ðð¡ ð¥ð¡
ð=1
To show that (ðð¡ ) â ðâ , for ð = 1, 2, . . . , consider the sequence xð¡ = (0, 0, . . . , ð¥ð¡ , 0, 0, . . . ) where ⧠⚠â£ðð¡ ⣠ð = ð and ð â= 0 ð¡ ðð¡ ð¥ð¡ = â© 0 otherwise Then xð¡ â ð1 , â¥xð¡ â¥1 = 1 and ð (xð¡ ) = â£ðð¡ ⣠Since ð â
ð1â ,
ð is bounded and therefore ð ð = ð (xð ) †â¥ð ⥠â¥xð ⥠= â¥ð â¥
for every ð . Consequently (ðð ) â ðâ . We conclude that ð1â = ðâ 123
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.47 By linearity ð(ð¥, ð¡) = ð(ð¥, 0) + ð(0, ð¡) = ð(ð¥, 0) + ð(0, 1)ð¡
Considered as a function of ð¥, ð(ð¥, 0) is a linear functional on ð. Deï¬ne ð(ð¥) = ð(ð¥, 0) ðŒ = ð(0, 1) Then ð(ð¥, ð¡) = ð(ð¥) + ðŒð¡ 3.48 Suppose ð â©
kernel ðð â kernel ð
ð=1
Deï¬ne the function ðº : ð â âð by ðº(x) = (ð1 (x), ð2 (x), . . . , ðð (x)) Then kernel ðº = { x â ð : ðð (x) = 0, ð = 1, 2, . . . ð } ð â© = kernel ðð ð=1
â kernel ð ð : ð â â and ðº : ð â âð . By Exercise 3.22, there exists a linear function ð» : âð â â such that ð = ð» â ðº. That is, for every ð¥ â ð ð (x) = ð» â ðº(x) = ð»(ð1 (x), ð2 (x), . . . , ðð (x)) Let ðŒð = ð»(eð ) where eð is the ð-th unit vector in âð . Since every linear mapping is determined by its action on a basis, we must have ð (x) = ðŒ1 ð1 (x) + ðŒ2 ð2 (x) + â
â
â
+ ðŒð ðð (x)
for every ð¥ â ð
That is ð â lin ð1 , ð2 , . . . , ðð Conversely, suppose ð â lin ð1 , ð2 , . . . , ðð That is ð (x) = ðŒ1 ð1 (x) + ðŒ2 ð2 (x) + â
â
â
+ ðŒð ðð (x) for every ð¥ â ð â©ð For every x â ð=1 kernel ðð , ðð (x) = 0, ð = 1, 2, . . . , ð and therefore ð (x) = 0. Therefore ð¥ â kernel ð . That is ð â©
kernel ðð â kernel ð
ð=1
124
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.49 Let ð» be a hyperplane in ð. Then there exists a unique subspace ð such that ð» = x0 + ð for some x0 â ð» (Exercise 1.153). There are two cases to consider. Case 1: x0 â / ð . For every x â ð, there exists unique ðŒx â â such x = ðŒx x0 + ð£ for some ð£ â ð Deï¬ne ð (x) = ðŒx . Then ð : ð â â. It is straightforward to show that ð is linear. Since ð» = x0 + ð , ðŒx = 1 if and only if x â ð». Therefore ð» = { x â ð : ð (x) = 1 } Case 2: x0 â ð . In this case, choose some x1 â / ð . Again, for every x â ð, there exists a unique ðŒx â â such x = ðŒx x1 + ð£ for some ð£ â ð and ð (x) = ðŒx is a linear functional on ð. Furthermore x0 â ð implies ð» = ð (Exercise 1.153) and therefore ð (x) = 0 if and only if x â ð». Therefore ð» = { x â ð : ð (x) = 0 } Conversely, let ð be a nonzero linear functional in ð â² . Let ð = kernel ð and choose x0 â ð â1 (1). (This is why we require ð â= 0). For any x â ð ð (x â ð (x)x0 ) = ð (x) â ð (x) à 1 = 0 so that x â ð (x)x0 â ð . That is, x = ð (x)x0 + ð£ for some ð£ â ð . Therefore, ð = lin (x0 , ð ) so that ð is a maximal proper subspace. For any ð â â, let x1 â ð â1 (ð). Then, for every x â ð â1 (ð), ð (x â x1 ) = 0 and { x : ð (x) = ð} = {x : ð (x â x1 ) = 0 } = x1 + ð which is a hyperplane. 3.50 By the previous exercise, there exists a linear functional ð such that ð» = { ð¥ â ð : ð (ð¥) = ð } for some ð â â. Since 0 â / ð», ð â= 0. Without loss of generality, we can assume that ð = 1. (Otherwise, take the linear functional 1ð ð ). To show that ð is unique, assume that ð is another linear functional with ð» = { x : ð (ð¥) = 1} = {x : ð(ð¥) = 1 } Then ð» â { x : ð (ð¥) â ð(ð¥) = 0 } Since ð» is a maximal subset, ð is the smallest subspace containing ð». Therefore ð (ð¥) = ð(ð¥) for every ð¥ â ð.
125
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.51 By Exercise 3.49, there exists a linear functional ð such that ð» = { ð¥ â ð : ð (ð¥) = 0 } Since x0 â / ð», ð (x0 ) â= 0. Without loss of generality, we can normalize so that ð (x0 ) = 1. (If ð (x0 ) = ð â= 1, then the linear functional ð â² = 1/cð has ð â² (x0 ) = 1 and kernel ð â² = ð».) To show that ð is unique, suppose that ð is another linear functional with kernel ð = ð» and ð(x0 ) = 1. For any x â ð, there exists ðŒ â â such that x = ðŒx0 + v with ð£ â ð» (Exercise 1.153). Since ð (v) = ð(v) = 0 and ð (x0 ) = ð(x0 ) = 1 ð(x) = ð(ðŒx0 + v) = ðŒð(ð¥0 ) = ðŒð (x0 ) = ð (ðŒx0 + v) = ð (x) 3.52 Assume ð = ðð, ð â= 0. Then ð (ð¥) = 0 ââ ð(ð¥) = 0 Conversely, let ð» = ð â1 (0) = ð â1 (0). If ð» = ð, then ð = ð = 0. Otherwise, ð» is a hyperplane containing 0. Choose some x0 â / ð». Every x â ð has a unique representation x = ðŒx0 + v with v â ð» (Exercise 1.153) and ð (x) = ðŒð (x0 ) ð(x) = ðŒð(x0 ) Let ð = ð (x0 )/ð(x0 ) so that ð (x0 ) = ðð(x0 ). Substituting ð (x) = ðŒð (x0 ) = ðŒðð(x0 ) = ðð(x) 3.53 ð continuous implies that the set { ð¥ â ð : ð (ð¥) = ð } = ð â1 (ð) is closed for every ð â â (Exercise 2.67). Conversely, let ð = 0 and assume that ð» = { ð¥ â ð : ð (ð¥) = 0 } is closed. There exists x0 â= 0 such that ð = lin {ð¥0 , ð»} (Exercise 1.153). Let xð â x be a convergent sequence in ð. Then there exist ðŒð , ðŒ â â and vð , ð£ â ð» such that xð = ðŒð x0 + vð , x = ðŒx0 + v and â¥xð â x⥠= â¥ðŒð x0 + vð â ðŒx0 + v⥠= â¥ðŒð x0 â ðŒx0 + vð â v⥠†â£ðŒð â ðŒâ£ â¥x0 ⥠+ â¥vð â v⥠â0 which implies that ðŒð â ðŒ. By linearity ð (xð ) = ðŒð ð (x0 ) + ð (vð ) = ðŒð ð (x0 ) since vð â ð» and therefore ð (xð ) = ðŒð ð (x0 ) â ðŒð (x0 ) = ð (x) ð is continuous.
126
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.54 ð (x + xâ² , y) =
ð â ð â
ððð (ð¥ð + ð¥â²ð )ðŠð
ð=1 ð=1
=
ð â ð â
ððð ð¥ð ðŠð +
ð=1 ð=1
ð â ð â
ððð ð¥â²ð ðŠð
ð=1 ð=1
= ð (x, y) + ð (xâ² , y) Similarly, we can show that ð (x, y + yâ² ) = ð (x, y) + ð (x, yâ² ) and ð (ðŒx, y) = ðŒð (x, y) = ð (x, ðŒy) for every ðŒ â â 3.55 Let x1 , x2 , . . . , xð be a basis for ð and y1 , y2 , . . . , yð be a basis for ð . Let the numbers ððð represent the action of ð on these bases, that is ððð = ð (xð , yð )
ð = 1, 2, . . . , ð, ð = 1, 2, . . . , ð
and let ðŽ be the ð à ð matrix of numbers ððð . Choose any x â ð and y â ð and let their representations in terms of the bases be x=
ð â
ðŒð xð and y =
ð=1
ð â
ðœð yð
ð=1
respectively. By the bilinearity of ð â â ð (x, y) = ð ( ðŒð xð , ðŒð yð ) =
â
ð
ðŒð ð (xð ,
ð
=
â
ðŒð
ð
=
â
ð
â
â
ðŒð yð )
ð
ðŒð ð (xð , yð )
ð
ðŒð
â
ð
=
â
ðŒð ððð
ð
ðŒð ðŽy
ð â²
= x ðŽy 3.56 Every y â ð â² is a linear functional on ð. Hence y(x + xâ² ) = y(x) + y(xâ² ) y(ðŒx) = ðŒy(x) and therefore ð (x + xâ² , y) = y(x + xâ² ) = y(x) + y(xâ² ) = ð (x, y) + ð (xâ² , y) ð (ðŒx, y) = y(ðŒx) = ðŒy(x) = ðŒð (x, y) 127
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
In the dual space ð â² (y + yâ² )(x) â¡ y(x) + yâ² (x) (ðŒy)(x) â¡ ðŒy(x) and therefore ð (x, y + yâ² ) = (y + yâ² )(x) = y(x) + yâ² (x) = ð (x, y) + ð (x, yâ² ) ð (x, ðŒy) = (ðŒy)(x) = ðŒy(x) = ðŒð (x, y) 3.57 Assume ð1 , ð2 â ðµðð¿(ð à ð, ð). Deï¬ne the mapping ð1 + ð2 : ð à ð â ð by (ð1 + ð2 )(x, y) = ð1 (x, y) + ð2 (x, y) We have to conï¬rm that ð1 + ð2 is bilinear, that is (ð1 + ð2 )(x1 + x2 , y) = ð1 (x1 + x2 , y) + ð2 (x1 + x2 , y) = ð1 (x1 , y) + ð1 (x2 , y) + ð2 (x1 , y) + ð2 (x2 , y) = ð1 (x1 , y) + ð2 (x1 , y) + ð1 (x1 , y) + ð2 (x2 , y) = (ð1 + ð2 )(x1 , y) + (ð1 + ð2 )(x2 , y) Similarly, we can show that (ð1 + ð2 )(x, y1 + y2 ) = (ð1 + ð2 )(x, y1 ) + (ð1 + ð2 )(x, y2 ) and (ð1 + ð2 )(ðŒx, y) = ðŒ(ð1 + ð2 )(x, y) = (ð1 + ð2 )(x, ðŒy) For every ð â ðµðð¿(ð à ð, ð) deï¬ne the function ðŒð : ð à ð â ð by (ðŒð )(x, y) = ðŒð (x, y) ðŒð is also bilinear, since (ðŒð )(x1 + x2 , y) = ðŒð (x1 + x2 , y) = ðŒð (x1 , y) + ðŒð (x2 , y) = (ðŒð )(x1 , y) + (ðŒð (x2 , y) Similarly (ðŒð )(x, y1 + y2 ) = (ðŒð )(x, y1 ) + (ðŒð )(x, y2 ) (ðŒð )(ðœx, y) = ðœ(ðŒð )(x, y) = (ðŒð )(x, ðœy) Analogous to (Exercise 2.78), ð1 + ð2 and ðŒð are also continuous 3.58
1. ðµð¿(ð, ð) is a linear space and therefore so is ðµð¿(ð, ðµð¿(ð, ð)) (Exercise 3.33).
128
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. ðx is linear and therefore ð (x, y1 + y2 ) = ð(x)(y1 + y2 ) = ð(x)(y1 ) + ð(x)(y2 ) = ð (x, y1 ) + ð (x, y2 ) and ð (x, ðŒy) = ð(x)(ðŒy) = ðŒð(x)(y) = ðŒð (x, y) Similarly, ð is linear and therefore ð (x1 + x2 , y) = ðx1 +x2 (y) = ðx1 (y) + ðx2 (y) = ð (x1 , y) + ð (x2 , y) and ð (ðŒx, y) = ððŒx (y) = ðŒðx (y) = ðŒð (x, y) ð is bilinear 3. Let ð â ðµðð¿(ð à ð, ð). For every x â ð, the partial function ðx : ð â ð is linear. Therefore ðx â ðµð¿(ð, ð) and ð â ðµð¿(ð, ðµð¿(ð, ð)). 3.59 Bilinearity and symmetry imply ð (x â ðŒy, x â ðŒy) = ð (x, x â ðŒy) â ðŒð (y, x â ðŒy) = ð (x, x) â ðŒð (x, y) â ðŒð (y, x) + ðŒ2 ð (y, y) = ð (x, x) â 2ðŒð (x, y) + ðŒ2 ð (y, y) Nonnegativity implies ð (x â ðŒy, x â ðŒy) = ð (x, x) â 2ðŒð (x, y) + ðŒ2 ð (y, y) ⥠0
(3.38)
for every x, y â ð and ðŒ â â Case 1 ð (x, x) = ð (y, y) = 0 Then (3.38) becomes â2ðŒð (x, y) ⥠0 Setting ðŒ = ð (x, y) generates ( )2 â2 ð (x, y) ⥠0 which implies that ð (x, y) = 0 Case 2 Either ð (x, x) > 0 or ð (y, y) > 0. Without loss of generality, assume ð (y, y) > 0 and set ðŒ = ð (x, y)/ð (y, y) in (3.38). That is ( ) ( )2 ð (x, y) ð (x, y) ð (x, x) â 2 ð (x, y) + ð (y, y) ⥠0 ð (y, y) ð (y, y) or ð (x, x) â
ð (x, y)2 â¥0 ð (y, y)
which implies ( )2 ð (x, y) †ð (x, x)ð (y, y) for every x, y â ð 129
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
3.60 A Euclidean space is a ï¬nite-dimensional normed space, which is complete (Proposition 1.4). 3.61 ð (x, y) = xð y satisï¬es the requirements of Exercise 3.59 and therefore (xð y)2 †(xð x)(yð y) Taking square roots ð x y †â¥x⥠â¥y⥠3.62 By deï¬nition, the inner product is a bilinear functional. To show that it is continuous, let ð be an inner product space with inner product denote by xð y. Let xð â x and yð â y be sequences in ð. ð ð ð (x ) y â xð y = (xð )ð yð â (xð )ð y + (xð )ð y â xð y †(xð )ð yð â (xð )ð y + (xð )ð y â xð y †(xð )ð (yð â y) + (xð â x)ð y Applying the Cauchy-Schwartz inequality ð ð ð (x ) y â xð y †â¥xð ⥠â¥yð â y⥠+ â¥xð â x⥠â¥y⥠Since the sequence xð converges, it is bounded, that is there exists ð such that â¥xð ⥠†ð for every ð. Therefore ð ð ð (x ) y â xð y †â¥xð ⥠â¥yð â y⥠+ â¥xð â x⥠â¥y⥠†ð â¥yð â y⥠+ â¥xð â x⥠â¥y⥠â 0 3.63 Applying the properties of the inner product â â â¥x⥠= xð x ⥠0 â â â¥x⥠= xð x = 0 if and only if x = 0 â â â â¥ðŒx⥠= (ðŒx)ð (ðŒx) = ðŒ2 xð x = â£ðŒâ£ â¥x⥠To prove the triangle inequality, observe that bilinearity and symmetry imply 2
â¥x + y⥠= (x + y)ð (x + y) = xð x + xð y + yð x + zð z = xð x + 2xð y + yð y 2
2
= â¥x⥠+ 2xð y + â¥y⥠†â¥xâ¥2 + 2 xð y + â¥yâ¥2 Applying the Cauchy-Schwartz inequality 2
2
â¥x + y⥠†â¥x⥠+ 2 â¥x⥠â¥y⥠+ â¥yâ¥
2
= (â¥x⥠+ â¥yâ¥)2 3.64 For every y â ð, the partial function ðy (x) = xð y is a linear functional on ð (since xð y is bilinear). Continuity follows from the Cauchy-Schwartz inequality, since for every x â ð â£ðy (x)⣠= xð y †â¥y⥠â¥x⥠130
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics which shows that â¥ðy ⥠†â¥yâ¥. In fact, â¥ðy ⥠= â¥y⥠since â¥ðy ⥠= sup â£ðy (x)⣠â¥xâ¥=1
( ) y ⥠ðy â¥y⥠( y )ð = y â¥y⥠=
1 ð y y = â¥y⥠â¥yâ¥
3.65 By the Weierstrass Theorem (Theorem 2.2), the continuous function ð(x) = â¥x⥠attains a maximum on the compact set ð at some point x0 . We claim that x0 is an extreme point. Suppose not. Then, there exist x1 , x2 â ð such that x0 = ðŒx1 + (1 â ðŒ)x2 = x2 + ðŒ(x1 â x2 ) Since x0 maximizes â¥x⥠on ð
( )ð ( ) 2 2 x2 + ðŒ(x1 â x2 ) â¥x2 ⥠†â¥x0 ⥠= x2 + ðŒ(x1 â x2 ) 2
= â¥x2 ⥠+ 2ðŒxð2 (x1 â x2 ) + ðŒ2 â¥x1 â x2 â¥
2
or 2
2xð2 (x1 â x2 ) + ðŒ â¥x1 â x2 ⥠⥠0
(3.39)
Similarly, interchanging the role of x1 and x2 2
2xð1 (x2 â x1 ) + ðŒ â¥x2 â x1 ⥠⥠0 or â2xð1 (x1 â x2 ) + ðŒ â¥x1 â x2 â¥2 ⥠0
(3.40)
Adding the inequalities (3.39) and (3.40) yields 2
2(x2 â x1 )ð (x1 â x2 ) + 2ðŒ â¥x1 â x2 ⥠⥠0 or 2(x2 â x1 )ð (x2 â x1 ) = â2(x2 â x1 )ð (x1 â x2 ) †2ðŒ â¥x1 â x2 â¥2 and therefore â¥x2 â x1 ⥠†ðŒ â¥x2 â x1 ⥠Since 0 < ðŒ < 1, this implies that â¥x1 â x2 ⥠= 0 or x1 = x2 which contradicts our premise that x0 is not an extreme point. 3.66 Using bilinearity and symmetry of the inner product â¥x + yâ¥2 + â¥x â yâ¥2 = (x + y)ð (x + y) + (x â y)ð (x â y) = xð x + xð y + yð x + yð y + xð x â xð y â yð x + yð y = 2xð x + 2yð y 2
= 2 â¥x⥠+ 2 â¥y⥠131
2
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.67 Note that â¥ð¥â¥ = â¥ðŠâ¥ = 1 and â¥ð¥ + ðŠâ¥ = sup
( ) ð¥(ð¡) + ðŠ(ð¡) = sup (1 + ð¡) = 2
â¥ð¥ â ðŠâ¥ = sup
( ) ð¥(ð¡) â ðŠ(ð¡) = sup (1 â ð¡) = 1
0â€ð¡â€1 0â€ð¡â€1
0â€ð¡â€1 0â€ð¡â€1
so that 2
2
2
â¥ð¥ + ðŠâ¥ + â¥ð¥ â ðŠâ¥ = 5 â= 2 â¥ð¥â¥ + 2 â¥ð¥â¥
2
Since ð¥ and ðŠ do not satisfy the parallelogram law (Exercise 3.66), ð¶(ð) cannot be an inner product space. 3.68 Let {x1 , x2 , . . . , xð } be a set of pairwise orthogonal vectors. Assume 0 = ðŒx1 + ðŒ2 x2 + â
â
â
+ ðŒð xð Using bilinearity, this implies 0 = 0ð xð =
ð â
ðŒð xðð xð = ðŒð â¥xð â¥
ð=1
for every ð = 1, 2, . . . , ð. Since xð â= 0, this implies ðŒð = 0 for every ð = 1, 2, . . . , ð. We conclude that the set {x1 , x2 , . . . , xð } is linearly independent (Exercise 1.133). 3.69 Let x1 , x2 , . . . , xð be a orthonormal basis for ð. Since ðŽ represents ð ð (xð ) =
ð â
ððð xð
ð=1
for ð = 1, 2, . . . , ð. Taking the inner product with xð , ( ð ) ð â â ð ð xð ð (xð ) = xð ððð xð = ððð xðð xð ð=1
ð=1
Since {x1 , x2 , . . . , xð } is orthonormal { xðð xð
=
1 0
if ð = ð otherwise
so that the last sum simpliï¬es to xðð ð (xð ) = ððð for every ð, ð 3.70
1. By the Cauchy-Schwartz inequality ð x y †â¥x⥠â¥x⥠for every x and y, so that ð x y â€1 â£cos ð⣠= â¥x⥠â¥y⥠which implies â1 †cos ð †1 132
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
2. Since cos 90 = 0, ð = 90 implies that xð y = 0 or x ⥠y. Conversely, if x ⥠y, xð y = 0 and cos ð = 0 which implies ð = 90 degrees. 3.71 By bilinearity 2
2
â¥x + y⥠= (x + y)ð (x + y) = â¥x⥠+ xð y + yð x + â¥yâ¥
2
If x ⥠y, xð y = yð x = 0 and 2
2
â¥x + y⥠= â¥x⥠+ â¥y⥠3.72
2
Ë â ð and let ðË be the set of all x â ð which are closer to y 1. Choose some x Ë , that is than x ðË = { x â ð : â¥x â y⥠†â¥Ë x â y⥠} ðË is compact (Proposition 1.4). By the Weierstrass theorem (Theorem 2.2), the continuous function ð(x) = â¥xyâ¥ Ë That is attains a minimum on ðË at some point x0 â ð. â¥x0 â y⥠†â¥x â y⥠for every x â ðË A fortiori â¥x0 â y⥠†â¥x â y⥠for every x â ð
2. Suppose there exists some x1 â ð such that â¥x1 â y⥠= â¥x0 â y⥠= ð¿ By the parallelogram law (Exercise 3.66) 2
â¥x0 â x1 ⥠= â¥x0 â y + y â x1 ⥠2
2 2
= 2 â¥x0 â y⥠+ 2 â¥x1 â y⥠â â¥(x0 â y) â (y â x1 )⥠2 2 2 2 1 = 2 â¥x0 â y⥠+ 2 â¥x1 â y⥠â 2 (x0 + x1 ) â y 2 2 1 = 2ð¿ 2 + 2ð¿ 2 â 22 2 (x0 + x1 ) â y
2
since 12 (x0 + x1 ) â ð and therefore 12 (x0 + x1 ) â y ⥠ð¿ so that 2
â¥x0 â x1 ⥠†2ð¿ 2 + 2ð¿ 2 â 4ð¿ 2 = 0 which implies that x1 = x0 . 3. Let x â ð. Since ð is convex, the line segment ðŒx+(1âðŒ)x0 = x0 +ðŒ(xâx0 ) â ð and therefore (since x0 is the closest point) ( 2 ) â¥x0 â yâ¥2 †x0 + ðŒ(x â x0 ) â y 2
= â¥(x0 â y) + ðŒ(x â x0 )⥠( )ð ( ) = (x0 â y) + ðŒ(x â x0 ) (x0 â y) + ðŒ(x â x0 ) 2
= â¥x0 â y⥠+ 2ðŒ(x0 â y)ð (x â x0 ) + ðŒ2 â¥x â x0 ⥠133
2
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics which implies that 2
2ðŒ(x0 â y)ð (x â x0 ) + ðŒ2 â¥x â x0 ⥠⥠0 Dividing through by ðŒ 2
2(x0 â y)ð (x â x0 ) + ðŒ â¥x â x0 ⥠⥠0 which inequality must hold for every 0 < ðŒ < 1. Letting ðŒ â 0, we must have (x0 â y)ð (x â x0 ) ⥠0 as required. 3.73
1. Using the parallelogram law (Exercise 3.66), 2
â¥xð â xð ⥠= â¥(xð â y) + (y â xð )â¥
2
2
2
2
= 2 â¥xð â y⥠+ 2 â¥y â xð ⥠â 2 â¥xð + xð â¥
for every ð, ð. Since ð is convex, (xð +xð )/2 â ð and therefore â¥xð + xð ⥠⥠2ð. Therefore 2
2
2
â¥xð â xð ⥠= 2 â¥xð â y⥠+ 2 â¥y â xð ⥠â 4ð2 Since â¥xð â y⥠â ð and â¥xð â y⥠â ð as ð, ð â â, we conclude that 2 â¥xð â xð ⥠â 0. That is, (xð ) is a Cauchy sequence. 2. Since ð is a closed subspace of complete space, there exists x0 â ð such that xð â x0 . By continuity of the norm â¥x0 â y⥠= lim â¥xð â y⥠= ð ðââ
Therefore â¥x0 â y⥠†â¥x â y⥠for every x â ð Uniqueness follows in the same manner as the ï¬nite-dimensional case. 3.74 Deï¬ne ð : ð â ð by ð(y) = { x â ð : x is closest to y } The function ð is well-deï¬ned since for every y â ð there exists a unique point x â ð which is closest to y (Exercise 3.72). Clearly, for every x â ð, x is the closest point to x. Therefore ð(x) = x for every x â ð. To show that ð is continuous, choose any y1 and y2 in ð x1 = ð(y1 ) and x2 = ð(y2 )
134
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
be the corresponding closest points in ð. Then ( )ð ( ) 2 â¥(y1 â y2 ) â (x1 â x2 )⥠= (y1 â y2 ) â (x1 â x2 ) (y1 â y1 ) â (x1 â x2 ) = (y1 â y2 )ð (y1 â y2 ) + (x1 â x2 )ð (x1 â x2 ) â 2(y1 â y2 )ð (x1 â x2 ) = â¥y1 â y2 â¥2 + â¥x1 â x2 â¥2 â 2(y1 â y2 )ð (x1 â x2 ) 2
2
= â¥y1 â y2 ⥠+ â¥x1 â x2 ⥠â 2(y1 â y2 )ð (x1 â x2 ) 2
â 2 â¥x1 â x2 ⥠+ 2(x1 â x2 )ð (x1 â x2 ) 2
2
= â¥y1 â y2 ⥠â â¥x1 â x2 ⥠( )ð + 2 (x1 â x2 ) â (y1 â y2 ) (x1 â x2 ) = â¥y1 â y2 â¥2 â â¥x1 â x2 â¥2 + 2(x1 â y1 )ð (x1 â x2 ) â 2(x2 â y2 )ð (x1 â x2 ) 2
2
= â¥y1 â y2 ⥠â â¥x1 â x2 â¥
â 2(x1 â y1 )ð (x2 â x1 ) â 2(x2 â y2 )ð (x1 â x2 ) so that 2
2
â¥y1 â y2 ⥠â â¥x1 â x2 ⥠= â¥(y1 â y2 ) â (x1 â x2 )â¥
2
+ 2(x1 â y1 )ð (x2 â x1 ) + 2(x2 â y2 )ð (x1 â ð¥2 ) Using Exercise 3.72 (x1 â y1 )ð (x2 â x1 ) ⥠0 and (x2 â y2 )ð (x1 â x2 ) ⥠0 which implies that the left-hand side â¥y1 â y2 â¥2 â â¥x1 â x2 â¥2 ⥠0 or â¥x1 â x2 ⥠= â¥ð(y1 ) â ð(y2 )⥠†â¥y1 â y2 ⥠ð is Lipschitz continuous. 3.75 Let ð = kernel ð . Then ð is a closed subspace of ð. If ð = ð, then ð is the zero functional and y = 0 is the required element. Otherwise chose any y â / ð and let x0 be the closest point in ð (Exercise 3.72). Deï¬ne z = x0 â y. Then z â= 0 and zð x ⥠0 for every x â ð Since ð is subspace, this implies that zð x = 0 for every x â ð that is z is orthogonal to ð. Let ðË be the subset of ð deï¬ned by ðË = { ð (x)z â ð (z)x : x â ð } For every x â ðË
( ) ð (x) = ð ð (x)z â ð (z)x = ð (x)ð (z) â ð (z)ð (x) = 0 135
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Therefore ðË â ð. For every x â ð
( )ð ð (x)z â ð (z)x z = ð (x)zð z â ð (z)xð z = 0 since z â ð ⥠. Therefore ð (x) =
ð (z) â¥zâ¥
(
ð 2x z
= xð
zð (z) â¥zâ¥
2
) = xð y
where y=
zð (z) â¥zâ¥
2
3.76 ð â is always complete (Proposition 3.3). To show that it is a Hilbert space, we have to that it has an inner product. For this purpose, it will be clearer if we use an alternative notation < x, y > to denote the inner product of x and y. Assume ð is a Hilbert space. By the Riesz representation theorem (Exercise 3.75), for every ð â ð â there exists yð â ð such that ð (x) =< x, yð > for every x â ð Furthermore, if yð represents ð and yð represents ð â ð â , then yð + yð represents ð + ð and ðŒyð represents ðŒð since (ð + ð)(x) = ð (x) + ð(x) =< x, yð > + < x, yð >=< x, yð + yð > (ðŒð )(x) = ðŒð (x) = ðŒ < x, yð >=< x, ðŒyð > Deï¬ne an inner product on ð â by < ð, ð >=< yð , yð > We show that it satisï¬es the properties of an inner product, namely symmetry < ð, ð >=< yð , yð >=< yð , yð >=< ð, ð > additivity < ð1 + ð2 , ð >=< yð , yð1 +ð2 >=< yð , yð1 + yð2 >=< ð1 , ð > + < ð2 , ð > homogeneity < ðŒð, ð >=< yð , ðŒyð >= ðŒ < yð , yð >= ðŒ < ð, ð > positive deï¬niteness < ð, ð >=< yð , yð >⥠0 and < ð, ð >=< yð , yð >= 0 if and only if ð = ð. Therefore, ð â is a complete inner product space, that is a Hilbert space. 3.77 Let ð be a Hilbert space. Applying the previous exercise a second time, ð ââ is also a Hilbert space. Let ð¹ be an arbitrary functional in ð ââ . By the Riesz representation theorem, there exists ð â ð â such that ð¹ (ð ) =< ð, ð > for every ð â ð â Again by the Riesz representation theorem, there exists xð (representing ð ) and xð¹ (representing ð) in ð such that ð¹ (ð ) =< ð, ð >=< xð¹ , xð > and ð (x) =< x, xð > 136
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics In particular, ð (xð¹ ) =< xð¹ , xð >= ð¹ (ð )
That is, for every ð¹ â ð ââ , there exists an element xð¹ â ð such that ð¹ (ð ) = ð (xð¹ ) ð is reï¬exive. 3.78
1. Adapt Exercise 3.64.
2. By Exercise 3.75, there exists unique xâ â ð such that ðy (x) = xð xâ 3. Substituting ð (x)ð y = ðy (x) = xð xâ = ð¥ð ð â (y) 4. For every y1 , y2 â ð ( ( ) ) xð ð â (y1 + y2 ) = ð (x)ð y1 + y2 = ð (x)ð y1 + ð (x)ð y1 = xð ð â (y1 ) + xð ð â (y1 ) and for every y â ð xð ð â (ðŒy) = ð (x)ð ðŒy = ðŒð (x)ð y = ðŒxð ð â (y) = xð ðŒð â (y) 3.79 The zero element 0ð is a ï¬xed point of every linear operator (Exercise 3.13). 3.80 ðŽðŽâ1 = ðŒ so that det(ðŽ) det(ðŽâ1 ) = det(ðŒ) = 1 3.81 Expanding along the ðth row using (3.8) det(ð¶) =
ð â
(â1)ð+ð (ðŒððð + ðœððð ) det(ð¶ðð )
ð=1 ð â
=ðŒ
(â1)ð+ð ððð det(ð¶ðð ) + ðœ
ð=1
ð â
(â1)ð+ð ððð det(ð¶ðð )
ð=1
But the matrices diï¬er only in the ðth row and therefore ðŽðð = ðµðð = ð¶ðð ,
ð = 1, 2, . . . ð
so that det(ð¶) = ðŒ
ð â
(â1)ð+ð ððð det(ðŽðð ) + ðœ
ð=1
ð â
(â1)ð+ð ððð det(ðµðð )
ð=1
= ðŒ det(ðŽ) + ðœ det(ðµ) 3.82 Suppose that x1 and x2 are eigenvectors corresponding to the eigenvalue ð. By linearity ð (x1 + x2 ) = ð (x1 ) + ð (x2 ) = ðx1 + ðx2 = ð(x1 + x2 ) and ð (ðŒx1 ) = ðŒð (x1 ) = ðŒðx Therefore x1 + x2 and ðŒx1 are also eigenvectors. 137
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
3.83 Suppose ð is singular. Then there exists x â= 0 such that ð (x) = 0. Therefore x is an eigenvector with eigenvalue 0. Conversely, if 0 is an eigenvalue ð (x) = 0x = 0 for any x â= 0. Therefore ð is singular. 3.84 Since ð (x) = ðx ð (x)ð x = ðxð x = ðxð x 3.85 By Exercise 3.69 ððð = xðð ð (xð ) ððð = xðð ð (xð ) = ð (xð )ð xð and therefore ððð = ððð ââ xðð ð (xð ) = ð (xð )ð xð 3.86 By bilinearity xð1 ð (x2 ) = xð1 ð2 x2 = ð2 xð1 x2 ð (x1 )ð x2 = ð1 xð1 x2 = ð1 xð1 x2 Since ð is symmetric, this implies (ð1 â ð2 )xð1 x2 = 0 and ð1 â= ð2 implies xð1 x2 = 0. 3.87
1. Since ð compact and ð is continuous (Exercises 3.31, 3.62), the maximum is attained at some x0 â ð (Theorem 2.2), that is ð = ð (x0 )ð x0 ⥠ð (x)ð x for every x â ð Hence ( )ð ð(x, y) = ðx â ð (x) y is well-deï¬ned.
2. For any x â ð ( )ð ð(x, x) = ðx â ð (x) x = ðxð x â ð (x)ð x = ð â¥xâ¥2 â ð (x)ð x ( 2
2
= ð â¥x⥠â â¥x⥠ð
x
)ð ( 2
â¥x⥠) 2( ð = â¥x⥠ð â ð (z) z ⥠0
since z = x/ â¥x⥠â ð. 138
x â¥xâ¥
) 2
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3. Since ð is symmetric ( )ð ð(y, x) = ðy â ð (y) x = ðyð x â ð (y)ð x = ðxð y â ð (x)ð ðŠ ( )ð = ðx â ð (x) y = ð(x, y) 4. ð satisï¬es the conditions of Exercise 3.59 and therefore (ð(x, y))2 †ð(x, x)ð(y, y) for every x, y â ð
(3.41)
By deï¬nition ð(x0 , x0 ) = 0 and (3.41) implies that ð(x0 , y) = 0 for every y â ð That is ( )ð ð(x0 , y) = ðx0 â ð (x0 ) y = 0 for every ðŠ â ð and therefore ðx0 â ð (x0 ) = 0 or ð (x0 ) = ðx0 In other words, x0 is an eigenvector. By construction, â¥x0 ⥠= 1. 3.88
1. Suppose x2 , x3 â ð. Then ( )ð ðŒx2 + ðœx3 x1 = ðŒxð2 x1 + ðœxð3 x1 = 0 so that ðŒx2 + ðœx3 â ð. ð is a subspace. Let {x1 , x2 , . . . , xð } be a basis for ð (Exercise 1.142). For x â ð, there exists (Exercise 1.137) unique ðŒ1 , ðŒ2 , . . . , ðŒð such that x = ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð If x â ð xð x1 = ðŒ1 xð1 x1 = 0 which implies that ðŒ1 = 0. dim ð = ð â 1.
Therefore, x2 , x3 , . . . , xð span ð and therefore
2. For every x â ð, ð (x)ð x0 = xð ð (x0 ) = xð ðx0 = ðxð x0 = 0 since ð is symmetric. Therefore ð (x) â {x0 }⥠= ð. 3.89 Let ð be a symmetric operator. By the spectral theorem (Proposition 3.6), there exists a diagonal matrix ðŽ which represents ð . The elements of ðŽ are the eigenvalues of ð . By Proposition 3.5, the determinant of ðŽ is the product of these diagonal elements. 139
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.90 By linearity ð (x) =
â
ð¥ð ð (xð )
ð
ð deï¬nes a quadratic form since â ( )ð â â ââ ââ â â ð(x) = xð ð (x) = ð¥ð xð ð¥ð ð (xð )â = ð¥ð ð¥ð xðð ð (xð ) = ððð ð¥ð ð¥ð ð
ð
ð
ð
ð
ð
by Exercise 3.69. 3.91 Let ð be the symmetric linear operator deï¬ning ð ð(x) = xð ð (x) By the spectral theorem (Proposition 3.6), there exists an orthonormal basis x1 , x2 , . . . , xð comprising the eigenvectors of ð . Let ð1 , ð2 , . . . , ðð be the corresponding eigenvalues, that is ð (xð ) = ðð xð
ð = 1, 2 . . . , ð
Then for x = ð¥1 x1 + ð¥2 x2 + â
â
â
+ ð¥ð xð ð(x) = xð ð (x) = (ð¥1 x1 + ð¥2 x2 + â
â
â
+ ð¥ð xð )ð ð (ð¥1 x1 + ð¥2 x2 + â
â
â
+ ð¥ð xð ) = (ð¥1 x1 + ð¥2 x2 + â
â
â
+ ð¥ð xð )ð (ð¥1 ð (x1 ) + ð¥2 ð (x2 ) + â
â
â
+ ð¥ð ð (xð )) = (ð¥1 x1 + ð¥2 x2 + â
â
â
+ ð¥ð xð )ð (ð¥1 ð1 x1 + ð¥2 ð2 x2 + â
â
â
+ ð¥ð ðð xð ) = ð¥1 ð1 ð¥1 + ð¥2 ð2 ð¥2 + â
â
â
+ ð¥ð ðð ð¥ð = ð1 ð¥21 + ð2 ð¥22 + â
â
â
+ ðð ð¥2ð 3.92
1. Assuming that ð11 â= 0, the quadratic form can be rewritten as follows ð(ð¥1 , ð¥2 ) = ð11 ð¥21 + 2ð12 ð¥1 ð¥2 + ð22 ð¥22 = ð11 ð¥21 + 2ð12 ð¥1 ð¥2 +
ð212 2 ð212 2 ð¥ â ð¥ + ð22 ð¥22 ð11 2 ð11 2
(
( )2 ) ( ) ð ð ð212 12 12 2 = ð11 ð¥1 + 2 ð¥1 ð¥2 + ð¥2 + ð22 â ð¥22 ð11 ð11 ð11 ( )2 ( ) ð11 ð22 â ð212 ð12 = ð11 ð¥1 + ð¥2 + ð¥22 ð11 ð11 2. We observe that ð must be positive for every ð¥1 and ð¥2 provided ð11 > 0 and ð11 ð22 â ð212 > 0. Similarly ð must be negative for every ð¥1 and ð¥2 if ð11 > 0 and ð11 ð22 â ð212 > 0. Otherwise, we can choose values for ð¥1 and ð¥2 which make ð both positive and negative. Note that the condition ð11 ð22 > ð212 > 0 implies that ð11 and ð12 must have the same sign. 3. If ð11 = ð22 = 0, then ð is indeï¬nite. Otherwise, if ð11 = 0 but ð22 â= 0, then the ð can we can âcomplete the squareâ using ð22 and deduce { } { } nonnegative ð11 , ð22 ⥠0 ð is deï¬nite if and only if and ð11 ð22 ⥠ð212 nonpositive ð11 , ð22 †0 140
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.93 Let ð : ð â â be a quadratic form on ð. Then there exists a linear operator ð such that ð(x) = xð ð (x) and (Exercise 3.13) ð(0) = 0ð ð (0) = 0 3.94 Suppose to the contrary that the positive (negative) deï¬nite matrix ðŽ is singular. Then there exists x â= 0 such that ðŽx = 0 and therefore xâ² ðŽx = 0 contradicting the deï¬niteness of ðŽ. 3.95 Let e1 , e2 , . . . , eð be the standard basis for âð (Example 1.79). Then for every ð eâ²ð ðŽeð = ððð > 0 3.96 Let ð be the quadratic form deï¬ned by ðŽ. By Exercise 3.91, there exists an orthonormal basis such that ð(x) = ð1 x21 + ð2 x22 + â
â
â
+ ðð x2ð where ð1 , ð2 , . . . , ðð are the eigenvalues of ðŽ. This implies ⧠⧠⫠⫠ðð > 0 ð(x) > 0       ⚠⚠⬠⬠ð(x) ⥠0 ðð ⥠0 ââ ð = 1, 2, . . . , ð ðð < 0 ð(x) < 0     â© â â© â ð(x) †0 ðð †0 3.97 Let ð1 , ð2 , . . . , ðð be the eigenvalues of ðŽ. By Exercise 3.89 det(ðŽ) = ð1 ð2 . . . ðð By Exercise 3.96, ðð ⥠0 for every ð and therefore det(ðŽ) ⥠0. We conclude that det(ðŽ) > 0 ââ ðð > 0 for every ð ââ ðŽ is positive deï¬nite by Exercise 3.96. 3.98
1. ðŽ0 = 0. Therefore, 0 is always a solution.
2. Assume x1 and x2 are solutions, that is ðŽx1 = 0 and ðŽx2 = 0 Then ðŽ(x1 + x2 ) = ðŽx1 + ðŽx2 = 0 x1 + x2 is also a solution. 3. Let ð be the linear function deï¬ned by ð (x) = ðŽx The system of equations ðŽx = 0 has a nontrivial solution if and only if kernel ð â= {0} ââ nullity ð > 0 By the rank theorem (Exercise 3.24) rankð + nullityð = dim ð so that nullity ð > 0 ââ rankð < dim ð = ð 141
Solutions for Foundations of Mathematical Economics 3.99
c 2001 Michael Carter â All rights reserved
1. Assume x1 and x2 are solutions of (3.16). That is ðŽx1 = c and ðŽx2 = c Subtracting ðŽx1 â ðŽx2 = ðŽ(x1 â x2 ) = 0
2. Assume xð solves (3.16) while x is any solution to (3.17). That is ðŽxð = c and ðŽx = 0 Adding ðŽxð + ðŽx = ðŽ(xð + x) = c We conclude that xð + x solves (3.16) for every x â ðŸ. 3. If 0 is the only solution of (3.17), ðŸ = {0}. Assume x1 and x2 are solutions of (3.16). Then x1 â x2 â ðŸ = {0} which implies x1 = x2 . 3.100 Let ð = { x : ðŽx = ð }. For every x, y â ð and ðŒ â â ðŽðŒx + (1 â ðŒ)y = ðŒðŽx + (1 â ðŒ)ðŽy = ðŒc + (1 â ðŒ)c = ð Therefore, z = ðŒx + (1 â ðŒ)y â ð. ð is aï¬ne. 3.101 Let ð â= â
be an aï¬ne set âð . Then there exists a unique subspace ð such that ð = x0 + ð for some x0 â ð (Exercise 1.150). The orthogonal complement of ð is ð ⥠= { a â ð : ax = 0 for every x â ð } Let (a1 , a2 , . . . , að ) be a basis for ð ⥠. Then ð = (ð ⥠)⥠= {x : að x = 0,
ð = 1, 2, . . . ð}
Let ðŽ be the ðÃð matrix whose rows are a1 , a2 , . . . , að . Then ð is the set of solutions to the homogeneous linear system ðŽx = 0, that is ð = { ð¥ : ðŽx = 0 } Therefore ð = x0 + ð = x0 + { x : ðŽx = 0 } = { x : ðŽ(x â x0 ) = 0 } = { x : ðŽx = c } where c = ðŽx0 . 3.102 Consider corresponding homogeneous system ð¥1 + 3ð¥2 = 0 ð¥1 â ð¥2 = 0 142
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Multiplying the second equation by 3 ð¥1 + 6ð¥2 = 0 3ð¥1 â 3ð¥2 = 0 and adding yields 4ð¥1 = 0
for which the only solution is ð¥1 = 0. Substituting in the ï¬rst equation implies ð¥2 = 0. The kernel of ð = ðŽx is {0}. Therefore dim ð (â2 ) = 2, and the system ðŽx = ð has a unique solution for every ð1 , ð2 . 3.103 We can write the system ðŽx = c in the form â â â â â â â â ð11 ð1ð ð1ð ð1 â .. â â .. â â .. â â .. â ð¥1 â . â + â
â
â
+ ð¥ð â . â + â
â
â
+ ð¥ð â . â = â . â ðð1
ððð
ððð
ðð
Subtracting c from the ðth column gives â â â â â â ð11 ð1ð ð¥ð ð1ð â ð1 â â â â . â â .. ð¥1 â ... â + â
â
â
+ â â + â
â
â
+ ð¥ð â .. â = 0 . ð¥ð ððð â ðð
ðð1
so that the columns of the matrix â ð11 . . . â .. ð¶=â . ...
ðð1
ððð
(ð¥ð ð1ð â ð1 ) .. .
(ð¥ð ððð â ðð )
... ...
â ð1ð .. â . â
ððð
are linearly dependent (Exercise 1.133). Therefore det(ð¶) = 0. Let ðµð denote the matrix obtained from ðŽ by replacing the ðth column with c. Then ðŽ, ðµð and ð¶ diï¬er only in the ðth column, with the ðth column of ð¶ being a linear combination of the ðth columns of ðŽ and ðµð . â â â â â â ð1ð ð1ð ð1ð â .. â â .. â â .. â = ð¥ â â . â â . â ðâ . â ððð
ððð
ððð
By Exercise 3.81 det(ð¶) = xð det(ðŽ) â det(ðµð ) = 0 and therefore ð¥ð =
det(ðµð ) det(ðŽ)
as required. 3.104 Let ( ð ð
)â1 ( ð ðŽ = ð ð¶ 143
ðµ ð·
)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics The inverse satisï¬es the equation ( )( ð ð ðŽ ð ð ð¶ In particular, this means that ðŽ and ( ð ð By Cramerâs rule (Exercise 3.103)
where Î = ðð â ðð. Similarly
ðµ ð·
)
( 1 = 0
) 0 1
ð¶ satisfy the equation )( ) ( ) ð ðŽ 1 = ð ð¶ 0
1 ð 0 ð ð = ðŽ = Î ð ð ð ð
ð 1 ð 0 âð = ð¶ = Î ð ð ð ð
ðµ and ð· are determined analogously. 3.105 A portfolio is duplicable if and only if there is a diï¬erent portfolio y â= x such that ð
x = ð
y or ð
(x â y) = 0 There exists a duplicable portfolio if and only if this homogeneous system has a nontrivial solution, that is if rank ð
< ðŽ. 3.106 State ð ¯ is insurable if there is a solution to the linear system ð
x = eð ¯
(3.42)
where eð ¯ is the ð ¯-th unit vector (the ð ¯ Arrow-Debreu security). (3.42) has a solution for every state ð if and only if ð (âðŽ ) = âð , that is rank ð
= ð. 3.107 Figure 3.1. 3.108 Let ð be an aï¬ne subset of âð . Then there exists (Exercise 3.101) a system of linear equations ðŽx = c such that ð = { x : ðŽx = c } Let að denote the ð-th row of ðŽ. Then ð = { ð¥ : að x = ðð , ð = 1, 2, . . . , ð } =
ð â©
{ x : að x = ðð }
ð=1
where each { x : að x = ðð } is a hyperplane in âð (Example 3.21). 144
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Figure 3.1: The solutions of three equations in two unknowns 3.109 Let ð = { x : ðŽx †c }. For every x, y â ð and 0 †ðŒ †1 ðŽx †c ðŽy †c and therefore ðŽðŒx + (1 â ðŒ)y = ðŒðŽx + (1 â ðŒ)ðŽy †ðŒc + (1 â ðŒ)c = ð Therefore, z = ðŒx + (1 â ðŒ)y â ð. ð is a convex set. 3.110 We have already seen that ð = { x : ðŽx †0 } is convex. To show that it is a cone, let x â ð. Then ðŽx †0 ðŽðŒx †0 so that ðŒx â ð. ð is a convex cone. 3.111
1. Each column ðŽð is a vector in âð . If the set {ðŽ1 , ðŽ2 , . . . , ðŽð } is linearly independent, it has at most ð elements, that is ð †ð and x is a basic feasible solution.
2. (a) Assume {ðŽ1 , ðŽ2 , . . . , ðŽð } are linearly dependent. Then (Exercise 1.133) there exist numbers ðŠ1 , ðŠ2 , . . . , ðŠð , not all zero, such that ðŠ1 ðŽ1 + ðŠ2 ðŽ2 + â
â
â
+ ðŠð ðŽð = 0 y = (ðŠ1 , ðŠ2 , . . . , ðŠð ) is a nontrivial solution to the homogeneous system. (b) For every ð¡ â â, âð¡y â kernel ð = ðŽx and xâ² = x â ð¡y is a solution of the corresponding nonhomogeneous system ðŽx = c. To see this directly, subtract ðŽð¡y = 0 145
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
from ðŽx = c to give ðŽxâ² = ðŽ(x â ð¡y) = c (c) Note that x > 0 and therefore ð¡Ë > 0 which implies that ð¥ Ëð > 0 for every ðŠð †0. For every ðŠð > 0, ð¥ð /ðŠð ⥠ð¡Ë, which implies that ð¥ð ⥠ð¡ËðŠð , so that ð¥Ëð ⥠ð¥ð â ð¡ËðŠð ⥠0 Ë is a feasible solution. Therefore, x (d) There exists some coordinate â such that ð¡Ë = ð¥â /ðŠâ so that ð¥ Ëâ = ð¥â â ð¡ËðŠâ = 0 so that ð â
c=
ð¥Ëð ðŽðœ
ð =1
ðâ=â
ð¥ Ë is a feasible solution with one less positive component. 3. Starting with any nonbasic feasible solution, this elimination technique can be repeated until the remaining vectors are linearly independent and a basic feasible solution is obtained. 3.112
1. Exercise 1.173.
2. For each ð, there exists ðð elements xðð and coeï¬cients ððð > 0 such that xð =
ðð â
ððð xðð
ð=1
and
â ðð
ð=1
ððð = 1. Hence x=
ð â
xð =
ð=1
ð â ðð â
ððð xðð
ð=1 ð=1
3. Direct computation. 4. Regarding the ððð as âvariablesâ and the points ð§ðð as coeï¬cents, z=
ðð ð â â
ððð zðð
ð=1 ð=1
is a linear equation system in which variables are restricted to be nonnegative. By the fundamental theorem of linear programming (Exercise 3.111), there exists
146
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
a basic feasible solution. That is, there exists coeï¬cients ððð ⥠0 and ððð > 0 for at most (ð + ð) components such that z=
ðð ð â â
ððð zðð
(3.43)
ð=1 ð=1
Decomposing, (3.43) implies x=
ðð ð â â
ððð xðð
ð=1 ð=1
and ðð â
ððð = 1
for every ð
ð=1
5. (3.43) implies that at least one ððð > 0 for every ð. This accounts for at least ð of the positive ððð . Since there are at most (ð + ð) coeï¬cients ððð which are strictly positive, there are at most ð indices ð which have more than one positive coeï¬cient ððð . For the remaining ð â ð indices, xð = xðð for some ð; that is xð â ðð . 3.113
1. Since ðŽ is productive, there exists x ⥠0 such that ðŽx > 0. Consider any z for which ðŽz ⥠0. For every ðŒ > 0 ðŽ(x + ðŒz) = ðŽx + ðŒðŽz > 0
(3.44)
Suppose to the contrary that z â⥠0. That is, there exists some component ð§ð < 0. Let ð§ð ðŒ = max{â } ð¥ð Without loss of generality, ð§1 attains this maximum, that is assume ðŒ = ð§1 /ð¥1 . Then ð¥1 + ðŒð§1 = 0 and ð¥ð + ðŒð§ð ⥠0 for every ð. Now consider the matrix ðµ = ðŒ â ðŽ. By the assumptions of the Leontief model (Example 3.35), the matrix ðŽ has 1 along the diagonal and negative oï¬-diagonal elements. That is ððð = 1
ð = 1, 2, . . . , ð
ððð †0
ð, ð = 1, 2, . . . , ð,
Therefore ðµ =ðŒ âðŽâ¥0 147
ð â= ð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
That is, every element of ðµ is nonnegative. Consequently since x + ðŒz ⥠0 ðµ(x + ðŒz) ⥠0
(3.45)
On the other hand, substituting ðŽ = ðŒ â ðµ in (3.45) (ðŒ â ðµ)(x + ðŒz) > 0 x + ðŒz > ðµ(x + ðŒz) which implies that the ï¬rst component of ðµ(x + ðŒz) is negative, contradicting (3.45). This contradiction establishes that z ⥠0. Suppose ðŽx = 0. A fortiori ðŽx ⥠0. By the previous part this implies x ⥠0. On the other hand, it also implies that âðŽx = ðŽ(âx) = 0 so that âx ⥠0. We conclude that x = 0 is the only solution to ðŽx = 0. ðŽ is nonsingular. Since ðŽ is nonsingular, the system ðŽx = y has a unique solution x for any y ⥠0. By the ï¬rst part, x ⥠0. 3.114 Suppose ðŽ is productive. By the previous exercise, ðŽ is nonsingular with inverse ðŽâ1 . Let eð be the ðth unit vector. Since eð ⥠0, there exists xð ⥠0 such that ðŽxð = eð Multiplying by ðŽâ1 xð = ðŽâ1 ðŽxð = ðŽâ1 eð = ðŽâ1 ð where ðŽâ1 is the ð column of ðŽâ1 . Since xð ⥠0 for every ð, we conclude that ðŽâ1 ⥠0. ð Conversely, assume that ðŽâ1 ⥠0. Let 1 = (1, 1, . . . , 1) denote a net output of 1 for each commodity. Then x = ðŽâ1 1 ⥠0 and ðŽx = 1 > 0 ðŽ is productive. 3.115 Takayama 1985, p.383, Theorem 4.C.4. 3.116 Let a0 = (ð01 , ð02 , . . . , ð0ð ) be the vector of labour requirements and ð€ the wage rate. The unit proï¬t of industry ð is â ðð = ðð + ððð ðð â ð€ð0 ðâ=ð
Recall that ððð †0 for ð â= ð. The vector of unit proï¬ts for all industries is Î = ðŽp â ð€ð0 Proï¬ts will be zero in all industries if there exists a price system p such that Î = ðŽp â ð€ð0 = 0 or ðŽp = ð€ð0
(3.46)
By the previous results, (3.46) has a unique nonnegative solution p = ðŽâ1 ð€ð0 if the technology ðŽ is productive. Furthermore, ðŽâ1 is nonnegative. Since ð0 > 0, so is p > 0. 148
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.117 Let ð¢ðµ denote the steady state unemployment rate for blacks. Then ð¢ðµ satisï¬es the equation ð¢ðµ = 0.0038(1 â ð¢ðµ ) + 0.8975ð¢ðµ which implies that ð¢ðµ = 0.036. That is, the data implies an unemployment rate of 3.6 percent for blacks. Similarly, the unemployment rate for white males ð¢ð satisï¬es the equation ð¢ð = 0.0022(1 â ð¢ð ) + 0.8614ð¢ð which implies that ð¢ð = 0.016 or 1.6 percent. 3.118 The transition matrix is
( ð =
) .6 .25 .4 .75
If the current state vector is x0 = (.4, .6), the state vector after a single mailing will be x1 = ð x0 ( )( ) .6 .25 .4 = .4 .75 .6 ( ) 0.39 = .61 Following a single mailing, the number of subscribers will drop to 30 percent of the mailing list, comprising 24 percent from renewals and 15 percent new subscriptions. 3.119 Let ð (ð¥) = ð¥2 . For every ð¥1 , ð¥2 â â and 0 †ðŒ †1 ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) = (ðŒð¥1 + (1 â ðŒ)ð¥2 )2 = (ðŒð¥1 + (1 â ðŒ)ð¥2 )(ðŒð¥1 + (1 â ðŒ)ð¥2 ) = ðŒ2 ð¥21 + 2ðŒ(1 â ðŒ)ð¥1 ð¥2 + (1 â ðŒ)2 ð¥22 = ðŒð¥21 + (1 â ðŒ)ð¥22 â ðŒð¥21 â (1 â ðŒ)ð¥22 + ðŒ2 ð¥21 + 2ðŒ(1 â ðŒ)ð¥1 ð¥2 + (1 â ðŒ)2 ð¥22 ) ( = ðŒð¥21 + (1 â ðŒ)ð¥22 â ðŒ(1 â ðŒ)ð¥21 â 2ðŒ(1 â ðŒ)ð¥1 ð¥2 + ðŒ(1 â ðŒ)ð¥22 = ðŒð¥21 + (1 â ðŒ)ð¥22 â ðŒ(1 â ðŒ)(ð¥1 â ð¥2 )2 †ðŒð¥21 + (1 â ðŒ)ð¥22 = ðŒð (ð¥1 ) + (1 â ðŒ)(ð¥2 ) 3.120 ð (ð¥) = ð¥ is linear and therefore convex. In the previous exercise we showed that ð¥2 is convex. Therefore ð (ð¥) = ð¥ð is convex for ð = 1, 2. Assume that ð is convex for
149
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ð â 1. Then ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) = (ðŒð¥1 + (1 â ðŒ)ð¥2 )ð = (ðŒð¥1 + (1 â ðŒ)ð¥2 )(ðŒð¥1 + (1 â ðŒ)ð¥2 )ðâ1 †(ðŒð¥1 + (1 â ðŒ)ð¥2 )(ðŒð¥ðâ1 + (1 â ðŒ)ð¥ðâ1 ) 1 2
(since ð¥ðâ1 is convex)
= ðŒ2 ð¥ð1 + ðŒ(1 â ðŒ)ð¥ðâ1 ð¥2 + ðŒ(1 â ðŒ)ð¥1 ð¥ðâ1 + (1 â ðŒ)2 ð¥ð2 1 2
= ðŒð¥ð1 + (1 â ðŒ)ð¥ð2 â ðŒð¥ð1 â (1 â ðŒ)ð¥ð2
+ ðŒ2 ð¥ð1 + ðŒ(1 â ðŒ)ð¥ðâ1 ð¥2 + ðŒ(1 â ðŒ)ð¥1 ð¥ðâ1 + (1 â ðŒ)2 ð¥ð2 1 2 ) ( ðâ1 ð = ðŒð¥ð1 + (1 â ðŒ)ð¥ð2 â ðŒ(1 â ðŒ) ð¥ð1 â ð¥1 ð¥ðâ1 â ð¥ ð¥ + ð¥ 2 2 2 1 ( ) ðâ1 ðâ1 ð ð = ðŒð¥1 + (1 â ðŒ)ð¥2 â ðŒ(1 â ðŒ) ð¥1 (ð¥1 â ð¥2 ) â ð¥2 (ð¥1 â ð¥2 ) ( ) ðâ1 â ð¥ ) = ðŒð¥ð1 + (1 â ðŒ)ð¥ð2 â ðŒ(1 â ðŒ) (ð¥1 â ð¥2 )(ð¥ðâ1 1 2 Since ð¥ð is monotonic (Example 2.53) ð¥ðâ1 â ð¥ðâ1 ⥠0 ââ ð¥1 â ð¥2 ⥠0 1 2 and therefore (ð¥1 â ð¥2 )(ð¥ðâ1 â ð¥ðâ1 )â¥0 1 2 We conclude that ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) †ðŒð¥ð1 + (1 â ðŒ)ð¥ð2 = ðŒð (ð¥1 ) + (1 â ðŒ)(ð¥2 ) ð is convex for all ð = 1, 2, . . . . 3.121 For given x1 , x2 â ð, deï¬ne ð : [0, 1] â ð by ð(ð¡) = (1 â ð¡)x1 + ð¡x2 Then ð(0) = x1 , ð(1) = x2 and â = ð â ð . Assume ð is convex. For any ð¡1 , ð¡2 â [0, 1], let ¯ 1 and ð(ð¡2 ) = x ¯2 ð(ð¡1 ) = x For any ðŒ â [0, 1]
) ( x1 + (1 â ðŒ)¯ ð ðŒð¡1 + (1 â ðŒ)ð¡2 = ðŒÂ¯ x2 ) ( ) ( x1 + (1 â ðŒ)¯ â ðŒð¡1 + (1 â ðŒ)ð¡2 = ð ðŒÂ¯ x2 †ðŒð (¯ x1 ) + (1 â ðŒ)ð (¯ x2 ) †ðŒâ(ð¡1 ) + (1 â ðŒ)ð¡2 )
â is convex. Conversely, assume â is convex for any x1 , x2 â ð. For any ðŒ â [0, 1] ð(ðŒ) = ðŒx1 + (1 â ðŒ)x2 and
) ( ð ðŒx1 + (1 â ðŒ)x2 = â(ðŒ) †ðŒâ(0) + (1 â ðŒ)â(1) = ðŒð (x1 ) + (1 â ðŒ)ð (x2 )
Since this is true for any x1 , x2 â ð, we conclude that ð is convex. 150
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
3.122 Assume ð is convex which implies epi ð is convex. The points (xð , ð (xð )) â epi ð . Since epi ð is convex ðŒ1 (x1 , ð (x1 )) + ðŒ2 (x1 , ð (x1 )) + â
â
â
+ (xð , ð (xð )) â epi ð that is ð (ðŒ1 x1 + ðŒ2 x2 + â
â
â
+ ðŒð xð ) †ðŒ1 ð (x1 ) + ðŒ2 ð (x1 ) + â
â
â
+ ðŒð ð (xð )) Conversely, letting ð = 2 and ðŒ = ðŒ1 , (3.25) implies that ð (ðŒx1 + (1 â ðŒ)x2 ) †ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) Jensenâs inequality can also be proved by induction from the deï¬nition of a convex function (see for example Sydsaeter + Hammond 1995; p.624). 3.123 For each ð, let ðŠð = log ð¥ð so that ð¥ð = ððŠð ðŒð ðŠð ð ð¥ðŒ ð = ð
Since ðð¥ is convex (Example 3.41) ð¥ð1 1 ð¥ð2 2 . . . ð¥ððð
ðð > 0 =
â
exp(ðŒð ðŠð ) = exp
(â
) â â ðŒð ððŠð = ðŒð ð¥ð ðŒð ðŠð â€
by Jensenâs inequality. Setting ðŒð = 1/ð, we have (ð¥1 ð¥2 . . . ð¥ð )1/ð â€
ð
1â ð¥ð ð ð=1
as required. 3.124 Assume ð is concave. That is for every x1 , x2 â ð and 0 †ðŒ †1 ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) Multiplying through by â1 reverses the inequality so that âð (ðŒx1 + (1 â ðŒ)x2 ) †âðŒð (x1 ) + (1 â ðŒ)ð (x2 ) = ðŒ â ð (x1 ) + (1 â ðŒ) â ð (x2 ) which shows that âð is concave. The converse follows analogously. 3.125 Assume that ð is concave. Then âð is convex and by Theorem 3.7 epi â ð = { (ð¥, ðŠ) â ð à â : ðŠ ⥠âð (ð¥), ð¥ â ð } is convex. But epi â ð = { (ð¥, ðŠ) â ð à â : ðŠ ⥠âð (ð¥), ð¥ â ð } = { (ð¥, ðŠ) â ð à â : ðŠ †ð (ð¥), ð¥ â ð } = hypo ð Therefore hypo ð is convex. Conversely, if hypo ð is convex, epi â ð is convex which implies that âð is convex and hence ð is concave.
151
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.126 Suppose that x1 minimizes the cost of producing ðŠ at input prices w1 while x2 ¯ be the weighted average price, that is minimizes cost at w2 . For some ðŒ â [0, 1], let w ¯ = ðŒw1 + (1 â ðŒ)w2 w ¯ minimizes cost at w. ¯ Then and suppose that x ¯ ðŠ) = w¯ ¯x ð(w, x = (ðŒw1 + (1 â ðŒ)w2 )¯ ¯ + (1 â ðŒ)w2 x ¯ = ðŒw1 x But since x1 and x2 minimize cost at w1 and w2 respectively ¯ ⥠ðŒw1 x1 = ðŒð(w1 , ðŠ) ðŒw1 x ¯ ⥠(1 â ðŒ)w2 x2 = (1 â ðŒ)ð(w2 , ðŠ) (1 â ðŒ)w2 x so that ¯ + (1 â ðŒ)w2 x ¯ ⥠ðŒð(w1 , ðŠ) + (1 â ðŒ)ðw2 , ðŠ) ¯ ðŠ) = ð(ðŒw1 + (1 â ðŒ)w2 , ðŠ) = ðŒw1 x ð(w, This establishes that the cost function ð is concave in w. 3.127 Since ð¢ is concave, Jensenâs inequality implies ( ð ) ð ð â1 â 1 1â ðð¡ ⥠ð¢(ðð¡ ) = ð¢ ð¢(ðð¡ ) ð ð ð ð¡=1 ð¡=1 ð¡=1 for any consumption stream ð1 , ð2 , . . . , ðð so that ( ð ) ð â â1 ðð¡ = ð ð¢(¯ ð= ð¢(ðð¡ ) †ð ð¢ ð) ð ð¡=1 ð¡=1 It is impossible to do better than consume a constant fraction ð¯ = ð€/ð of wealth in each period. 3.128 If ð¥1 = ð¥3 , the inequality is trivially satisï¬ed. Now assume ð¥1 â= ð¥3 . Since ð¥2 â [ð¥1 , ð¥3 ], there exists ðŒ â [0, 1] such that ð¥2 = ðŒð¥1 + (1 â ðŒ)ð¥2 Let 𥠯 = ð¥1 â ð¥2 + ð¥3 . Then 𥠯 â [ð¥1 , ð¥3 ] and there exists ðœ â [0, 1] such that ð¥Â¯ = ðœð¥1 + (1 â ðœ)ð¥2 Adding
( ) ð¥Â¯ + ð¥2 = (ðŒ + ðœ)ð¥1 + (1 â ðŒ) + (1 â ðœ) ð¥3
or ð¥1 â ð¥3 = (ðŒ + ðœ)(ð¥3 â ð¥1 ) which implies that ðŒ + ðœ = 1 and therefore ðœ = 1 â ðŒ. Since ð is convex ð (ð¥2 ) †ðŒð (ð¥1 ) + (1 â ðŒ)ð (ð¥2 ð (¯ ð¥) †ðœð (ð¥1 ) + (1 â ðœ)ð (ð¥2 ) = (1 â ðŒ)ð (ð¥1 ) + ðŒð (ð¥3 ) Adding ð (¯ ð¥) + ð (ð¥2 ) †ð (ð¥1 ) + ð (ð¥3 ) 152
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.129 Let ð¥1 , ð¥2 , ðŠ1 , ðŠ2 â â with ð¥1 < ð¥2 and ðŠ1 < ðŠ2 . Note that ð¥1 â ðŠ2 †ð¥2 â ðŠ2 †ð¥2 â ðŠ1 and therefore (Exercise 3.128) ( ) ð ð¥1 â ðŠ2 ) â (ð¥2 â ðŠ2 ) + (ð¥2 â ðŠ1 ) > ð (ð¥1 â ðŠ2 ) â ð (ð¥2 â ðŠ2 ) + ð (ð¥2 â ðŠ1 ) That is ð (ð¥1 â ðŠ1 ) > ð (ð¥1 â ðŠ2 ) â ð (ð¥2 â ðŠ2 ) + ð (ð¥2 â ðŠ1 ) Rearranging ð (ð¥2 â ðŠ2 ) â ð (ð¥1 â ðŠ2 ) > ð (ð¥2 â ðŠ1 ) â ð (ð¥1 â ðŠ1 ) as required. 3.130 A functional is aï¬ne if and only if inequalities (3.24) and (3.26) are satisï¬ed as equalities. 3.131 Since ð and ð are convex on ð ð (ðœx1 + (1 â ðœ)x2 ) †ðœð (x1 ) + (1 â ðœ)ð (x2 ) 1
2
1
2
ð(ðœx + (1 â ðœ)x ) †ðœð(x ) + (1 â ðœ)ð(x )
(3.47) (3.48)
for every x1 , x2 â ð and ðœ â [0, 1]. Adding (ð + ð)(ðœx1 + (1 â ðœ)x2 ) †ðœ(ð + ð)(x1 ) + (1 â ðœ)ð (x2 ) ð + ð is convex. Multiplying (3.47) by ðŒ ⥠0 ðŒð (ðœx1 + (1 â ðœ)x2 ) †ðŒ(ðœð (x1 ) + (1 â ðœ)ð (x2 )) = (ðœðŒð (x1 ) + (1 â ðœ)ðŒð (x2 )) ðŒð is convex. Moreover, if ð is strictly convex, ð (ðœx1 + (1 â ðœ)x2 ) < ðœð (x1 ) + (1 â ðœ)ð (x2 )
(3.49)
for every x1 , x2 â ð, x1 â= x2 and ðœ â (0, 1). Adding this to (3.48) (ð + ð)(ðœx1 + (1 â ðœ)x2 ) < ðœ(ð + ð)(x1 ) + (1 â ðœ)ð (x2 ) so that ð + ð is strictly convex. Multiplying (3.49) by ðŒ > 0 ðŒð (ðœx1 + (1 â ðœ)x2 ) < ðŒ(ðœð (x1 ) + (1 â ðœ)ð (x2 )) = (ðœðŒð (x1 ) + (1 â ðœ)ðŒð (x2 )) ðŒð is strictly convex. 3.132 x â epi (ð âš ð) ââ x â epi ð and x â epi ð That is epi (ð âš ð) = epi ð â© epi ð Therefore epi ð âš ð is convex (Exercise 1.162) and therefore ð is convex (Proposition 3.7). 153
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.133 If ð is convex ð (ðŒx1 + (1 â ðŒ)x2 ) †ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) Since ð is increasing ( ) ( ) ð ð (ðŒx1 + (1 â ðŒ)x2 ) †ð ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) ) ( ) ( †ðŒð ð (x1 ) + (1 â ðŒ)ð ð (x2 ) since ð is also convex. The concave case is proved similarly. 3.134 Let ð¹ = log ð . If ð¹ is convex, ð (x) = ðð¹ (x) is an increasing convex function of a convex function and is therefore convex (Exercise 3.133). 3.135 If ð is positive and concave, then log ð is concave (Exercise 3.51). Therefore log
1 = log 1 â log ð = â log ð ð
is convex. By the previous exercise (Exercise 3.134), this implies that 1/ð is convex. If ð is negative and convex, then âð is positive and concave, 1/ â ð is convex, and therefore 1/ð is concave. 3.136 Consider the identity ) ( ) ( ) ( ) ( ð ð (ð¥1 âš ð¥2 ) + ð ð (ð¥1 ⧠ð¥2 ) â ð ð (ð¥1 ) â ð ð (ð¥2 ) ) ( ) ( ) ( ) ) ) ( = ð ð (ð¥1 âš ð¥2 ) + ð ð (ð¥1 ⧠ð¥2 ) â ð ð (ð¥1 ) â ð ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) â ð (ð¥1 ) ( ) ( ) + ð ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) â ð (ð¥1 ) â ð ð (ð¥2 ) (3.50) Deï¬ne
( ) ( ) ( ) ( ) ð(ð¥1 , ð¥2 ) = ð ð (ð¥1 âš ð¥2 ) + ð ð (ð¥1 ⧠ð¥2 ) â ð ð (ð¥1 ) â ð ð (ð¥2 )
Then ð â ð is supermodular if ð is nonnegative deï¬nite and submodular if ð is nonpositive deï¬nite. Using the identity (3.50), ð can be decomposed into two components ð(ð¥1 , ð¥2 ) = ð1 (ð¥1 , ð¥2 ) + ð2 (ð¥1 , ð¥2 ) ( ) ( ) ( ) ð1 (ð¥1 , ð¥2 ) = ð ð (ð¥1 âš ð¥2 ) + ð ð (ð¥1 ⧠ð¥2 ) â ð ð (ð¥1 ) ( ) â ð ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) â ð (ð¥1 ) ( ) ( ) ð2 (ð¥1 , ð¥2 ) = ð ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) â ð (ð¥1 ) â ð ð (ð¥2 )
(3.51)
ð will deï¬nite if both components are deï¬nite. For any ð¥1 , ð¥2 â ð¥1 , let ð = ð (ð¥1 ⧠ð¥2 ), ð = ð (ð¥1 ) and ð = ð (ð¥1 âš ð¥2 ). Provided ð is monotone, ð lies between ð and ð. Substituting in (3.51) ð1 (ð¥1 , ð¥2 ) = ð(ð) + ð(ð) â ð(ð) â ð(ð + ð â ð) and Exercise 3.128 implies
{
ð1 (ð¥1 , ð¥2 ) = ð(ð) + ð(ð) â ð(ð) â ð(ð + ð â ð) Now consider ð2 .
} { } â¥ð convex if ð is â€0 concave
{ } { } ⥠supermodular ð (ð¥1 âš ð¥2 ) + ð (ð¥1 ⧠ð¥2 ) â ð (ð¥1 ) is ð (ð¥2 ) if ð is †submodular 154
(3.52)
Solutions for Foundations of Mathematical Economics and therefore since ð is increasing
{
ð2 (ð¥1 , ð¥2 ) =
c 2001 Michael Carter â All rights reserved
⥠0 if ð is supermodular †0 if ð is submodular
(3.53)
Together (3.52) and (3.53) gives the desired result. 3.137
1. Assume that ð is bounded above in a neighborhood of x0 . Then there exists a ball ðµ(ð¥0 ) and constant ð such that ð (x) †ð for every x â ðµ(ð¥0 ) Since ð is convex ð (ðŒx + (1 â ðŒ)x0 ) †ðŒð (x) + (1 â ðŒ)ð (x0 ) †ðŒð + (1 â ðŒ)ð (x0 )
(3.54)
2. Given x â ðµ(ð¥0 ) and ðŒ â [0, 1] let z = ðŒx + (1 â ðŒ)x0
(3.55)
Subtracting ð (x0 ) from (3.54) gives ð (z) â ð (x0 ) †ðŒ(ð â ð (x0 )) Rewriting (3.55) (1 â ðŒ)x0 = z â ðŒx (1 + ðŒ)x0 = z + ðŒ(2x0 â x) ðŒ 1 z+ (2x0 â x) x0 = 1+ðŒ 1+ðŒ 3. Note that (2x0 â x) = x0 â (x â x0 ) â ðµ(x0 ) so that ð (2x0 â x) †ð and therefore ð (x0 ) â€
ðŒ ðŒ 1 1 ð (z) + ð (2x0 â x) †ð (z) + ð 1+ðŒ 1+ðŒ 1+ðŒ 1+ðŒ
which implies (1 + ðŒ)ð (x0 ) †ð (z) + ðŒð ðŒ(ð (x0 ) â ð ) †ð (z) â ð (x0 ) 4. Combined with (3.56) we have ðŒ(ð (x0 ) â ð ) †ð (z) â ð (x0 ) †ðŒ(ð â ð (x0 )) or â£ð (z) â ð (x0 )⣠†ðŒ(ð â ð (x0 )) and therefore ð (z) â ð (x0 ) as z â x0 . ð is continuous. 155
(3.56)
Solutions for Foundations of Mathematical Economics 3.138
c 2001 Michael Carter â All rights reserved
1. Since ð is open, there exists a ball ðµð (x1 ) â ð. Let ð¡ = 1 + ð2 . Then x0 + ð¡(x1 â x0 ) â ðµð (ð¥1 ) â ð.
2. Let ð = ð¡â1 ð¡ ð. The open ball ðµð (x1 ) of radius ð centered on x1 is contained in ð . Therefore ð is a neighborhood of x1 . 3. Since ð is convex, for every y â ð ð (y) †(1 â ðŒ)ð (x) + ðŒð (z) †(1 â ðŒ)ð + ðŒð (z) †ð + ð (z) Therefore ð is bounded on ð . 3.139 The previous exercise showed that ð is locally bounded from above for every x â ð. To show that it is also locally bounded from below, choose some x0 â ð. There exists some ðµ(x0 and ð such that ð (x) †ð for every x â ðµ(x0 ) Choose some ð¥1 â ðµ(x0 ) and let x2 = 2x0 â x1 . Then x2 = 2x0 â x1 = x0 â (x1 â x0 ) â ðµ(x0 ) and ð (x2 ) †ð . Since ð¹ is convex ð (x) â€
1 1 ð (x1 ) + ð (x2 ) 2 2
and therefore ð (x1 ) ⥠2ð (x) â ð (x2 ) Since ð (x2 ) †ð , âð (x2 ) ⥠âð and therefore ð (x1 ) ⥠2ð (x) â ð so that ð is bounded from below. 3.140 Let ð be a convex function deï¬ned on an open convex set ð in a normed linear space, which is bounded from above in a neighborhood of a single point x0 â ð. By Exercise 3.138, ð is bounded above at every x â ð. This implies (Exercise 3.137) that ð is continuous at every x â ð. 3.141 Without loss of generality, assume 0 â ð. Assume ð has dimension ð and let x1 , x2 , . . . , xð be a basis for the subspace containing ð. Choose some ð > 0 small enough so that ð = conv {0, ðx1 , ðx2 , . . . , ðð¥ð } â ð Any x â ð is a convexâ combination of the points 0, x1 , x2 , . . . , xð and so there exists ðŒ0 , ðŒ1 , ðŒ2 , . . . , ðŒð ⥠0, ðŒð = 1 such that x = ðŒ0 0 + ðŒ1 x1 + â
â
â
+ ðŒð xð . By Jensenâs inequality ð (x) = ð (ðŒ0 0 + ðŒ1 x1 + â
â
â
+ ðŒð xð ) †ðŒ0 ð (0) + ðŒ1 ð (x1 ) + â
â
â
+ ðŒð ð (xð ) †max{ ð (0), ð (x1 ), . . . , ð (xð ) } Therefore, ð is bounded above on a neighbourhood of some x0 â int ð (which is nonempty by Exercise 1.229). By Proposition 3.8, ð is continuous on ð.
156
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.142 Clearly, if ð is convex, it is locally convex at every x â ð, where ð is the required neighborhood. To prove the converse, assume to the contrary that ð is locally convex at every x â ð but it is not globally convex. That is, there exists x1 , x2 â ð such that ð (ðŒx1 + (1 â ðŒ)x2 ) > ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) Let ) ( â(ð¡) = ð ð¡x1 + (1 â ð¡)x2 Local convexity implies that ð is continuous at every x â ð (Corollary 3.8.1), and therefore continuous on ð. Therefore, â is continuous on [0, 1]. By the continuous maximum theorem (Theorem 2.3), ð = arg max â(ð¡) xâ[x1 ,x2 ]
is nonempty and compact. Let ð¡0 = max ð . For every ð > 0, â(ð¡0 â ð) †â(ð¡0 ) and â(ð¡0 + ð) < â(ð¡0 ) Let x0 = ð¡0 x1 + (1 â ð¡0 )x2 and xð = (ð¡0 + ð)x1 + (1 â ð¡0 â ð)x2 Every neighborhood ð of x0 contains xâð , xð â [x1 , x2 ] with 1 1 1 1 ð (xâð ) + ð (xð ) = â(ð¡0 â ð) + â(ð¡0 + ð) < â(ð¡0 ) = ð (x0 ) = ð 2 2 2 2
(
1 1 xâð + xð 2 2
)
contradicting the local convexity of ð at x0 . 3.143 Assume ð is quasiconcave. That is for every x1 , x2 â ð and 0 †ðŒ †1 ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠min{ð (x1 ), (x2 )} Multiplying through by â1 reverses the inequality so that âð (ðŒx1 + (1 â ðŒ)x2 ) †â min{ð (x1 ), ð (x2 )} = max{âð (x1 ), âð (x2 )} which shows that âð is quasiconvex. The converse follows analogously. 3.144 Assume ð is concave, that is ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) for every x1 , x2 â ð and 0 †ðŒ †1 Without loss of generality assume that ð (x1 ) †ð (x2 ). Then ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) ⥠ðŒð (x1 ) + (1 â ðŒ)ð (x1 ) = ð (x1 ) = min{ð (x1 ), ð (x2 )} ð is quasiconcave. 3.145 Let ð : â â â. Choose any ð¥1 , ð¥2 in â with ð¥1 < ð¥2 . If ð is increasing, then ð (ð¥1 ) †ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) †ð (ð¥2 ) for every 0 †ðŒ †1. The ï¬rst inequality implies that ð (ð¥1 ) = min{ð (ð¥1 ), ð (ð¥2 )} †ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) 157
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
so that ð is quasiconcave. The second inequality implies that ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) †max{ð (ð¥1 ), ð (ð¥2 )} = ð (ð¥2 ) so that ð is also quasiconvex. Conversely, if ð is decreasing ð (ð¥1 ) ⥠ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) ⥠ð (ð¥2 ) for every 0 †ðŒ †1. The ï¬rst inequality implies that ð (ð¥1 ) = max{ð (ð¥1 ), ð (ð¥2 )} ⥠ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) so that ð is quasiconvex. The second inequality implies that ð (ðŒð¥1 + (1 â ðŒ)ð¥2 ) †max{ð (ð¥1 ), ð (ð¥2 )} = ð (ð¥2 ) so that ð is also quasiconcave. 3.146 âŸð (ð) = { x â ð : ð (x) †ð } = {x â ð : âð (x) ⥠âð} = â¿âð (âð) 3.147 For given ð and ð, choose any p1 and p2 in âŸð£ (ð). For any 0 †ðŒ †1, let ¯ = ðŒp1 + (1 â ðŒ)p2 . The key step is to show that any commodity bundle x which is p ¯ is also aï¬ordable at either p1 or p2 . Assume that x is aï¬ordable at p ¯, aï¬ordable at p that is x is in the budget set ¯x †ð } x â ð(¯ p, ð) = { x : p To show that x is aï¬ordable at either p1 or p2 , that is x â ð(p1 , ð) or x â ð(p2 , ð) assume to the contrary that xâ / ð(p1 , ð) and x â / ð(p2 , ð) This implies that p1 x > ð and p2 x > ð so that ðŒp1 x > ðŒð and (1 â ðŒ)p2 > (1 â ðŒ)ð Summing these two inequalities ¯ x = (ðŒp1 + (1 â ðŒ)p2 )x > ð p contradicting the assumption that x â ð(¯ p, ð). We conclude that ð(¯ p, ð) â ð(p1 , ð) ⪠ð(p2 , ð) Now ð£(¯ ð, ð) = sup{ ð¢(x) : x â ð(¯ p, ð) } †sup{ ð¢(x) : x â ð(p1 , ð) ⪠ð(p2 , ð) } â€ð ¯ â âŸð£ (ð) for every 0 †ðŒ †1. Thus, âŸð£ (ð) is convex and so ð£ is quasiconvex Therefore p (Exercise 3.146). 158
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.148 Since ð is quasiconcave ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠min{ð (x1 ), ð (x2 )} for every x1 , x2 â ð and 0 †ðŒ †1 Since ð is increasing ) ( ) ( ) ( ) ( ð ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ð( min{ð (x1 ), ð (x2 )}) ⥠min{ð ð (x1 ) , ð ð (x2 ) } ð â ð is quasiconcave. 3.149 When ð ⥠1, the function â(x) = ðŒ1 ð¥ð1 + ðŒ2 ð¥ð2 + . . . ðŒð ð¥ðð is convex (Example 3.58) as is ðŠ 1/ð . Therefore ð (x) = (â(x))
1/ð
is an increasing convex function of a convex function and is therefore convex (Exercise 3.133). 3.150 ð is a monotonic transformation of the concave function â(x) = x. 3.151 By Exercise 3.39, there exist linear functionals ðË and ðË and scalars ð and ð such that ð (x) = ðË(x) + ð and ð(x) = ðË(x) + ð The upper contour set â¿â (ð) = { ð¥ â ð : â(x) ⥠ð } ðË(ð¥) + ð ⥠ð} = {ð¥ â ð : ðË(ð¥) + ð = { ð¥ â âð : ðË(x) + ð ⥠ðË ð(x) + ðð } +
ð (x) ⥠ð â ðð } = { ð¥ â âð+ : ðË(x) â ðË which is a halfspace in ð and therefore convex. Similarly, the lower contour set âŸâ (ð) = { ð¥ â ð : â(x) ⥠ð } is also a halfspace and hence convex. Therefore â is both quasiconcave and quasiconvex. 3.152 For ð †0 â¿(ð) = { ð¥ â ð : â(x) ⥠0 } = ð which is convex. For ð > 0 â¿â (ð) = { ð¥ â ð : â(x) ⥠ð } ð (x) ⥠ð} ð(x) = { ð¥ â ð : ð (x) ⥠ðð(x) } = {ð¥ â ð :
= { ð¥ â ð : ð (x) â ðð(x) ⥠0 } is convex since ð â ðð = ð + ð(âð) is concave (Exercises 3.124 and 3.131). Since â¿â (ð) is convex for every ð, â is quasiconcave. 159
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.153 ð (x) ðË(x)
â(x) =
where ðË = 1/ð is positive and convex by Exercise 3.135. By the previous exercise, â is quasiconcave. 3.154 Let ð¹ = log ð . If ð¹ is concave, ð (x) = ðð¹ (x) is an increasing function of (quasi)concave function, and hence is quasiconcave (Exercise 3.148). 3.155 Let ð¹ (x) = log ð (x) =
ð â
ðŒð log ðð (x)
ð=1
As the sum of concave functions, ð¹ is concave (Exercise 3.131). By the previous exercise, ð is quasiconcave. ¯ = ðŒðœ1 + (1 â ðŒ)ðœ 2 ¯ are optimal solutions for ðœ1 , ðœ2 and ðœ 3.156 Assume x1 , x2 and x respectively. That is ð (x1 , ðœ 1 ) = ð£(ðœ 1 ) ð (x2 , ðœ 2 ) = ð£(ðœ 2 ) ¯ = ð£(ðœ) ¯ ð (¯ x, ðœ) Since ð is convex in 𜠯 = ð (¯ ¯ ð£(ðœ) x, ðœ) = ð (¯ x, ðŒðœ 1 + (1 â ðŒ)ðœ 2 ) †ðŒð (¯ x, ðœ 1 ) + (1 â ðŒ)ð (xâ , ðœ2 ) †ðŒð (x1 , ðœ1 ) + (1 â ðŒ)ð (x2 , ðœ 2 ) = ðŒð£(ðœ 1 ) + (1 â ðŒ)ð£(ðœ 2 ) ð£ is convex. 3.157 Assume to the contrary that x1 and x2 are distinct optimal solutions, that is â x2 , for some ðœ â Îâ , so that x1 , x2 â ð(ðœ), x1 = ð (x1 , ðœ) = ð (x2 , ðœ) = ð£(ðœ) ⥠ð (x, ðœ) for every x â ðº(ðœ) ¯ = ðŒx1 + (1 â ðŒ)x2 for ðŒ â (0, 1). Since ðº(ðœ) is convex, x ¯ is feasible. Since ð is Let x strictly quasiconcave ð (¯ x, ðœ) > min{ ð (x1 , ðœ), ð (x2 , ðœ) } = ð£(ðœ) contradicting the optimality of x1 and x2 . We conclude that ð(ðœ) is single-valued for every ðœ â Îâ . In other words, ð is a function. 3.158
1. The value function is ð£(ð¥0 ) =
sup ð (x)
xâÎ(ð¥0 )
where ð (x) =
â â ð¡=0
160
ðœ ð¡ ð (ð¥ð¡ , ð¥ð¡+1 )
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
and Î(ð¥0 ) = {x â ð â : ð¥ð¡+1 â ðº(ð¥ð¡ ), ð¡ = 0, 1, 2, . . . } Since an optimal policy exists (Exercise 2.125), the maximum is attained and ð£(ð¥0 ) = max ð (x) xâÎ(ð¥0 )
(3.57)
It is straightforward to show that â ð (x) is strictly concave and â Î(ð¥0 ) is convex Applying the Concave Maximum Theorem (Theorem 3.1) to (3.57), we conclude that the value function ð£ is strictly concave. 2. Assume to the contrary that xâ² and xâ²â² are distinct optimal plans, so that ð£(ð¥0 ) = ð (xâ² ) = ð (xâ²â² ) ¯ is feasible and ¯ = ðŒxâ² + (1 â ðŒ)xâ²â² . Since Î(ð¥0 ) is convex, x Let x ð (¯ x) > ðŒð (xâ² ) + (1 â ðŒ)ð (xâ²â² ) = ð (xâ² ) which contradicts the optimality of xâ² . We conclude that the optimal plan is unique. 3.159 We observe that â ð¢(ð¹ (ð) â ðŠ) is supermodular in ðŠ (Exercise 2.51) â ð¢(ð¹ (ð) â ðŠ) displays strictly increasing diï¬erences in (ð, ðŠ) (Exercise 3.129) â ðº(ð) = [0, ð¹ (ð)] is increasing. Applying Exercise 2.126, we can conclude that the optimal policy (ð0 , ð1â , ð2â , . . . ) is a monotone sequence. Since ð is compact, kâ is a bounded monotone sequence, which converges monotonically to some steady state ð â (Exercise 1.101). 3.160 Suppose there exists (xâ , yâ ) â ð à ð such that ð (x, yâ ) †ð (xâ , yâ ) †ð (xâ , y) for every x â ð and y â ð Let ð£ = ð (xâ , yâ ). Since ð (x, yâ ) †ð£ for every x â ð max ð (ð¥, yâ ) †ð£ xâð
and therefore min max ð (x, y) †max ð (x, yâ ) †ð£
yâð xâð
xâð
Similarly max min ð (x, y) ⥠ð£ xâð yâðŠ
Combining the last two inequalities, we have max min ð (x, y) ⥠ð£ ⥠min max ð (ð¥, y) xâð yâðŠ
yâð xâð
161
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Together with (3.28), this implies equality max min ð (x, y) = min max ð (x, y) xâð yâð
yâð xâð
Conversely, suppose that max min ð (x, y) = ð£ = min max ð (x, y) xâð yâð
yâð xâð
The function ð(x) = min ð (x, y) yâð
is a continuous function (Theorem 2.3) on a compact set ð. By the Weierstrass theorem (Theorem 2.2), there exists xâ which maximizes ð on ð, that is ð(xâ ) = min ð (xâ , y) = max ð(x) = max min ð (x, y) = ð£ yâð
xâð
xâð yâð
which implies that ð (xâ , y) ⥠ð£ for every y â ð Similarly, there exists y â ð such that ð (x, yâ ) †ð£ for every x â ð Combining these inequalities, we have ð (x, yâ ) †ð£ †ð (xâ , y) for every x â ð and y â ð In particular, we have ð (xâ , yâ ) †ð£ †ð (xâ , yâ ) so that ð£ = ð (xâ , yâ ) as required. 3.161 For any ð¥ â ð and ðŠ â ð , let ð(ð¥) = min ð (ð¥, ðŠ) and â(ðŠ) = max ð (ð¥, ðŠ) ðŠâð
ð¥âð
Then ð(ð¥) = min ð (ð¥, ðŠ) †max ð (ð¥, ðŠ) = â(ðŠ) ðŠâð
ð¥âð
and therefore max ð(ð¥) †max â(ðŠ) ð¥âð
ðŠâð
That is max min ð (ð¥, ðŠ) †ððŠ max ð (ð¥, ðŠ) ð¥
ðŠ
ð¥
162
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.162 Clearly ð (ð¥) = ð¥ð his homogeneous of degree ð. Conversely assume ð is homogeneous of degree ð, that is ð (ð¡ð¥) = ð¡ð ð (ð¥) Letting ð¥ = 1 ð (ð¡) = ð¡ð ð (1) Setting ð (1) = ðŽ â â and interchanging ð¥ and ð¡ yields the result. 3.163 1/ð
ð (ð¡x) = (ð1 (ð¡ð¥1 )ð + ð2 (ð¡ð¥2 )ð + . . . ðð (ð¡ð¥ð )ð ) 1/ð
= ð¡ (ð1 ð¥ð1 + ð2 ð¥ð2 + . . . ðð ð¥ðð ) = ð¡ð (x) 3.164 For ðœ â â++
â(ðœð¡) = ð (ðœð¡x0 ) = ðœ ð ð (ð¡x0 ) = ðœ ð â(ð¡) 3.165 Suppose that xâ minimizes the cost of producing output ðŠ at prices w. That is wð xâ †wð x
for every x â ð (ðŠ)
It follows that ð¡wð xâ †ð¡wð x
for every x â ð (ðŠ)
for every ð¡ > 0, verifying that xâ minimizes the cost of producing ðŠ at prices ð¡w. Therefore ð(ð¡w, ðŠ) = (ð¡w)xâ = ð¡(wð xâ ) = ð¡ð(w, ðŠ) ð(w, ðŠ) homogeneous of degree one in input prices w. 3.166 For given prices w, let xâ minimize the cost of producing one unit of output, so that ð(w, 1) = wð xâ . Clearly ð (xâ ) = 1 where ð is the production function. Now consider any output ðŠ. Since ð is homogeneous ð (ðŠxâ ) = ðŠð (xâ ) = ðŠ Therefore ðŠxâ is suï¬cient to produce ðŠ, so that ð(w, ðŠ) †wð (ðŠxâ ) = ðŠwð xâ = ðŠð(w, 1) Suppose that ð(w, ðŠ) < wð (ðŠxâ ) = ðŠð(w, 1) Then there exists xâ² such that ð (xâ² ) = ðŠ and wð xâ² < wð (ðŠxâ ) which implies that w
ð
(
xâ² ðŠ
)
< wð xâ = ð(w, 1) 163
Solutions for Foundations of Mathematical Economics Since ð is homogeneous
( ð
xâ² ðŠ
) =
c 2001 Michael Carter â All rights reserved
1 ð (xâ² ) = 1 ðŠ
Therefore, xâ² is a lower cost method of producing one unit of output, contradicting the deï¬nition of xâ . We conclude that ð(w, ðŠ) = ðŠð(w, 1) ð(w, ðŠ) is homogeneous of degree one in ðŠ. 3.167 If the consumerâs demand is invariant to proportionate changes in all prices and income, so also will the derived utility. More formally, suppose that xâ maximizes utility at prices p and income ð, that is xâ â¿ x
for every x â ð(p, ð)
Then ð£(p, ð) = ð¢(xâ ) Since ð(ð¡p, ð¡ð) = ð(p, ð) xâ â¿ x
for every x â ð(ð¡p, ð¡ð)
and ð£(ð¡p, ð¡ð) = ð¢(xâ ) = ð£(p, ð) 3.168 Assume ð is homogeneous of degree one, so that ð (ð¡x) = ð¡ð (x)
for every ð¡ > 0
Let (x, ðŠ) â epi ð , so that ð (x) †ðŠ For any ð¡ > 0 ð (ð¡x) = ð¡ð (x) †ð¡ðŠ which implies that (ð¡x, ð¡ðŠ) â epi ð . Therefore epi ð is a cone. Conversely assume epi ð is a cone. Let x â ð and deï¬ne ðŠ = ð (x). Then (x, ðŠ) â epi ð and therefore (ð¡x, ð¡ðŠ) â epi ð so ð (ð¡x) †ð¡ðŠ Now suppose to the contrary that ð (ð¡x) = ð§ < ð¡ðŠ = ð¡ð (x)
(3.58)
Then (ð¡x, ð§) â epi ð . Since epi ð is a cone, we must have (x, ð§/ð¡) â epi ð so that ð (x) â€
ð§ ð¡
and ð¡ð (x) †ð§ = ð (ð¡x) contradicting (3.58). We conclude that ð (ð¡x) = ð¡ð (x) for every ð¡ > 0 164
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.169 Take any x1 and x2 in ð and let ðŠ1 = ð (x1 ) > 0 and ðŠ2 = ð (x2 ) > 0 Since ð is homogeneous of degree one, ( ) ( ) x1 x2 ð =ð =1 ðŠ1 ðŠ2 Since ð is quasiconcave
( ) x2 x1 + (1 â ðŒ) ð ðŒ â¥1 ðŠ1 ðŠ2
for every 0 †ðŒ †1. Choose ðŒ = ðŠ1 /(ðŠ1 + ðŠ2 ) so that (1 â ðŒ) = ðŠ2 /(ðŠ1 + ðŠ2 ). Then ( ) x1 x2 ð + â¥1 ðŠ1 + ðŠ2 ðŠ1 + ðŠ2 Again using the homogeneity of ð , this implies ð (x1 + x2 ) ⥠ðŠ1 + ðŠ2 = ð (x1 ) + ð (x2 ) 3.170 Let ð â ð¹ (ð) be a strictly positive deï¬nite, quasiconcave functional which is homogeneous of degree one. For any x1 , x2 in ð and 0 †ðŒ †1 ðŒx1 , (1 â ðŒ)x2 in ð and therefore ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ð (ðŒx1 ) + ð ((1 â ðŒ)x2 ) since ð is superadditive (Exercise 3.169). But ð (ðŒx1 ) = ðŒð (x1 ) ð ((1 â ðŒ)x2 ) = (1 â ðŒ)ð (x2 ) by homogeneity. Substituting in (3.58), we conclude that ð (ðŒx1 + (1 â ðŒ)x2 ) ⥠ðŒð (x1 ) + (1 â ðŒ)ð ((1 â ðŒ)x2 ) ð is concave. 3.171 Assume that ð is strictly positive deï¬nite, quasiconcave and homogeneous of degree ð, 0 < ð < 1. Deï¬ne â(x) = (ð (x))
1/ð
Then â is quasiconcave (Exercise 3.148. Further, for every ð¡ > 0 1/ð
â(ð¡x) = (ð (ð¡x)) ( )1/ð = ð¡ð ð (x) = ð¡ (ð (x))
1/ð
= ð¡â(x) so that â is homogeneous of degree 1. By Exercise 3.170, â is concave. ð (x) = (â(x))
ð
That is ð = ð â â where ð(ðŠ) = ðŠ ð is monotone and concave provided ð †1. By Exercise 3.133, ð = ð â â is concave. 165
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.172 Continuity is a necessary and suï¬cient condition for the existence of a utility function representing â¿ (Remark 2.9). Suppose ð¢ represents the homothetic preference relation â¿. For any x1 , x2 â ð ð¢(x1 ) = ð¢(x2 ) =â x1 ⌠x2 =â ð¡x1 ⌠ð¡x2 =â ð¢(ð¡x1 ) = ð¢(ð¡x2 ) for every ð¡ > 0 Conversely, if ð¢ is a homothetic functional, x1 ⌠x2 =â ð¢(x1 ) = ð¢(x2 ) =â ð¢(ð¡x1 ) = ð¢(ð¡x2 ) =â ð¡x1 ⌠ð¡x2 for every ð¡ > 0 3.173 Suppose that ð = ð â â where ð is strictly increasing and â is homogeneous of degree ð. Then ( )1/ð Ë â(x) = â(x) Ë where is homogeneous of degree one and ð = ðË â â ( ) ðË(ðŠ) = ð ðŠ ð ) is increasing. 3.174 Assume x1 , x2 â ð with ð (x1 ) = ð(â(x1 )) = ð(âx2 )) = ð (x2 ) Since ð is strictly increasing, this implies that â(x1 ) = â(x2 ) Since â is homogeneous â(ð¡x1 ) = ð¡ð â(x1 ) = ð¡ð â(x2 ) = â(ð¡x2 ) for some ð. Therefore ð (ð¡x1 ) = ð(â(ð¡x1 )) = ð(â(ð¡x2 )) = ð (ð¡x2 ) 3.175 Let x0 â= 0 be any point in ð, and deï¬ne ð : â â â by ð(ðŒ) = ð (ðŒx0 ) Since ð is strictly increasing, so is ð and therefore ð has a strictly increasing inverse ð â1 . Let â = ð â1 â ð so that ð = ð â â. We need to show that â is homogeneous. For any x â ð, there exists ðŒ such that ð(ðŒ) = ð (ðŒx0 ) = ð (x) that is ðŒ = â(x) = ð â1 (ð (x)). Since ð is homothetic ð(ð¡ðŒ) = ð (ð¡ðŒx0 )ð (ð¡x) for every ð¡ > 0 and therefore â(ð¡x) = ð â1 (ð (ð¡x)) = ð â1 (ð (ð¡ðŒx0 )) = ð â1 ð(ð¡ðŒ) = ð¡ðŒ = ð¡â(x) â is homogeneous of degree one. 166
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.176 Let ð be the production function. If ð is homothetic, there exists (Exercise 3.175) a linearly homogeneous function â and strictly increasing function ð such that ð = ð ââ. ð(w, ðŠ) = min{ wð x : ð (x) ⥠ðŠ } x
= min{ wð x : ð(â(x)) ⥠ðŠ } x
= min{ wð x : â(x) ⥠ð â1 (ðŠ) } x
= ð â1 (ðŠ)ð(w, 1) by Exercise 3.166. 3.177 Let ð : ð â â be positive, strictly increasing, homothetic and quasiconcave. By Exercise 3.175, there exists a linearly homogeneous function â : ð â â and strictly increasing function ð â ð¹ (ð
) such that ð = ð â â. â = ð â1 â ð is positive, quasiconcave (Exercise 3.148) and homogeneous of degree one. By Proposition 3.12, â is concave and therefore ð = ð â â is concaviï¬able. 3.178 Since ð»ð (ð) is a supporting hyperplane to ð at x0 , then ð (x0 ) = ð and either ð (x) ⥠ð = ð (x0 ) for every x â ð or ð (x) †ð = ð (x0 ) for every x â ð 3.179 Suppose to the contrary that y = (â, ð) â int ðŽ â© ðµ. Then y â¿ yâ . By strict convexity yðŒ = ðŒy + (1 â ðŒ)yâ â» yâ for every ðŒ â (0, 1) Since y â int ðŽ, yðŒ â ðŽ for ðŒ suï¬ciently small. That is, there exists some ðŒ such that yðŒ is feasible and yðŒ â» yâ , contradicting the optimality of yâ . 3.180 For notational simplicity, let ð be the linear functional which separates ðŽ and ðµ in Example 3.77. ð (y) measure the cost of the plan y = (â, ð), that is ð (y) = ð€â + ðð. Assume to the contrary there exists a preferred lifestyle in ð, that is there exists some y = (â, ð) â ð such that y â» yâ = (ââ , ð â ). Since y â ðµ, ð (y) ⥠ð (yâ ) by (3.29). On the other hand, y â ð which implies that ð (y) †ð (yâ ). Consequently, ð (y) = ð (yâ ). By continuity, there exists some ðŒ < 1 such that ðŒy â» yâ which implies that ðŒy â ðµ. By linearity ð (ðŒy) = ðŒð (y) < ð (y) = ð (yâ ) = ðŒ contrary to (3.29). This contradiction establishes that yâ is the best choice in budget set ð. 3.181 By Proposition 3.7, epi ð is a convex set in ð à â with (x0 , ð (x0 )) a point on its boundary. By Corollary 3.2.2 of the Separating Hyperplane Theorem, there exists linear a functional ð â (ð à â)â² such that ð(x, ðŠ) ⥠ð(x0 , ð (x0 )) for every (x, ðŠ) â epi ð 167
(3.59)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ð can be decomposed into two components (Exercise 3.47) ð(x, ðŠ) = âð(x) + ðŒðŠ The assumption that x0 â int ð ensures that ðŒ > 0 and we can normalize so that ðŒ = 1. Substituting in (3.59) âð(x) + ð (x) ⥠âð(x0 ) + ð (x0 ) ð (x) ⥠ð (x0 ) + ð(x â x0 ) for every x â ð. 3.182 By Exercise 3.72, there exists a unique point x0 â ð such that (x0 â y)ð (x â x0 ) ⥠0 for every x â ð Deï¬ne the linear functional (Exercise 3.64) ð (x) = (x0 â y)ð x and let ð = ð (x0 ). For all x â ð ð (x) â ð (x0 ) = ð (x â x0 ) = (x0 â y)ð (x â x0 ) ⥠0 and therefore ð (x) ⥠ð (x0 ) = ð for every x â ð Furthermore 2
ð (x0 ) â ð (y) = ð (x0 â y) = (x0 â y)ð (x0 â y) = â¥x0 â y⥠> 0 since y â= x0 . Therefore ð (x0 ) > ð (y) and ð (y) < ð †ð (x) for every x â ð 3.183 If y â b(ð), y â ð ð and there exists a sequence of points {yð } â ð ð converging to y (Exercise 1.105). That is, there exists a sequence of nonboundary points {yð } â /ð converging to y. For every point yð , there is a linear functional ð ð â ð â and ðð such that ð ð (yð ) < ðð †ð ð (x)
for every x â ð
Deï¬ne ð ð = ð ð / â¥ð ð â¥. By construction, the sequence of linear functionals ð ð belong to the unit ball in ð â (since â¥ð ⥠= 1). Since ð â is ï¬nite dimensional, the unit ball is compact as so ð ð has a convergent subsequence with limit ð such that ð (y) †ð (x)
for every ð¥ â ð
ð (y) †ð (x)
for every ð¥ â ð
A fortiori
3.184 There are two possible cases. yâ / ð By Exercise 3.182, there exists a hyperplane which separates y and ð which a fortiori separates y and ð, that is ð (y) †ð (x)
for every x â ð 168
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
y â ð Since y â / ð, y must be a boundary point of ð. By the previous exercise, there exists a supporting hyperplane at y, that is there exists a continuous linear functional ð â ð â such that ð (y) †ð (x) 3.185
for every x â ð
1. ð (ð) â â.
2. ð (ð) is convex and hence an interval (Exercise 1.160. 3. ð (ð) is open in â (Proposition 3.2). 3.186 ð is nonempty and convex and 0 â / ð. (Otherwise, there exists x â ðŽ and y â ðµ such that 0 = y + (âx) which implies that x = y contradicting the assumption that ðŽ â© ðµ = â
.) Thus there exists a continuous linear functional ð â ð â such that ð (y â x) ⥠ð (0) = 0
for every x â ðŽ, y â ðµ
so that ð (x) †ð (y) for every x â ðŽ, y â ðµ Let ð = supxâðŽ ð (x). Then ð (x) †ð †ð (y) for every x â ðŽ, y â ðµ By Exercise 3.185, ð (int ðŽ) is an open interval in (ââ, ð], hence ð (int ðŽ) â (ââ, ð), so that ð (x) < ð for every x â int ðŽ. Similarly, ð (int ðµ) > ð and ð (x) < ð < ð (y) for every x â int ðŽ, y â int ðµ 3.187 Since int ðŽ â© ðµ = â
, int ðŽ and ðµ can be separated. That is, there exists a continuous linear functional ð â ð â and a number ð such that ð (x) †ð †ð (y)
for every x â ðŽ, y â int ðµ
which implies that ð (x) †ð †ð (y)
for every x â ðŽ, y â ðµ
since ðâ€
inf
yâint ðµ
ð (y) = inf ð (y) yâðµ
Conversely, suppose that ðŽ and ðµ can be separated. That is, there exists ð â ð â such that ð (x) †ð †ð (y)
for every x â ðŽ, y â ðµ
Then ð (int ðŽ) is an open interval in [ð, â), which is disjoint from the interval ð (ðµ) â (ââ, ð]. This implies that int ðŽ â© ðµ = â
. 3.188 Since x0 â b(ð), {x0 } â© int ð = â
and int ð â= â
. By Corollary 3.2.1, {x0 } and ð can be separated, that is there exist ð â ð â such that ð (x0 ) †ð (x) for every x â ð
169
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.189 Let x â ð¶. Since ð¶ is a cone, ðx â ð¶ for every ð ⥠0 and therefore ð (ðx) ⥠ð or ð (x) ⥠ð/ð
for every ð ⥠0
Taking the limit as ð â â implies that ð (x) ⥠0
for every x â ð¶
3.190 First note that 0 â ð and therefore ð (0) = 0 †ð so that ð ⥠0. Suppose that there exists some z â ð for which ð (z) = ð â= 0. By linearity, this implies ð(
2ð 2ð z) = ð (z) = 2ð > ð ð ð
which contradicts the requirement ð (z) †ð for every z â ð 3.191 By Corollary 3.2.1, there exists ð â ð â such that ð (z) †ð †ð (x)
for every x â ð, z â ð
By Exercise 3.190 ð (z) = 0
for every z â ð
ð (x) ⥠0
for every x â ð
and therefore
Therefore ð is contained in the hyperplane ð»ð (0) which separates ð from ð. 3.192 Combining Theorem 3.2 and Corollary 3.2.1, there exists a hyperplane ð»ð (ð) such that ð (x) †ð †ð (y)
for every x â ðŽ, y â ðµ
and such that ð (x) < ð †ð (y)
for every x â int ðŽ, y â ðµ
Since int ðŽ â= â
, there exists some x â int ðŽ with ð (x) < ð. Hence ðŽ â ð â1 (ð) = ð»ð (ð). 3.193 Follows directly from the basic separation theorem, since ðŽ = int ðŽ and ðµ = int ðµ. 3.194 Let ð = ðµ â ðŽ. Then 1. ð is a nonempty, closed, convex set (Exercise 1.203). 2. 0 â / ð. There exists a continuous linear functional ð â ð â such that ð (x) ⥠ð > ð (0) = 0 170
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
ðŽ
ðµ
Figure 3.2: ðŽ and ðµ cannot be strongly separated. for every z â ð (Exercise 3.182). For every x â ðŽ, y â ðµ, z = y â x â ð and ð (z) = ð (y) â ð (x) ⥠ð > 0 or ð (x) + ð †ð (y) which implies that sup ð (x) + ð †inf ð (y) yâðµ
xâðŽ
and sup ð (x) < inf ð (y)
xâðŽ
yâðµ
3.195 No. See Figure 3.2. 3.196
1. Assume that there exists a convex neighborhood ð â 0 such that (ðŽ + ð ) â© ðµ = â
Then (ðŽ + ð ) is convex and ðŽ â int (ðŽ + ð ) â= â
and int (ðŽ + ð ) â© ðµ = â
. By Corollary 3.2.1, there exists continuous linear functional such that ð (x + u) †ð (y)
for every x â ðŽ, u â ð, y â ðµ
Since ð (ð ) is an open interval containing 0, there exists some u0 with ð (u0 ) = ð > 0. ð (x) + ð †ð (y)
for every x â ðŽ, y â ðµ
which implies that sup ð (x) < inf ð (y) yâðµ
xâðŽ
Conversely, assume that ðŽ and ðµ can be strongly separated. That is, there exists a continuous linear functional ð â ð â and number ð > 0 such that ð (x) †ð â ð < ð + ð †ð (y) for every x â ðŽ, y â ðµ Let ð = { ð¥ â ð : â£ð (ð¥)⣠< ð }. ð is a convex neighborhood of 0 such that (ðŽ + ð ) â© ðµ = â
. 171
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. Let ðŽ and ðµ be nonempty, disjoint, convex subsets in a normed linear space ð with ðŽ compact and ðµ closed. By Exercise 1.208, there exists a convex neighborhood ð â 0 such that (ðŽ + ð ) â© ðµ = â
. By the previous part, ðŽ and ðµ can be strongly separated. 3.197 Assume ð(ðŽ, ðµ) = inf{ â¥x â y⥠: x â ðŽ, y â ðµ } = 2ð > 0. Let ð = ðµð (0) be the open ball around 0 of radius ð. For every x â ðŽ, u â ð, y â ðµ â¥x + (âu) â y⥠= â¥x â y â u⥠⥠â¥x â y⥠â â¥u⥠so that ð(ðŽ + ð, ðµ) = inf â¥x + (âu) â y⥠⥠inf (â¥x â y⥠â â¥uâ¥) x,u,y
x,u,y
⥠inf â¥x â y⥠â sup â¥uâ¥) x,y
u
= 2ð â ð =ð>0 Therefore (ðŽ + ð ) â© ðµ = â
and so ðŽ and ðµ can be strongly separated. Conversely, assume that ðŽ and ðµ can be strongly separated, so that there exists a convex neighborhood ð of 0 such that (ðŽ + ð ) â© ðµ = â
. Therefore, there exists ð > 0 such that ðµð (0) â ð and ðŽ + ðµð â© ðµ = â
This implies that ð(ðŽ, ðµ) = inf{ â¥x â y⥠: x â ðŽ, y â ðµ } > ð > 0 3.198 Take ðŽ = {y} and ðµ = ð in Proposition 3.14. There exists ð â ð â such that ð (y) < ð †ð (x)
for every x â ð
By Corollary 3.2.3, ð = 0. 3.199
1. Consider the set ð = { ð (ð¥), âð1 (ð¥), âð2 (ð¥), . . . , âðð (ð¥) : ð¥ â ð } ð is the image of a linear mapping from ð to ð = âð+1 and hence is a subspace of âð+1 .
2. By hypothesis, the point e0 = (1, 0, 0, . . . , 0) â âð+1 does not belong to ð. Otherwise, we have an ð¥ â ð such that ðð (ð¥) = 0 for every ð but ð (ð¥) = 1. 3. By the previous exercise, there exists a linear functional ð â ð â such that ð(e0 ) > 0 ð(z) = 0
for every z â ð
4. In other words, there exists a vector ð = (ð0 , ð1 , . . . , ðð ) â ð = â(ð+1)â such that ðe0 > 0 ðz = 0
(3.60) for every z â ð 172
(3.61)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Equation (3.61) states that ðz = ð0 z0 + ð1 z1 + â
â
â
+ ðð zð = 0
for every z â ð
That is, for every ð¥ â ð, ð0 ð (x) â ð1 ð1 (x) â ð2 ð2 (x) â . . . â ðð ðð (x) = 0 5. Inequality (3.60) establishes that ð0 > 0. Without loss of generality we can normalize so that ð0 = 1. 6. Therefore ð (ð¥) =
ð â
ðð ðð (ð¥)
ð=1
3.200 For every x â ð, ðð (x) = 0, ð = 1, 2 . . . ð and therefore ð (x) =
ð â
ðð ðð (x) = 0
ð=1
3.201 The set ð = { ð1 (ð¥), ð2 (ð¥), . . . , ðð (ð¥) : ð¥ â ð } is a closed subspace in âð . If the system is inconsistent, c = (ð1 , ð2 , . . . , ðð ) â / ð. By Exercise 3.198, there exists a linear functional ð on âð such ð(z) = 0 for every z â ð ð(c) > 0 That is, there exist numbers ð1 , ð2 , . . . , ðð such that ð â
ðð ðð (x) = 0
ð=1
and ð â
ðð ðð > 0
ð=1
which contradicts the hypothesis ð â
ðð ðð = 0 =â
ð=1
ð â
ðð ðð = 0
ð=1
Conversely, if for some x â ð ðð (x) = ðð
ð = 1, 2, . . . , ð
then ð â
ðð ðð (x) =
ð=1
ð â
ðð ðð
ð=1
and ð â
ðð ðð = 0 =â
ð=1
ð â ð=1
173
ðð ðð = 0
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Ë = { x â ðŸ : â¥x⥠= 1 } is 3.202 The set ðŸ 1 â compact (the unit ball is compact if and only if ð is ï¬nite-dimensional) â convex (which is why we need the 1 norm) By Proposition 3.14, there exists a linear functional ð â ð â such that Ë ËâðŸ for every x for every x â ð
ð (Ë x) > 0 ð (x) = 0
Ë Then Ë = x/ â¥xâ¥1 â ðŸ. For any x â ðŸ, x â= 0, deï¬ne x Ë ) = â¥xâ¥1 ð (Ë ð (x) = ð (â¥xâ¥1 x x) > 0 3.203
1. Let ðŽ = { (x, ðŠ) : ðŠ ⥠ð(x), x â ð } ðµ = { (x, ðŠ) : ðŠ = ð0 (x), x â ð } ðŽ is the epigraph of a convex functional and hence convex. ðµ is a subspace of ð = ð à â and also convex.
2. Since ð is convex, int ðŽ â= â
. Furthermore ð0 (x) †ð(x) =â int ðŽ â© ðµ = â
3. By Exercise 3.2.3, there exists linear functional ð â ð â such that ð(x, ðŠ) ⥠0 ð(x, ðŠ) = 0
for every (x, ðŠ) â ðŽ for every (x, ðŠ) â ðµ
There exists ðŠ such that ðŠ > ð(0) and therefore (0, ðŠ) â int ðŽ and ð(0, ðŠ) > 0. Therefore ð(0, 1) =
1 ð(0, 1) > 0 ðŠ
4. Let ð â ð â be deï¬ned by 1 ð (x) = â ð(x, 0) ð where ð = ð(0, 1). Since ð(x, 0) = ð(x, ðŠ) â ð(0, ðŠ) = ð(x, ðŠ) â ððŠ 1 1 ð (x) = â (ð(x, ðŠ) â ððŠ) = â ð(x, ðŠ) + ðŠ ð ð for every ðŠ â â 5. For every x â ð 1 ð (x) = â ð(x, ð0 (x)) + ð0 (x) ð = ð0 (x) since ð(x, ð0 (x)) = 0 for every x â ð. Thus ð is an extension of ð0 . 174
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
6. For any x â ð, let ðŠ = ð(x). Then (x, ðŠ) â ðŽ and ð(x, ðŠ) ⥠0. Therefore 1 ð (x) = â ð(x, ðŠ) + ðŠ ð 1 = â ð(x, ðŠ) + ð(x) ð †ð(x) Therefore ð is bounded by ð as required. 3.204 Let ð â ð â be deï¬ned by ð(x) = â¥ð0 â¥ð â¥x⥠Then ð0 (x) †ð(x) for all x â ð. By the Hahn-Banach theorem (Exercise 3.15), there exists an extension ð â ð â such that ð (x) †ð(x) = â¥ð0 â¥ð â¥x⥠Therefore â¥ð â¥ð = sup â¥ð (x)⥠= â¥ð0 â¥ð â¥xâ¥=1
3.205 If x0 = 0, any bounded linear functional will do. Therefore, assume x0 â= 0. On the subspace lin {x0 } = {ðŒx0 : ðŒ â â}, deï¬ne the function ð0 (ðŒx0 ) = ðŒ â¥x0 ⥠ð0 is a bounded linear functional on lin {x0 } with norm 1. By the previous part, ð0 can be extended to a bounded linear functional ð â ð â with the same norm, that is â¥ð ⥠= 1 and ð (x0 ) = â¥x0 â¥. 3.206 Since x1 â= x2 , x1 â x2 â= 0. There exists a bounded linear functional such that ð (x1 â x2 ) = â¥x1 â x2 ⥠â= 0 so that ð (x1 ) â= ð (x2 ) 3.207
1.
â ð is a complete lattice (Exercise 1.179).
â The intersection of any chain is â nonempty (since ð is compact) â a face (Exercise 1.179) Hence every chain has a minimal element. â By Zornâs lemma (Remark 1.5), ð has a minimal element ð¹0 . 2. Assume to the contrary that ð¹0 contains two distinct elements x1 , x2 . Then (Exercise 3.206) there exists a continuous linear functional ð â ð â such that ð (x1 ) â= ð (x2 ) Let ð be in the minimum value of ð (x) on ð¹0 and let ð¹1 be the set on which it attains this minimum. (Since ð¹0 is compact, ð is well-deï¬ned and ð¹1 is nonempty. That is ð = min{ ð (ð¥) : x â ð¹0 } ð¹1 = { x â ð : ð (x) = ð } 175
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Now ð¹1 â ð¹0 since ð (x1 ) â= ð (x2 ). To show that ð¹1 is a face of ð¹0 , assume that ðŒx+(1âðŒ)y â ð¹1 for some x, y â ð¹0 . Then ð = ð (ðŒx + (1 â ðŒ)y) = ðŒð (x) + (1 â ðŒ)ð (y) = ð. Since x, y â ð¹0 , this implies that ð (x) = ð (y) = ð so that x, y â ð¹1 . Therefore ð¹1 is a face. We have shown that, if ð¹0 contains two distinct elements, there exists a smaller face ð¹1 â ð¹0 , contradicting the minimality of ð¹0 . We conclude that ð¹0 comprises a single element x0 . 3. ð¹0 = {x0 } which is an extreme point of ð. 3.208 Let ð» = ð»ð (ð) be a supporting hyperplane to ð. Without loss of generality assume ð (x) †ð for every x â ð
(3.62)
and there exists some xâ â ð such that ð (xâ ) = ð That is ð is maximized at xâ . Version 1 By the previous exercise, ð achieves its maximum at an extreme point. That is, there exists an extreme point x0 â ð such that ð (x0 ) ⥠ð (x) for every x â ð In particular, ð (x0 ) ⥠ð (xâ ) = ð. But (3.62) implies ð (x0 ) †ð. Therefore, we conclude that ð (x0 ) = ð and therefore x0 â ð». Version 2 The set ð» â© ð is a nonempty, compact, convex subset of a linear space. Hence, by Exercise 3.207, ð» â© ð contains an extreme point, say x0 . We show that x0 is an extreme point of ð. Assume not, that is assume that there exists x1 , x2 â ð such that x0 = ðŒx1 + (1 â ðŒ)x2 for some ðŒ â (0, 1). Since x0 is an extreme point of ð» â© ð, at least one of the points x1 , x2 must lie outside ð». Assume x1 â / ð» which implies that ð (x1 ) < ð. Since ð (x2 ) †ð ð (x0 ) = ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) < ð
(3.63)
However, since x0 â ð» â© ð, we must have ð (x0 ) = ð which contradicts (3.63). Therefore x0 is an extreme point of ð. In fact, we have shown that every extreme point of ð» â© ð must be an extreme point of ð. 3.209 Let ðË denote the closed, convex hull of the extreme points of ð. (The closed, convex hull of a set is simply the closure of the convex hull.) Clearly ðË â ð and it remains to show that ðË contains all of ð. Ë By the Strong Separation Assume not. That is, assume ðË â ð and let x0 â ð â ð. Theorem, there exists a linear functional ð â ð â such that ð (x0 ) > ð (x) for every x â ðË 176
(3.64)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
On the other hand, by Exercise 3.16, ð attains its maximum at an extreme point of ð. That is, there exists x1 â ðË such that ð (x1 ) ⥠ð (x) for every x â ð In particular ð (x1 ) ⥠ð (x0 ) Ë since x0 â ðË â ð. This contradicts (3.64) since x1 â ð. Thus our assumption that ð â ðË yields a contradiction. We conclude that ð = ðË 3.210
1. (a) ð is compact and convex, since it is the product of compact, convex sets (Proposition 1.2, Exercise 1.165). âð âð (b) Since x â ð=1 conv ðð , there exist xð â conv ðð such that x = ð=1 xð . (x1 , x2 , . . . , xð ) â ð (x) so that ð (x) â= â
. (c) By the Krein-Millman theorem (or Exercise 3.207), ð (x) has an extreme point z = (z1 , z2 , . . . , zð ) such that â zð â conv ðð for every ð âð â ð=1 zð = x. since z â ð (x).
2. (a) Exercise 1.176 (b) Since ð > ð = dim ð, the vectors y1 , y2 , . . . , yð are linearly dependent (Exercise 1.143). Consequently, there exists numbers ðŒâ²1 , ðŒâ²2 , . . . , ðŒâ²ð , not all zero, such that ðŒâ²1 y1 + ðŒâ²2 y2 + â
â
â
+ ðŒâ²ð yð = 0 (Exercise 1.133). Let ðŒð =
ðŒâ²ð maxð â£ðŒð â£
Then â£ðŒð ⣠†1 for every ð and ðŒ1 y1 + ðŒ2 y2 + â
â
â
+ ðŒð yð = 0 (c) Since â£ðŒð ⣠†1, zð + ðŒð yð â conv ðð for every ð = 1, 2, . . . , ð. Furthermore ð â ð=1
z+ ð =
ð â
zð +
ð=1
ð â
ðŒð yð =
ð=1
ð â
zð = x
ð=1
Therefore, z+ â ð (x). Similarly, zâ â ð (x). (d) By direct computation z=
1 + 1 â z + z 2 2
which implies that z is not an extreme point of ð (x), contrary to our assumption. This establishes that at least ð â ð zð are extreme points of the corresponding conv ðð . 177
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
4 3 (0, 2.5)
conv ð2 2
(.5, 2) P(x)
1 0
1
2
3
4
conv ð1 Figure 3.3: Illustrating the proof of the Shapley Folkman theorem. 3. Every extreme point of conv ðð is an element of ðð . 3.211 See Figure 3.3. 3.212 Let {ð1 , ð2 , . . . , ðð } be a â collection of subsets of an ð-dimensional ânonempty ð ð linear space and let x â â conv ð = conv ð . That is, there exists xð â ð ð ð=1 ð=1 conv ðð such that x = ðð=1 xð . By CarathÂŽeodoryâs theorem, there exists for every xð a ï¬nite number of points xð1 , xð2 , . . . , xððð such that xð â conv {xð1 , xð2 , . . . , xððð }. For every ð = 1, 2, . . . , ð, let ðËð = { xðð : ð = 1, 2, . . . , ðð } Then x=
ð â
xð ,
xð â conv ðËð
ð=1
â âË ðð . Moreover, the sets ðð are compact (in fact That is, x â conv ðËð = conv ï¬nite). By the previous exercise, there exists ð points zð â ðËð such that x=
ð â
zð ,
zð â conv ðËð
ð=1
and moreover zð â ðËð â ðð for at least ð â ð indices ð. 3.213 Let ð be a closed convex set in a normed linear space. Clearly, ð is contained in the intersection of all the closed halfspaces which contain ð. For any y â / ð, there exists a hyperplane which strongly separates {y} and ð. One of its closed halfspaces contains ð but not y. Consequently, y does not belong to the intersection of all the closed halfspaces containing ð. 3.214
1. Since ð â (ðŠ) is the intersection of closed, convex sets, it is closed and convex. Assume x is feasible, that is x â ð (ðŠ). Then wð x †(ðw, ðŠ) and x â ð â (ðŠ). That is, ð (ðŠ) â ð â (ðŠ).
178
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. Assume ð (ðŠ) is convex. For any x0 â / ð (ðŠ) there exists w such that wð x0 <
inf wð x = ð(w, ðŠ)
xâð (ðŠ)
by the Strong Separation Theorem. Monotonicity ensures that w ⥠0 and hence x0 â / ð â (ðŠ). 3.215 Assume x â ð (ðŠ) = ð â (ðŠ). That is ð(w) for every x wð x ⥠ðŠË Therefore, for any ð¡ â â+ ð¡wð x ⥠ð¡ðŠð(w) for every x which implies that ð¡x â ð â (ðŠ) = ð (ðŠ). 3.216 A polyhedron ð = { ð¥ â ð : ðð (ð¥) †ðð , ð = 1, 2, . . . , ð } ð â© = { x â ð : ðð (x) †ðð } ð=1
is the intersection of a ï¬nite number of closed convex sets. 3.217 Each row að = (ðð1 , ðð2 , . . . ððð ) of ðŽ deï¬nes a linear functional ðð (x) = ðð1 ð¥1 + ðð2 ð¥2 + â
â
â
+ ððð ð¥ð on âð . The set ð of solutions to ðŽx †c is ð = { ð¥ â ð : ðð (x) †ðð , ð = 1, 2, . . . , ð } is a polyhedron. 3.218 For simplicity, we assume that the game is superadditive, so that ð€(ð) ⥠0 for every ð. Consequently, in every core allocation x, 0 †ð¥ð †ð€(ð ) and core â [0, ð€(ð )] à [0, ð€(ð )] à â
â
â
à [0, ð€(ð )] â âð Thus, the core is bounded. Since it is the intersection of closed halfspaces, the core is also closed. By Proposition 1.1, the core is compact. 3.219 polytope =â polyhedron Assume that ð is a polytope generated by the points { x1 , x2 , . . . , xð } and let ð¹1 , ð¹2 , . . . , ð¹ð denote the proper faces of ð . For each ð = 1, 2, . . . , ð, let ð»ð denote the hyperplane containing ð¹ð so that ð¹ð = ð â© ð»ð . For every such hyperplane, there exists a nonzero linear functional ðð and constant ðð such that ðð (x) = ðð for every x â ð»ð . Furthermore, every such hyperplane is a bounding hyperplane of ð . Without loss of generality, we can assume that ðð (x) †ð for every x â ð . Let ð = { x â ð : ðð (x) †ðð , ð = 1, 2, . . . , ð } Clearly ð â ð. To show that ð â ð , assume not. That is, assume that there exists y â ð âð and let x â ri ð . (ri ð is nonempty by exercise 1.229). Since ð is ¯ = ðŒx+(1âðŒ)y belongs closed (Exercise 1.227), there exists a some ðŒ such that x ¯ â ð¹ð â ð»ð . to the relative boundary of ð , and there exists some ð such that x Let ð»ð+ = { x â ð : ðð (x) †ðð } denote the closed half-space bounded by ð»ð and ¯ = ðŒx+(1âðŒ)y, which implies that containing ð . ð»ð is a face of ð»ð+ containing x x, y â ð»ð . This in turn implies that x â ð¹ð , which contradicts the assumption that x â ri ð . We conclude that ð = ð . 179
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
polyhedron =â polytope Conversely, assume ð is a nonempty compact polyhedral set in a normed linear space. Then, there exist linear functionals ð1 , ð2 , . . . , ðð in ð â and numbers ð1 , ð2 , . . . , ðð such that ð = { x â ð : ðð (x) †ðð , ð = 1, 2, . . . , ð }. We show that ð has a ï¬nite number of extreme points. Let ð denote the dimension of ð. If ð = 1, ð is either a single point or closed line segment (since ð is compact), and therefore has a ï¬nite number of extreme points (that is, 1 or 2). Now assume that every compact polyhedral set of dimension ð â 1 has a ï¬nite number of extreme points. Let ð»ð , ð = ð = 1, 2, . . . , ð denote the hyperplanes associated with the linear functionals ðð deï¬ning ð (Exercise 3.49). Let x be an extreme point of ð. Then ð is a boundary point of ð (Exercise 1.220) and therefore belongs to some ð»ð . We claim that x is also an extreme point of the set ð â© ð»ð . To see this, assume otherwise. That is, assume that x is not an extreme point of ð â©ð»ð . Then, there exists x1 , x2 â ð â©ð»ð such that x = ðŒx1 + (1 â ðŒ)x2 . But then x1 , x2 â ð and x is not an extreme point of ð. Therefore, every extreme point of ð is an extreme point of some ð â© ð»ð , which is a compact polyhedral set of dimension ð â 1. By hypothesis, each ð â© ð»ð has a ï¬nite number of extreme points. Since there are only ð such hyperplanes ð»ð , ð has a ï¬nite number of extreme points. By the Krein-Milman theorem (Exercise 3.209), ð is the closed convex hull of its extreme points. Since there are only ï¬nite extreme points, ð is a polytope. 3.220
1. Let ð, ð â ð â so that ð (x) †0 and ð(x) †0 for every x â ð. For every ðŒ, ðœ ⥠0 ðŒð (x) + ðœð (x) †0 for every ð¥ â ð. This shows that ðŒð + ðœð â ð â . ð â is a convex cone. To show that ð â is closed, let ð be the limit of a sequence (ðð ) of functionals in ð â . Then, for every x â ð, ðð (x) †0 so that ð (x) = lim ðð (x) †0
2. Let x, y â ð ââ . Then, for every ð â ð â ð (x) †0 and ð (y) †0 and therefore ð (ðŒx + ðœy) = ðŒð (x) + ðœð (y) †0 for every ðŒ, ðœ ⥠0. There ðŒx + ðœy â ð ââ . ð ââ is a convex cone. To show that ð ââ is closed, let xð be a sequence of points in ð ââ converging to ð¥. For every ð = 1, 2, . . . ð (xð ) †0 for every ð â ð â By continuity ð (x) = lim ð (xð ) †0 for every ð â ð â Consequently x â ð ââ which is therefore closed. 180
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3. Let x â ð. Then ð (x) †0 for every ð â ð â so that x â ð ââ . 4. Exercise 1.79. 3.221 Let ð â ð2â . Then ð (x) †0 for every x â ð2 . A fortiori, since ð1 â ð2 , ð (x) †0 for every x â ð1 . Therefore ð â ð1â . 3.222 Exercise 3.220 showed that ð â ð ââ . To show the converse, let y â / ð. By Proposition 3.14, there exists some ð â ð â and ð such that ð (y) > ð ð (x) < ð
for every x â ð
Since ð is a cone, 0 â ð and ð (0) = 0 < ð. Since ðŒð = ð for every ðŒ > 0 then ð (x) < 0
for every x â ð
/ ð ââ . That is so that ð â ð â . ð (y) > 0, y â yâ / ð =â y â / ð ââ from which we conclude that ð ââ â ð. 3.223 Let ðŸ = cone {ð1 , ð2 , . . . , ðð } ð â = { ð â ðâ : ð = ðð ðð , ðð ⥠0 } ð=1
be the set of all nonnegative linear combinations of the linear functionals ðð . ðŸ is a closed convex cone. Suppose that ð â / cone {ð1 , ð2 , . . . , ðð }, that is assume that ð â / ðŸ. Then {ð } is a compact convex set disjoint from ðŸ. By Proposition 3.14, there exists a continuous linear functional ð and number ð such that sup ð(ð) < ð < ð(ð )
ðâðŸ
Since 0 â ðŸ, ð ⥠0 and so ð(ð ) > 0. Further, for every ð â ðº ð â ðð ðð ) ð(ð) = ð( ð=1
=
ð â
ðð ð(ðð ) < ð for every ðð ⥠0
ð=1
Since ðð can be made arbitrarily large, this last inequality implies that ð(ðð ) †0
ð = 1, 2, . . . , ð
By the Riesz representation theorem (Exercise 3.75), there exists x â ð ð(ðð ) = ðð (x) and ð(ð ) = ð (x) Since ð(ðð ) = ðð (x) †0 181
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
x â ð. By hypothesis ð (x) = ð(ð ) †0 contradicting the conclusion that ð(ð ) > 0. This contradiction establishes that ð â ðŸ, that is ð (ð¥) =
ð â
ðð ðð (ð¥),
ðð ⥠0
ð=1
3.224 Let a1 , a2 , . . . , að denote the rows of ðŽ and deï¬ne the linear functional ð, ð1 , ð2 , . . . , ðð by ð (x) = cx ðð (x) = að x ð = 1, 2, . . . , ð Assume cx †0 for every x satisfying ðŽx †0, that is ð (x) †0 for every x â ð where ð = { x â ð : ðð (x) †0, ð = 1, 2, . . . , ð } By Proposition 3.18, there exists y â âð + such that ð (x) =
ð â
ðŠð ðð (x)
ð=1
or c=
ð â
ðŠð að = ðŽð y
ð=1
Conversely, assume that c = ðŽð y =
ð â
ðŠð að
ð=1
Then ðŽx †0 =â að x †0 for every ð =â cx †0 3.225 Let ð = âð+ denote the positive orthant of âð . ð is a convex set (indeed cone) with a nonempty interior. By Corollary 3.2.1, there exists a hyperplane ð»p (ð) such that pð x †ð †py
for every x â ð, y â ð
Since 0 â ð p0 = 0 ⥠ð which implies that ð †0 and pð x †ð †0
for every ð¥ â ð
To show that p is nonnegative, let e1 , e2 , . . . , eð denote the standard basis for âð . Each eð belongs to ð so that peð = ðð ⥠0 182
for every ð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.226 Assume yâ is an eï¬cient production plan in ð and let ð = ð â ðŠ â . ð is convex. We claim that ð â©âð++ = â
. Otherwise, if there exists some z â ð â©âð++ , let yâ² = yâ +z â z â ð implies yâ² â ð while â z â âð++ implies yâ² > yâ contradicting the eï¬ciency of yâ . Therefore, ð is a convex set which contains no interior points of the nonnegative orthant âð+ . By Exercise 3.225, there exists a price system p such that pð x †0 for every x â ð Since ð = ð â ðŠ â , this implies p(y â yâ ) †0 for every y â ð or pyâ ⥠py for every y â ð ðŠ â maximizes the producerâs proï¬t at prices p. 3.227 Consider the set ð â = { x â âð : âx â ð }. ð â© int âðâ = â
=â ð â â© int âð+ = â
From the previous exercise, there exists a hyperplane with nonnegative normal p â© 0 such that pð x †0
for every x â ð â
pð x ⥠0
for every x â ð
Since p â© 0, this implies
3.228
1. Suppose x â â¿(xâ ). Then, there exists an allocation (x1 , x2 , . . . , xð ) such that x=
ð â
xð
ð=1
where xð â â¿(xâð ) for every ð = 1, 2, . . . , ð. Conversely, if (x1â , x2 , . . . , xð ) is an ð allocation with xð â â¿(xâð ) for every ð = 1, 2, . . . , ð, then x = ð=1 xð â â¿(xâ ). 2. For every agent ð, xâð â â¿(xâð ), which implies that xâ =
ð â
xâð â â¿(xâ )
ð=1
and therefore 0 â ð = â¿(xâ ) â xâ â= â
Since individual preferences are convex, â¿(xâð ) is convex for each ð and therefore â â â â ð = â¿(x ) â x = ð â¿(xð ) â xâ is convex (Exercise 1.164).
183
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
Assume to the contrary that ð â© int âðâ â= â
. That is, there exists some z â ð with z < 0. This implies that there exists some allocation (x1 , x2 , . . . , xð ) such that â xð â xâ < 0 z= ð
xâð
and xð â¿ for every ð â ð . Distribute z equally to all the consumers. That is, consider the allocation yð = xð + z/ð By strict monotonicity, yð â» xð â¿ xâð for every ð â ð . Since â â â yð = xð + z = xâ = xâð ð
ð
ð
(y1 , y2 , . . . , yð ) is a reallocation of the original allocation xâ which is strictly preferred by all consumers. This contradicts the assumed Pareto eï¬ciency of xâ . We conclude that ð â© int âðâ â= â
3. Applying Exercise 3.227, there exists a hyperplane with nonnegative normal pâ â© 0 such that pâ z ⥠0 for every z â ð That is pâ (x â xâ ) ⥠0 or pâ x ⥠pâ xâ for every x â â¿(xâ )
(3.65)
4. Consider any allocation which is strictly preferred to xâ by consumer ð, that is xð â â»ð (xâð ). Construct another allocation y by taking ð > 0 of each commodity away from agent ð and distributing amongst the other agents to give yð = (1 â ð)xð ð yð = xâð + xð , ðâ1
ð â= ð
By continuity, there exists some ð > 0 such that yð = (1 â ð)xð â»ð xâð . By monotonicity, yð â»ð xâð for every ð â= ð. We have constructed â an allocation y which is strictly preferred to xâ by all the agents, so that y = ð yð â â¿(xâ ). (3.65) implies that py ⥠pxâ That is â
â( xâð + p â(1 â ð)xð + ðâ=ð
â â â â â ) â â ð xð â = p âxð + xâð â ⥠p âxâð + xâð â ðâ1 ðâ=ð
ðâ=ð
which implies that pxð ⥠pxâð
for every xð â â»(xâð ) 184
(3.66)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5. Trivially, xâ is a feasible allocation with endowments wð = xâð and ðð = pâ wð = pâ xâð . To show that (pâ , xâ ) is a competitive equilibrium, we have to show that xâð is the best allocation in the budget set ðð (p, ðð ) for each consumer ð. Suppose to the contrary there exists some consumer ð and allocation yð such that yð â» xð and pyð †ðð = pxâð . By continuity, there exists some ðŒ â (0, 1) such that ðŒyð â»ð xâð and p(ðŒyð ) = ðŒpyð < pyð †pxâ contradicting (3.66). We conclude that xâð â¿ð xð for every x â ð(pâ , ðð ) for every consumer ð. (pâ , xâ ) is a competitive equilibrium. 3.229 By the previous exercise, there exists a price system pâ such that xâð is optimal for each consumer ð in the budget set ð(pâ , pâ xâð ), that is xâð â¿ð xð for every xð â ð(pâ , pâ xâð )
(3.67)
For each consumer, let ð¡ð be the diï¬erence between her endowed wealth pâ wð and her required wealth pâ xâð . That is, deï¬ne ð¡ð = pâ xâð â pâ wð = pâ (xâð â wð ) Then pâ xâð = pâ + wð
(3.68)
â
By assumption x is feasible, so that â â â xâð â wð = (xâð â wð ) = 0 ð
so that
ð
â ð
ð¡ð = pâ
ð
â
(xâð â wð ) = 0
ð
â
Furthermore, for ðð = ð wð + ð¡ð , (3.68) implies ð(pâ , ðð ) = { xð : pâ xð †pâ wð + ð¡ð } = { xð : pâ xð †pâ xâð } = ð(pâ , pâ xâð ) for each consumer ð. Using (3.67) we conclude that xâð â¿ð xð for every xð â ð(pâ , ðð ) for every agent ð. (pâ , xâ ) is a competitive equilibrium where each consumerâs after-tax wealth is ðð = pwð + ð¡ð 3.230 Apply Exercise 3.202 with ðŸ = âð+ . 3.231 ðŸ â = { p : pð x †0 for every x â ðŸ } No such hyperplane exists if and only if ðŸ â â© âð++ = â
. Assume this is the case. By Exercise 3.225, there exists x â© 0 such that xp = pð x †0 for every p â ðŸ â In other words, x â ðŸ ââ . By the duality theorem ðŸ ââ = ðŸ which implies that x â ðŸ as well as âð+ , contrary to the hypothesis that ðŸ â© âð+ = {0}. This contradiction establishes that ðŸ â â© âð++ â= â
. 185
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.232 Given a set of ï¬nancial assets with prices p and payoï¬ matrix ð
, let ð = { (âpx, ð
ð¥) : x â âð } ð is the set of all possible (cost, payoï¬) pairs. It is a subspace of âð+1 . Let ð be the nonnegative orthant in âð+1 . The no arbitrage condition ð
x ⥠0 =â pð x ⥠0 implies that ð â© ð = {0}. By Exercise 3.230, there exists a hyperplane with positive normal ð = ð0 , ð1 , . . . , ðð such that ðz = 0
for every z â ð
ðz > 0
â {0} for every z â âð+1 +
That is for every x â âð
âð0 px + ðð
x = 0 or pð x = ð/ð0 ð
x
for every x â âð
ð/ð0 is required state price vector. Conversely, if a state price vector exists ðð =
ð â
ð
ðð ðð
ð=1
then clearly ð
x ⥠0 =â pð x ⥠0 No arbitrage portfolios exist. 3.233 Apply the Farkas lemma to the system âðŽx †0 âcð x > 0 3.234 The inequality system ðŽð y ⥠c has a nonnegative solution if and only if the corresponding system of equations ðŽð y â z = c ð has a nonnegative solution y â âð + , z â â+ . This is equivalent to the system ( ) y â² ðµ =c z
(3.69)
where ðµ â² = (ðŽð , âðŒð ) and ðŒð is the ð à ð identity matrix. By the Farkas lemma, system (3.69) has no solution if and only if the system ðµx †0 and cð x > 0 ( ) ðŽ has a solution x â âð . Since ðµ = , ðµx †0 implies âðŒ ðŽx †0 and â ðŒx †0 and the latter inequality implies x â âð+ . Thus we have established that the system ðŽð y ⥠c has no nonnegative solution if and only if ðŽx †0 and cð x > 0 for some x â âð+ 186
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Ë â âð+ such that 3.235 Assume system I has a solution, that is there exists x Ëâ¥0 ðŽË x = 0, cË x > 0, x Ë /cË Then x = x x satisï¬es the system ðŽx = 0, cx = 1, x ⥠0
(3.70)
xâ² ðŽð = 0, xc = 1, x ⥠0
(3.71)
which is equivalent to
Suppose y â âð satisï¬es ðŽy ⥠c Multiplying by x ⥠0 gives xâ² ðŽð y ⥠xc Substituting (3.71), this implies the contradiction 0â¥1 We conclude that system II cannot have a solution if I has a solution. Now, assume system I has no solution. System I is equivalent to (3.70) which in turn is equivalent to the system ( ) ( ) ðŽ 0 x= c 1 or ðµx = b (3.72) ) ( ) ( âðŽ 0 is (ð + 1) à ð and b = â âð+1 . If (3.72) has no solution, where ðµ = c 1 there exists (by the Farkas alternative) some z â âð+1 such that ðµ â² z †0 and bz > 0 Decompose z into z = (y, ð§) with y â âð and ð§ â â. The second inequality implies that (0, 1)â² (y, ð§) = 0y + ð§ = ð§ > 0 Without loss of generality, we can normalize so that ð§ = 1 and z = (y, 1). Now ðµ â² = (âðŽð , c) and so the ï¬rst inequality implies that ( ) y ð (âðŽ , c) = âðŽð y + c †0 1 or ðŽð y ⥠c We conclude that II has a solution. 187
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.236 For every linear functional ðð , there exists a vector a â âð such that ðð (x) = að xË (Proposition 3.11). Let ðŽð be the matrix whose rows are að , that is â 1â a â a2 â â ðŽ=â â. . .â að Then, the system of inequalities (3.31) is ðŽð x ⥠c where c = (ð1 , ð2 , . . . , ðð ). By the preceding exercise, this system is consistent if and only there is no solution to the system ðŽð = 0
cð > 0
ðâ¥0
Now ð â
ðŽð = 0 ââ
ðð ðð = 0
ð = 1, 2, . . . , ð
ð=1
Therefore, the inequalities (3.31) is consistent if an only if ð â
ðð ðð = 0 =â
ð=1
ð â
ðð ðð †0
ð=1
for every set of nonnegative numbers ð1 , ð2 , . . . , ðð . 3.237 Let ðµ be the 2ð à ð matrix comprising ðŽ and âðŽ as follows ( ) ðŽ ðµ= âðŽ Then the Fredholm alternative I cð x = 1
ðŽx = 0 is equivalent to the system ðµx †0
cx > 0
(3.73)
2ð By the Farkas alternative theorem, either (3.73) has a solution or there exists ð â ð
+ such that
ðµâ²ð = c
(3.74)
Decompose ð into two ð-vectors ð ð = (ð, ð¿), ð, ð¿ â ð
+
so that (3.74) can be rewritten as ðµ â² ð = ðŽð ð â ðŽð ð¿ = ðŽð (ð â ð¿) = c Deï¬ne y = ð â ð¿ â âð We have established that either (3.73) has a solution or there exists a vector y â âð such that ðŽð y = c 188
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.238 Let að , ð = 1, 2, . . . , ð denote the rows of ðŽ. Each að deï¬nes linear functional ðð (ð¥) = að ð¥ on âð , and c deï¬nes another linear functional ð (ð¥) = cð x. Assume that ð (ð¥) = cð x = 0 for every x â ð where ð = { x : ðð (x) = að x = 0, ð = 1, 2, . . . , ð } Then the system ðŽð¥ = 0 has no solution satisfying the constraint cð x > 0. By Exercise 3.20, there exists scalars ðŠ1 , ðŠ2 , . . . , ðŠð such that ð (x)=
ð â
ðŠð ðð (x)
ð=1
or c=
ð â
ðŠð ðð = ðŽð y
ð=1
That is y = (ðŠ1 , ðŠ2 , . . . , ðŠð ) solves the related nonhomogeneous system ðŽð y = c Conversely, assume that ðŽð y = c for some ðŠ â âð . Then cð x = ðŠðŽð¥ = 0 for all ð¥ such that ðŽð¥ = 0 and therefore there is no solution satisfying the constraint cð x = 1. 3.239 Let ð = { z : z = ðŽx, x â â } the image of ð. ð is a subspace. Assume that system I has no solution, that is ð â© âð ++ = â
By Exercise 3.225, there exists y â âð + â {0} such that yz = 0 for every z â ð That is yðŽx = 0 for every x â âð Letting x = ðŽð y, we have yðŽðŽð y = 0 which implies that ðŽð y = 0 System II has a solution y. Ë is a solution to I. Suppose to the contrary there also exists Conversely, assume that x Ë to II. Then, since ðŽË Ë â© 0, we must have y Ë ðŽË Ë ðŽð y Ë > 0. a solution y x > 0 and y x=x ðË ð ËðŽ y Ë = 0, a contradiction. Hence, we On the other hand, ðŽ y = 0 which implies x conclude that II cannot have a solution if I has a solution. 189
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.240 We have already shown (Exercise 3.239) that the alternatives I and II are mutually incompatible. If Gordanâs system II ðŽð y = 0 has a semipositive solution y â© 0, then we can normalize y such that 1y = 1 and the system ðŽð y = 0 1y = 1 has a nonnegative solution. Conversely, if Gordanâs system II has no solution, the system ðµâ²y = c (
) ðŽð where ðµ = and c = (0, 1) = (0, 0, . . . , 0, 1), 0 â âð , is the (ð + 1)st unit 1 vector has no solution y ⥠0. By the Farkas lemma, there exists z â âð+1 such that â²
ðµz ⥠0 cz < 0 Decompose z into z = (x, ð¥) with x â âð . The second inequality implies that ð¥ < 0 since cz = (0, 1)â² (x, ð¥) = ð¥ < 0 Since ðµ = (ðŽ, 1), the ï¬rst inequality implies that ðµz = (ðŽ, 1)(x, ð¥) = ðŽx + 1ð¥ ⥠0 or ðŽx ⥠â1ð¥ > 0 x solves Gordanâs system I. 3.241 Let a1 , a2 , . . . , að be a basis for ð. Let ðŽ = (a1 , a2 , . . . , að ) be the matrix whose columns are að . To say that ð contains no positive vector means that the system ðŽx > 0 has no solution. By Gordanâs theorem, there exists some y â© 0 such that ðŽð y = 0 that is að y = yað = 0, ð = 1, 2, . . . , ð so that y â ð ⥠. 190
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.242 Let ð be the subspace ð = { z : ðŽx : x â âð }. System I has no solution ðŽx â© 0 if and only if ð has no nonnegative vector z â© 0. By the previous exercise, ð ⥠contains a positive vector y > 0 such that yz = 0 for every z â ð Letting x = ðŽð y, we have yðŽðŽð y = 0 which implies that ðŽð y = 0 System II has a solution y. 3.243 Let ð = { z : z = ðŽx, x â â } the image of ð. ð is a subspace. Assume that system I has no solution, that is ð â© âð + = {0} By Exercise 3.230, there exists y â âð ++ such that yz = 0 for every z â ð That is yðŽx = 0 for every x â âð Letting x = ðŽð y, we have yðŽðŽð y = 0 which implies that ðŽð y = 0 System II has a solution y. Ë is a solution to I. Suppose to the contrary there also exists Conversely, assume that x Ë > 0. Ë to II. Then, since ðŽË Ë > 0, we must have y Ë ðŽË Ë ðŽð y a solution y x â© 0 and y x=x ðË ð ËðŽ y Ë = 0, a contradiction. Hence, we On the other hand, ðŽ y = 0 which implies x conclude that II cannot have a solution if I has a solution. 3.244 The inequality system ðŽð y †0 has a nonnegative solution if and only if the corresponding system of equations ðŽð y + z = 0 ð has a nonnegative solution y â âð + , z â â+ . This is equivalent to the system ( ) y â² ðµ =0 z
(3.75)
where ðµ â² = (ðŽð , ðŒð ) and ðŒð is the ð à ð identity matrix. By Gordanâs theorem, system (3.75) has no solution if and only if the system ( has a solution x â âð . Since ðµ =
ðµx > 0 ) ðŽ , ðµx > 0 implies ðŒ
ðŽx > 0 and ðŒx > 0 and the latter inequality implies x â âð++ . Thus we have established that the system ðŽð y †0 has no nonnegative solution if and only if ðŽx > 0 for some x â âð++ 191
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.245 Assume system II has no solution, that is there is no y â âð such that ðŽy †0, y â© 0 This implies that the system âðŽy ⥠0 1y ⥠1 ( ) âðŽ ð â² has no solution y â â+ . Deï¬ning ðµ = , the latter can be written as 1â² ðµ â² y ⥠âeð+1
(3.76)
where âeð+1 = (0, 1), 0 â âð . By the Gale alternative (Exercise 3.234), if system (3.76) has no solution, the alternative system ðµz †0, âeð+1 z > 0 ð has a nonnegative solution z â âð+1 + . Decompose z into z = (x, ð§) where x â â+ and ð§ â â+ . The second inequality implies ð§ > 0 since eð+1 z = ð§.
ðµ = (âðŽð , 1) and the ï¬rst inequality implies ( ) x ðµz = (âðŽð , 1) = âðŽð x + 1𧠆0 ð§ or ðŽð x ⥠1ð§ > 0 Thus system I has a solution x â âð+ . Since x = 0 implies ðŽx = 0, we conclude that x â© 0. Conversely, assume that II has a solution y â© 0 such that ðŽy †0. Then, for every x â âð+ xðŽð y = yâ² ðŽð x †0 Since y â© 0, this implies ðŽð x †0 for every x â âð+ which contradicts I. 3.246 We give a constructive proof, by proposing an algorithm which will generate the desired decomposition. Assume that x satisï¬es ðŽx â© 0. Arrange the rows of ðŽ such that the positive elements of ðŽx are listed ï¬rst. That is, decompose ðŽ into two submatrices such that ðµ1x > 0 ð¶1x = 0 Either Case 1 ð¶ 1 x â© 0 has no solution and the result is proved or 192
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Case 2 ð¶ 1 x â© 0 has a solution xâ² . ¯ be a linear combination of x and xâ² . Speciï¬cally, deï¬ne Let x ¯ = ðŒx + xâ² x where ðŒ > max
âbð x bð xâ²
where bð is the ðth row of ðµ 1 . ðŒ is chosen so that ðŒðµ 1 x > ðµ 1 xâ² By direct computation ¯ = ðŒðµ 1 x + ðµ 1 xâ² > 0 ðµ1x ¯ = ðŒð¶ 1 x + ð¶ 1 xâ² â© 0 ð¶1x ¯ is another solution to ðŽx â© 0 since ð¶ 1 x = 0 and ð¶ 1 xâ² â© 0. By construction, x such that ðŽÂ¯ x has more positive components than ðŽx. Again, collect all the positive components together, decomposing ðŽ into two submatrices such that ¯>0 ðµ2x ¯=0 ð¶2x Either Case 1 ð¶ 2 x â© 0 has no solution and the result is proved or Case 2 ð¶ 2 x â© 0 has a solution xâ²â² . In the second case, we can repeat the previous procedure, generating another decomposition ðµ 3 , ð¶ 3 and so on. At each stage ð, the matrix ðµ ð get larger and ð¶ ð smaller. The algorithm must terminate before ðµ ð equals ðŽ, since we began with the assumption that ðŽx > 0 has no solution. 3.247 There are three possible cases to consider. Case 1: y = 0 is the only solution of ðŽð y = 0. Then ðŽx > 0 has a solution xâ² by Gordanâs theorem and ðŽxâ² + 0 > 0 Case 2: ðŽð y = 0 has a positive solution y > 0 Then 0 is the only solution ðŽx ⥠0 by Stiemkeâs theorem and ðŽ0 + y > 0 Case 3 ðŽð y = 0 has a solution y â© 0 but y â> 0. By Gordanâs theorem ðŽx > 0 has no solution. By the previous exercise, ðŽ can be decomposed into two consistent subsystems ðµx > 0 ð¶x = 0 193
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
such that ð¶x â© 0 has no solution. Assume that ðµ is ð à ð and ð¶ is ð à ð where ð = ð â ð. Applying Stiemkeâs theorem to ð¶, there exists z > 0, z â âð . Deï¬ne y â âð + by { 0 ð = 1, 2, . . . , ð ðŠð = ðŠð = ð§ðâð ð = ð + 1, ð + 2, . . . , ð Then x, y is the desired solution since for every ð, ð = 1, 2, . . . , ð either ðŠð > 0 or (ðŽx)ð = (ðµx)ð > 0. 3.248 Consider the dual pair ( ) ( ) y ðŽ = 0, y ⥠0, z ⥠0 x ⥠0 and (ðŽð , ðŒ) z ðŒ By Tuckerâs theorem, this has a solution xâ , yâ , zâ such that ðŽxâ ⥠0, xâ ⥠0, ðŽð yâ + zâ = 0, yâ ⥠0, zâ ⥠0 ðŽx + y > 0 ðŒxâ + ðŒz > 0 Substituting zâ = âðŽð yâ implies ðŽð y †0 and x â ðŽð yâ > 0 3.249 Consider the dual pair ðŽx ⥠0 and ðŽð y = 0, y ⥠0 where ðŽ is an ð à ð matrix. By Tuckerâs theorem, there exists a pair of solutions xâ â âð and yâ â âð such that ðŽxâ + yâ > 0
(3.77)
Assume that ðŽx > 0 has no solution (Gordan I). Then there exists some ð such that (ðŽxâ )ð = 0 and (3.77) implies that ðŠðâ > 0. Therefore yâ â© 0 and solves Gordan II. Conversely, assume that ðŽð y = 0 has no solution y > 0 (Stiemke II). Then, there exists some ð such that ðŠðâ = 0 and (3.77) implies that (ðŽxâ )ð > 0). Therefore xâ solves ðŽx â© 0 (Stiemke I). 3.250 We have already shown that Farkas I and II are mutually inconsistent. Assume that Farkas system I ðŽx ⥠0, cð x < 0 ( has no solution. Deï¬ne the (ð + 1) à ð matrix ðµ = the system ðµx ⥠0 194
ðŽ âcâ²
) . Our assumption is that
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
has no solution with (ðµx)ð+1 = âcx > 0. By Tuckerâs theorem, the dual system ðµâ²z = 0 has a solution z â âð+1 with zð+1 > 0. Without loss of generality, we can normalize + â² ð so that zð+1 = 1. Decompose z into z = (y, 1) with y â âð + . Since ðµ = (ðŽ , âc), â² ðµ z = 0 implies ðµ â² z = (ðŽð , âc)(y, 1) = ðŽð y â c = 0 or ðŽð y = c y â âð + solves Farkas II. 3.251 If x ⥠0 solves I, then xâ² (ðŽð y1 + ðµ â² y2 + ð¶ â² y3 ) = xâ² ðŽð y1 + xâ² ðµ â² y2 + xâ² ð¶ â² y3 ) > 0 since xâ² ðŽð y1 = y1 ðŽx > 0, xâ² ðµ â² y2 = y2 ðµx ⥠0 and xâ² ð¶ â² y3 = y3 ð¶x = 0 which contradicts II. The equation ð¶x = 0 is equivalent to the pair of inequalities ð¶x ⥠0, âð¶x ⥠0. By Tuckerâs theorem the dual pair ðŽð y1 + ðµ â² y2 + ð¶ â² y3 â ð¶ â² y4 = 0
ðŽx ⥠0 ðµx ⥠0 ð¶x ⥠0 âð¶x ⥠0
has solutions ð¥ â âð , y1 â âð1 , y2 â âð2 , u3 , v3 â âð3 such that y1 ⥠0
ðŽx + y1 > 0
y2 ⥠0 u3 ⥠0
ðµx + y2 > 0 ð¶x + u3 > 0
v3 ⥠0
âð¶x + v3 > 0
Assume Motzkin I has no solution. That is, there is y1 â© 0. Deï¬ne y3 = u3 â v3 . Then y1 , y2 , y3 satisï¬es Motzkin II. 3.252
1. For every a â ð, let ðaâ be the polar set ðaâ = { x â âð : â¥x⥠= 1, xa ⥠0 } ðaâ is nonempty since 0 â ðaâ . Let x be the limit of a sequence xð of points in ðaâ . Since xð a ⥠0 for every ð, xa ⥠0 so that x â ðaâ . Hence ðaâ is a closed subset of ðµ = { x â âð : â¥x⥠= 1 }.
2. Let {a1 , a2 , . . . , að } be any ï¬nite set of points in ð. Since 0 â / ð, the system ð â
ðŠð að = 0,
ð=1
ð â
ðŠð = 1, ðŠð ⥠0
ð=1
has no solution. A fortiori, the system ð â
ðŠð að = 0
ð=1
195
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
has no solution ðŠ â âð + . If ðŽ is the ðÃn matrix whose rows are að , the latter system can be written as ðŽð y = 0 3. By Gordanâs theorem, the system ðŽx > 0
(3.78)
¯ â= 0. has a solution x 4. Without loss of generality, we can take â¥Â¯ x⥠= 1. (3.78) implies that ¯=x ¯ að > 0 að x ¯ â ðaâð . Hence for every ð = 1, 2 . . . , ð so that x ð â©
¯â x
ð=1
ðaâð
â©ð 5. We have shown that for every ï¬nite set {a1 , a2 , . . . , að } â ð, ð=1 ðaâð is nonempty closed subset of the compact set ðµ = {ð¥ â âð : â¥x⥠= 1}. By the Finite intersection property (Exercise 1.116) â© ðaâ â= â
aâð
6. For every p â
â© aâð
ðaâ pa ⥠0 for every a â ð
p deï¬nes a hyperplane ð (a) = pa which separates ð from 0. 3.253 The expected outcome if player 1 adopts the mixed strategy p = (ð1 , ð2 , . . . , ðð ) and player 2 plays her ð pure strategy is ð¢(p, ð) =
ð â
ðð ððð = pað
ð=1
where að is the ðth column of ðŽ. The expected payoï¬ to 1 for all possible responses of player 2 is the vector (pðŽ)â² = ðŽð p. The mixed strategy p ensures player 1 a nonnegative security level provided ðŽð p ⥠0. Similarly, if 2 adopts the mixed strategy q = (ð1 , ð2 , . . . , ðð ), the expected payoï¬ to 2 if 1 plays his ð strategy is að q where að is the ðth row of ðŽ. The expected outcome for all the possible responses of player 1 is the vector ðŽq. The mixed strategy q ensures player 2 a nonpositive security level provided ðŽq †0. By the von Neumann alternative theorem (Exercise 3.245), at least one of these alternatives must be true. That is, either Either I ðŽð p > 0, p â© 0 for some p â âð or II ðŽq †0, q â© 0 for some q â âð Since p â© 0 and q â© 0, we can normalize so that p â Îðâ1 and q â Îðâ1 . At least one of the players has a strategy which guarantees she cannot lose. 196
Solutions for Foundations of Mathematical Economics 3.254
c 2001 Michael Carter â All rights reserved
1. For any ð â â, deï¬ne the game ð¢Ë(a1 , a2 ) = ð¢(a1 , a2 ) â ð with Ë(p, ð) = max min ð¢(p, ð) â ð = ð£1 â ð ð£Ë1 = max min ð¢ p
p
ð
ð
ð£Ë2 = min max ð¢ Ë(ð, q) = min max ð¢(ð, q) â ð = ð£2 â ð q
q
ð
ð
By the previous exercise, Either ð£Ë1 ⥠0 or ð£ËðŠ †0 That is Either ð£1 ⥠ð or ð£2 †ð 2. Since this applies for arbitrary ð â â, it implies that while ð£1 †ð£2 and there is no ð such that ð£1 < ð < ð£2 Therefore, we conclude that ð£1 = ð£2 as required. 3.255
1. The mixed strategies p of player 1 are elements of the simplex Îðâ1 , which is compact (Example 1.110). Since ð£1 (p) = minðð=1 ð¢(p, ð) is continuous (Maximum theorem 2.3), ð£1 (p) achieves its maximum on Îðâ1 (Weierstrass theorem 2.2). That is, there exists pâ â Îðâ1 such that ð£1 = ð£1 (pâ ) = max ð£1 (p) p
Similarly, there exists qâ â Îðâ1 such that ð£2 = ð£2 (qâ ) = min ð£2 (q) q
2. Let ð¢(p, q) denote the expected outcome when player 1 adopts mixed strategy p and player 2 plays q. That is ð¢(p, q) =
ð â ð â
ðð ðð ððð
ð=1 ð=1
Then ð£ = ð¢(pâ , qâ ) = max ð¢(ð, qâ ) ⥠ð
â
ðð ð¢(ð, qâ ) = ð¢(p, qâ ) for every p â Îðâ1
ð
Similarly ð£ = ð¢(pâ , qâ ) = min ð¢(pâ , ð) †ð
â
ðð ð¢(pâ , ð) = ð¢(pâ , q) for every q â Îðâ1
ð
(pâ , qâ ) is a Nash equilibrium. 197
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
3.256 By the Minimax theorem, every ï¬nite two person zero-sum game has a value. The previous result shows that this is attained at a Nash equilibrium. 3.257 If player 2 adopts the strategy ð¡1 ðp (ð¡1 ) = âð1 + 2ð2 < 0 if ð1 > 2ð2 If player 2 adopts the strategy ð¡5 ðp (ð¡5 ) = ð1 â 2ð2 < 0 if ð1 < 2ð2 Therefore ð£1 (p) = min ðp (z) †min{ðp (ð¡1 ), ðp (ð¡5 )} < 0 ð§âð
for every p such that ð1 â= ð2 . Since ð1 + ð2 = 1, we conclude that { = 0 p = pâ = ( 2/3, 1/3) ð£1 (p) < 0 otherwise We conclude that ð£1 = max ð£1 (p) = 0 p
which is attained at pâ = ( 2/3, 1/3). 3.258
1. ð
ð£2 = min max ð§ð zâð ð=1
Since ð is compact, ð£2 = 0 implies there exists z¯ â ð such that ð
max ð§Â¯ð = 0 ð=1
which implies that ¯ z †0. Consequently ð â© âðâ â= â
. 2. Assume to the contrary that there exists z â ð â© int âðâ That is, there exists some strategy q â Îðâ1 such that ðŽq < 0 and therefore ð£2 < 0, contrary to the hypothesis. 3. There exists a hyperplane with nonnegative normal separating ð from âðâ (Exercise 3.227). That is, there exists pâ â âð+ , pâ â= 0 such that ðpâ (z) ⥠0 for every z â ð and therefore ð£1 (pâ ) = min ðpâ (z) ⥠0 zâð
Without loss of generality, we can normalize so that pâ â Îðâ1 .
198
âð
ð=1
ðâð = 1 and therefore
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4. Consequently ð£1 = max ð£1 (p) ⥠ð£1 (pâ ) ⥠0 p
On the other hand, we know that ð contains a point z¯ †0. For every p ⥠0 ðp (¯z) †0 and therefore ð§) †0 ð£1 (p) = min ðp (z) †ðp (¯ zâð
so that ð£1 = max ð£1 (p) †0 p
We conclude that ð£1 = 0 = ð£2 3.259 Consider the game with the same strategies and the payoï¬ function ð¢ Ë(a1 , a2 ) = ð¢(a1 , a2 ) â ð The expected value to player 2 is Ë(ð, q) = min max ð¢(ð, q) â ð = ð£2 â ð = 0 ð£Ë2 = min max ð¢ q
q
ð
ð
By the previous exercise ð£Ë1 = ð£Ë2 = 0 and ð£1 = max min ð¢(p, ð) = max min ð¢ Ë(p, ð) + ð = ð£Ë1 + ð = ð = ð£2 p
q
ð
ð
3.260 Assume that p1 and p2 are both optimal strategies for player 1. Then ð¢(p1 , q) ⥠ð£ for every q â Îðâ1 ð¢(p2 , q) ⥠ð£ for every q â Îðâ1 ¯ = ðŒp1 , p2 + (1 â ðŒ). Since ð¢ is bilinear Let p ð¢(¯ p, q) = ðŒð¢(p1 , q) + (1 â ðŒ)ð¢(p2 , q) ⥠ð£ for every q â Îðâ1 ¯ is also an optimal strategy for player 1. Consequently, p 3.261 ð is the payoï¬ function of some 2 person zero-sum game in which the players have ð + 1 and ð + 1 strategies respectively. The result follows from the Minimax Theorem. 3.262
1. The possible partitions of ð = {1, 2, 3} are: {1}, {2}, {3} {ð, ð}, {ð},
ð, ð, ð =â ð, ð â= ð â= ð
{1, 2, 3} In any partition, at most one coalition can have two or more players, and therefore ðŸ â
ð€(ðð ) †1
ð=1
199
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
2. Assume x = (ð¥1 , ð¥2 , ð¥3 ) â core. Then x must satisfy the following system of inequalities ð¥1 + ð¥2 ⥠1 = ð€({1, 2}) ð¥1 + ð¥3 ⥠1 = ð€({1, 3}) ð¥2 + ð¥3 ⥠1 = ð€({2, 3}) which can be summed to yield 2(ð¥1 + ð¥2 + ð¥3 ) ⥠3 or ð¥1 + ð¥2 + ð¥3 ⥠3/2 which implies that x exceeds the sum available. This contradiction establishes that the core is empty. Alternatively, observe that the three person majority game is a simple game with no veto players. By Exercise 1.69, its core is empty. 3.263 Assume that the game (ð, ð€) is not cohesive. Then there exists a partition {ð1 , ð2 , . . . , ððŸ } of ð such that ð€(ð ) <
ðŸ â
ð€(ðð )
ð=1
Assume x â core. Then
â
ð¥ð ⥠ð€(ðð )
ð = 1, 2, . . . , ðŸ
ðâðð
Since {ð1 , ð2 , . . . , ððŸ } is a partition â
ð¥ð =
ðâð
ðŸ â â
ð¥ð â¥
ð=1 ðâðð
ð â
ð€(ðð ) > ð€(ð )
ð=1
which contradicts the assumption that x â core. This establishes that cohesivity is necessary for the existence of the core. To show that cohesivity is not suï¬cient, we observe that the three person majority game is cohesive, but its core is empty. 3.264 The other balanced families of coalitions in a three player game are 1. ⬠= {ð } with weights { ð€(ð) =
1 0
ð=ð otherwise
2. ⬠= {{1}, {2}, {3}} with weights ð€(ð) = 1 for every ð â ⬠3. ⬠= {{ð}, {ð, ð}}, ð, ð, ð â â¬, ð â= ð â= ð with weights ð€(ð) = 1 for every ð â ⬠3.265 The following table lists some nontrivial balanced families of coalitions for a four player game. Other balanced families can be obtained by permutation of the players. 200
Solutions for Foundations of Mathematical Economics
{123}, {124}, {34} {12}, {13}, {23}, {4} {123}, {14}, {24}, {3} {123}, {14}, {24}, {34} {123}, {124}, {134}, {234}
c 2001 Michael Carter â All rights reserved
Weights 1/2, 1/2, 1/2 1/2, 1/2, 1/2, 1 1/2, 1/2, 1/2, 1/2 2/3, 1/3, 1/3, 1/3 1/3, 1/3, 1/3, 1/3
3.266 Both sides of the expression eð =
â
ðð eð
ðââ¬
are vectors, with each component corresponding to a particular player. For player ð, the ðð¡ â component of eð is 1 and the ðð¡ â component of eð is 1 if ð â ð and 0 otherwise. Therefore, for each player ð, the preceding expression can be written â ðð = 1 ðââ¬â£ðâð
For each coalition ð, the share of the coalition ð at the allocation x is â ðð (x) = ð â ðð¥ð = eð xË The condition ðð =
â
(3.79)
ðð ðð
ðââ¬
means that for every x â ð ðð (x) =
â
ðð ðð (x)
ðââ¬
Substituting (3.79) eð xË =
â
ðð ðð xË
ðââ¬
which is equivalent to the condition â
ðð eð = eð
ðââ¬
3.267 By construction, ð ⥠0. If ð = 0, â ðð ðð â ððð = 0 ðâð
implies that ðð = 0 for all ð and consequently â ðð ð€(ð) â ðð€(ð ) †0 ðâð
is trivially satisï¬ed. On the other hand, if ð > 0, we can divide both conditions by ð.)
201
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3.268 Let (ð, ð€1 ) and (ð, ð€2 ) be balanced games. By the Bondareva-Shapley theorem, they have nonempty cores. Let x1 â core(ð, ð€1 ) and x2 â core(ð, ð€2 ). That is, ðð (x1 ) ⥠ð€1 (ð) for every ð â ð ðð (x2 ) ⥠ð€2 (ð) for every ð â ð Adding, we have ðð (x1 ) + ðð (x2 ) = ðð (x1 + x2 ) ⥠ð€1 (ð) + ð€2 (ð) for every ð â ð which implies that x1 + x2 belongs to core(ð, ð€1 + ð€2 ). Therefore (ð, ð€1 + ð€2 ) is balanced. Similarly, if x â core(ð, ð€), then ðŒx belongs to core(ð, ðŒð€) for every ðŒ â â+ . That is (ð, ðŒð€) is balanced for every ðŒ â â+ . 3.269
1. Assume otherwise. That is assume there exists some y â ðŽ â© ðµ. Taking the ï¬rst ð components, this implies that â eð = ðð eð ðâð
for some (ðð ⥠0 : ð â ð ). Let ⬠= {ð â ð ⣠ðð > 0} be the set of coalitions with strictly positive weights. Then ⬠is a balanced family of coalitions with weights ðð (Exercise 3.266). However, looking at the last coordinate, y â ðŽ â© ðµ implies â ðð ð€(ð) = ð€(ð ) + ð > ð€(ð ) ðââ¬
which contradicts the assumption that the game is balanced. We conclude that ðŽ and ðµ are disjoint if the game is balanced. 2. (a) Substituting y = (eâ
, 0) in (3.36) gives (z, ð§0 )â² (0, 0) = 0 ⥠ð which implies that ð †0. NOTE We still have to show that ð ⥠0. (b) Substituting (eð , ð€ðŠ(ð )) in (3.36) gives ð§eð + ð§0 ð€(ð ) > ð§eð + ð§0 ð€(ð ) + ð§0 ð for all ð > 0, which implies that ð§0 < 0. 3. Without loss of generality, we can normalize so that ð§0 = â1. Then the separating hyperplane conditions become (z, â1)â² y ⥠0 â²
(z, â1) (eð , ð€(ð ) + ð) < 0
for every y â ðŽ
(3.80)
for every ð > 0
(3.81)
For any ð â ð , (eð , ð€(ð)) â ðŽ. Substituting y = (eð , ð€(ð)) in (3.80) gives eâ²ð z â ð€(ð) ⥠0 that is ðð (z) = eâ²ð z =⥠ð€(ð) 202
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics while (3.81) implies ðð (z) = eâ²ð z > ð€(ð ) + ð
for every ð > 0
This establishes that z belongs to the core. Hence the core is nonempty. â 3.270 1. Let ðŒ = ð€(ð ) â ðâð ð€ð > 0 since (ð, ð€) is essential. For every ð â ð , deï¬ne ( ) â 1 0 ð€(ð) â ð€ð ð€ (ð) = ðŒ ðâð
Then ð€0 ({ð}) = 0 for every ð â ð ð€0 (ð ) = 1 ð€0 is 0â1 normalized. 2. Let y â core(ð, ð€0 ). Then for every ð â ð â ðŠð ⥠ð€0 (ð)
(3.82)
ðâð
â
ðŠð = 1
(3.83)
ðâð
Let w = (ð€1 , ð€2 , . . . , ð€ð ) where ð€ð = ð€({ð}). Let x = ðŒy + w. Using (3.82) and (3.83) â â ð¥ð = (ðŒðŠð + ð€ð ) ðâð
ðâð
=ðŒ
â
ðŠð +
ðâð
â ðâð
ð€ð
ðâð
⥠ðŒð€0 (ð) + 1 =ðŒ ðŒ
â â
ð€ð
ðâð
(
ð€(ð) â
â ðâð
) ð€ð
+
â
ð€ð
ðâð
= ð€(ð) â ð¥ð = (ðŒðŠð + ð€ð ) ðâð
=ðŒ+
â
ð€ð
ðâð
= ð€(ð ) Therefore, x = ðŒy + w â core(ð, ð€). Similarly, we can show that x â core(ð, ð€) =â y =
1 (x â w) â core(ð, ð€0 ) ðŒ
and therefore core(ð, ð€) = ðŒcore(ð, ð€0 ) + w 203
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3. This immediately implies
core(ð, ð€) = â
ââ core(ð, ð€0 ) = â
3.271 (ð, ð€) is 0â1 normalized, that is ð€({ð} = 0 for every ð â ð ð€(ð ) = 1 Consequently, x belongs to the core of (ð, ð€) if and only if â
ð¥ð ⥠ð€ð = 0
(3.84)
ð¥ð = ð€(ð ) = 1
(3.85)
ð¥ð ⥠ð€(ð) for every ð â ð
(3.86)
ðâð
â ðâð
(3.84) and (3.85) ensure that x = (ð¥1 , ð¥2 , . . . , ð¥ð ) is a mixed strategy for player 1 in the two-person zero-sum game. Using this mixed strategy, the expected payoï¬ to player I for any strategy ð of player II is ð¢(x, ð) =
â
ð¥ð ð¢(ð, ð) =
ðâð
â ðâð
ð¥ð
1 ð€(ð)
(3.86) implies ð¢(x, ð) =
â ðâð
ð¥ð
1 ⥠1 for every ð â ð ð€(ð)
That is any x â core(ð, ð€) provides a mixed strategy for player I which ensures a payoï¬ at least 1. That is core(ð, ð€) â= â
=â ð¿ ⥠1 Conversely, if the ð¿ < 1, there is no mixed strategy for player I which satisï¬es (3.86) and consequently no x which satisï¬es (3.84), (3.85) and (3.86). In other words, core(ð, ð€) = â
. 3.272 If ð¿ is the value of ðº, there exists a mixed strategy which will guarantee that II pays no more than ð¿. That is, there exists numbers ðŠð ⥠0 for every coalition ð â ð such that â ðŠð = 1 ðâð
and â
ðŠð ð¢(ð, ð) †ð¿
for every ð â ð
ðâð
that is â ðâð
ðŠð
1 â€ð¿ ð€(ð)
for every ð â ð
ðâð
204
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
or â ðâð
ðŠð â€1 ð¿ð€(ð)
for every ð â ð
(3.87)
ðâð
For each coalition ð â ð let ðð =
ðŠð ð¿ð€(ð)
in (3.87) â
ðð †1
ðâð ðâð
Augment the collection ð with the single-player coalitions to form the collection ⬠= ð ⪠{ {ð} : ð â ð } and with weights { ðð : ð â ð } and â
ð{ð} = 1 â
ðð
ðâð
Then ⬠is a balanced collection. Since the game (ð, ð€) is balanced 1 = ð€(ð ) â¥
â
ðð ð€(ð)
ðââ¬
=
â
ðð ð€(ð)
ðâð
â
ðŠð ð€(ð) ð¿ð€(ð) ðâ⬠1â = ðŠð ð¿ =
ðââ¬
1 = ð¿ that is 1â¥
1 ð¿
¯ = (1/ð, 1/ð, . . . , 1/ð), the payoï¬ is If I plays the mixed strategy x ð¢(¯ x, ð) =
â ðâð
1 1 = > 0 for every ð â ð ðð€(ð) ð€(ð)
Therefore ð¿ > 0 and (3.88) implies that ð¿â¥1
205
(3.88)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 3.273 Assume core(ð, ð€) â= â
and let ð¥ â core(ð, ð€). Then ðð (x) ⥠ð€(ð) for every ð â ð where ðð =
â ðâð
(3.89)
ð¥ð measures the share coalition ð at the allocation x.
Let ⬠be a balanced family of coalitions with weights ðð . For every ð â â¬, (3.89) implies ðð ðð (x) ⥠ðð ð€(ð) Summing over all ð â ⬠â
â
ðð ðð (x) â¥
ðââ¬
ðð ð€(ð)
ðââ¬
Evaluating the left hand side of this inequality â â â ðð ðð (x) = ð ð¥ð ðââ¬
ðââ¬
=
ðâð
ââ
ðð¥ð
ðâð ðââ¬
=
â
ðâð
ð¥ð
ðâð
=
â
â
ðâ⬠ðâð
ð¥ð
ðâð
= ð€(ð ) Substituting this in (3.90) gives ð€(ð ) â¥
â ðââ¬
The game is balanced.
206
ðð ð€(ð)
ð
(3.90)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 4: Smooth Functions 4.1 Along the demand curve, price and quantity are related according to the equation ð = 10 â ð¥ This is called the inverse demand function. Total revenue ð
(ð¥) (price times quantity) is given by ð
(ð¥) = ðð¥ = (10 â ð¥)ð¥ = 10ð¥ â ð¥2 = ð (ð¥) ð(ð¥) can be rewritten as ð(ð¥) = 21 + 4(ð¥ â 3) At ð¥ = 3, the price is 7 but the marginal revenue of an additional unit is only 4. The function ð decomposes (approximately) the total revenue into two components â the revenue from the sale of 3 units (21 = 3 à 7) plus the marginal revenue from the sale of additional units (4(ð¥ â 3)). 4.2 If your answer is 5 per cent, obtained by subtracting the inï¬ation rate from the growth rate of nominal GDP, you are implicitly using a linear approximation. To see this, let ð ð ðð ðð
= price level at the beginning of the year = real GDP at the beginning of the year = change in prices during year = change in output during year
We are told that nominal GDP at the end of the year, (ð + ðð)(ð + ðð), equals 1.10 times nominal GDP at the beginning of the year, ðð. That is (ð + ðð)(ð + ðð) = 1.10ðð
(4.42)
Furthermore, the price level at the end of the year, ð + ðð equals 1.05 times the price level of the start of year, ð: ð + ðð = 1.05ð Substituting this in equation (4.38) yields 1.05ð(ð + ðð) = 1.10ðð which can be solved to give ðð = (
1.10 â 1)ð = 0.0476 1.05
The growth rate of real GDP (ðð/ð) is equal to 4.76 per cent. 207
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
To show how the estimate of 5 per cent involves a linear approximation, we expand the expression for real GDP at the end of the year. (ð + ðð)(ð + ðð) = ðð + ððð + ððð + ðððð Dividing by ðð (ð + ðð)(ð + ðð) ðð ðð ðððð =1+ + + ðð ð ð ðð The growth rate of nominal GDP is (ð + ðð)(ð + ðð) â ðð (ð + ðð)(ð + ðð) = â1 ðð ðð ðð ðð ðððð = + + ð ðð ðð = Growth rate of output + Inï¬ation rate + Error term For small changes, the error term ðððð/ðð is insigniï¬cant, and we can approximate the growth rate of output according to the sum Growth rate of nominal GDP = Growth rate of output + Inï¬ation rate This is a linear approximation since it approximates the function (ð + ðð)(ð + ðð) by the linear function ðð + ððð + ððð. In eï¬ect, we are evaluating the change output at the old prices, and the change in prices at the old output, and ignoring in interaction between changes in prices and changes in quantities. The use of linear approximation in growth rates is extremely common in practice. 4.3 From (4.2) â¥x⥠ð(x) = ð (x0 + x) â ð (x0 ) â ð(x) and therefore ð(x) =
ð (x0 + x) â ð (x0 ) â ð(x) â¥xâ¥
ð(x) â 0ð as x â 0ð can be expressed as lim ð(x) = 0ð
xâ0ð
4.4 Suppose not. That is, there exist two linear maps such that ð (x0 + x) = ð (x0 ) + ð1 (x) + â¥x⥠ð1 (x) ð (x0 + x) = ð (x0 ) + ð2 (x) + â¥x⥠ð2 (x) with lim ðð (x) = 0,
xâ0
208
ð = 1, 2
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Subtracting we have ð¿1 (x) â ð¿2 (x) = â¥x⥠(ð1 (x) â ð2 (x)) and lim
xâ0
ð1 (x) â ð2 (x) =0 â¥xâ¥
Since ð1 â ð2 is linear, (4) implies that ð1 (x) = ð2 (x) for all x â ð. To see this, we proceed by contradiction. Again, suppose not. That is, suppose there exists some x â ð such that ð1 (x) â= ð2 (x) For this x, let ð=
ð1 (x) â ð2 (x) â¥xâ¥
By linearity, ð1 (ð¡x) â ð2 (ð¡x) = ð for every âð¡ > 0 â¥ð¡x⥠and therefore lim
ð¡â0
ð1 (ð¡x) â ð2 (ð¡x) = ð â= 0 â¥ð¡xâ¥
which contradicts (4). Therefore ð1 (x) = ð2 (x) for all x â ð. 4.5 If ð : ð â ð is diï¬erentiable at x0 , then ð (x0 + x) = ð (x0 ) + ð(x) + ð(x) â¥x⥠where ð(x) â 0ð as x â 0ð . Since ð is a continuous linear function, ð(x) â 0ð as x â 0ð . Therefore lim ð (x0 + x) = lim ð (x0 ) + lim ð(x) + lim ð(x) â¥xâ¥
xâ0
xâ0
xâ0
xâ0
= ð (x0 ) ð is continuous. 4.6 4.7 4.8 The approximation error at the point (2, 16) is ð (2, 16) ð(2, 16) Absolute error Percentage error Relative error
=8.0000 =11.3333 =-3.3333 =-41.6667 =-4.1667
By contrast, â(2, 16) = 8 = ð (2, 16). Table 4.1 shows that â gives a good approximation to ð in the neighborhood of (2, 16). 209
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Table 4.1: Approximating the Cobb-Douglas function at (2, 16) x x0 + x At their intersection: (0.0, 0.0) (2.0, 16.0)
Approximation Error Percentage Relative
ð (x0 + x)
â(x0 + x)
8.0000
8.0000
0.0000
NIL
Around the unit circle: (1.0, 0.0) (3.0, 16.0) (0.7, 0.7) (2.7, 16.7) (0.0, 1.0) (2.0, 17.0) (-0.7, 0.7) (1.3, 16.7) (-1.0, 0.0) (1.0, 16.0) (-0.7, -0.7) (1.3, 15.3) (0.0, -1.0) (2.0, 15.0) (0.7, -0.7) (2.7, 15.3)
9.1577 9.1083 8.3300 7.1196 6.3496 6.7119 7.6631 8.5867
9.3333 9.1785 8.3333 7.2929 6.6667 6.8215 7.6667 8.7071
-1.9177 -0.7712 -0.0406 -2.4342 -4.9934 -1.6323 -0.0466 -1.4018
-0.1756 -0.0702 -0.0034 -0.1733 -0.3171 -0.1096 -0.0036 -0.1204
Around a smaller circle: (0.10, 0.00) (2.1, 16.0) (0.07, 0.07) (2.1, 16.1) (0.00, 0.10) (2.0, 16.1) (-0.07, 0.07) (1.9, 16.1) (-0.10, 0.00) (1.9, 16.0) (-0.07, -0.07) (1.9, 15.9) (0.00, -0.10) (2.0, 15.9) (0.07, -0.07) (2.1, 15.9)
8.1312 8.1170 8.0333 7.9279 7.8644 7.8813 7.9666 8.0693
8.1333 8.1179 8.0333 7.9293 7.8667 7.8821 7.9667 8.0707
-0.0266 -0.0103 -0.0004 -0.0181 -0.0291 -0.0110 -0.0004 -0.0171
-0.0216 -0.0083 -0.0003 -0.0143 -0.0229 -0.0087 -0.0003 -0.0138
Parallel to the (-2.0, 0.0) (-1.0, 0.0) (-0.5, 0.0) (-0.1, 0.0) (0.0, 0.0) (0.1, 0.0) (0.5, 0.0) (1.0, 0.0) (2.0, 0.0) (4.0, 0.0)
x1 axis: (0.0, 16.0) (1.0, 16.0) (1.5, 16.0) (1.9, 16.0) (2.0, 16.0) (2.1, 16.0) (2.5, 16.0) (3.0, 16.0) (4.0, 16.0) (6.0, 16.0)
0.0000 6.3496 7.2685 7.8644 8.0000 8.1312 8.6177 9.1577 10.0794 11.5380
5.3333 6.6667 7.3333 7.8667 8.0000 8.1333 8.6667 9.3333 10.6667 13.3333
NIL -4.9934 -0.8922 -0.0291 0.0000 -0.0266 -0.5678 -1.9177 -5.8267 -15.5602
-2.6667 -0.3171 -0.1297 -0.0229 NIL -0.0216 -0.0979 -0.1756 -0.2936 -0.4488
Parallel to the (0.0, -4.0) (0.0, -2.0) (0.0, -1.0) (0.0, -0.5) (0.0, -0.1) (0.0, 0.0) (0.0, 0.1) (0.0, 0.5) (0.0, 1.0) (0.0, 2.0) (0.0, 4.0)
x2 axis: (2.0, 12.0) (2.0, 14.0) (2.0, 15.0) (2.0, 15.5) (2.0, 15.9) (2.0, 16.0) (2.0, 16.1) (2.0, 16.5) (2.0, 17.0) (2.0, 18.0) (2.0, 20.0)
6.6039 7.3186 7.6631 7.8325 7.9666 8.0000 8.0333 8.1658 8.3300 8.6535 9.2832
6.6667 7.3333 7.6667 7.8333 7.9667 8.0000 8.0333 8.1667 8.3333 8.6667 9.3333
-0.9511 -0.2012 -0.0466 -0.0112 -0.0004 0.0000 -0.0004 -0.0105 -0.0406 -0.1522 -0.5403
-0.0157 -0.0074 -0.0036 -0.0018 -0.0003 NIL -0.0003 -0.0017 -0.0034 -0.0066 -0.0125
210
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.9 To show that ð is nonlinear, consider ð((1, 2, 3, 4, 5) + (66, 55, 75, 81, 63)) = ð(67, 57, 78, 85, 68) = (85, 78, 68, 67, 58) â= (5, 4, 3, 2, 1) + (81, 75, 67, 63, 55) To show that ð is diï¬erentiable, consider a particular point, say (66, 55, 75, 81, 63). Consider the permutation ð : âð â âð deï¬ned by ð(ð¥1 , ð¥2 , . . . , ð¥5 ) = (ð¥4 , ð¥3 , ð¥1 , ð¥5 , ð¥2 ) ð is linear and ð(66, 55, 75, 81, 63) = (81, 75, 67, 63, 55) = ð(66, 55, 75, 81, 63) Furthermore, ð(x) = ð(x) for all x close to (66, 55, 75, 81, 63). Hence, ð(x) approximates ð(x) in a neighborhood of (66, 55, 75, 81, 63) and so ð is diï¬erentiable at (66, 55, 75, 81, 63). The choice of (66, 55, 75, 81, 63) was arbitrary, and the argument applies at every x such that xð â= xð . In summary, each application of ð involves a permutation, although the particular permutation depends upon the argument, x. However, for any given x0 with x0ð â= x0ð , the same permutation applies to all x in the neighborhood of x0 , so that the permutation (which is a linear function) is the derivative of ð at x0 . 4.10 Using (4.3), we have for any x ð (x0 + ð¡x) â ð (x0 ) â ð·ð [x0 ](ð¡x) =0 ð¡xâ0 â¥ð¡x⥠lim
or ð (x0 + ð¡x) â ð (x0 ) â ð¡ð·ð [x0 ](x) =0 ð¡â0 ð¡ â¥x⥠lim
For â¥x⥠= 1, this implies ð¡ð·ð [x0 ](x) ð (x0 + ð¡x) â ð (x0 ) = ð¡â0 ð¡ ð¡ lim
that is 0 0 â x ð [x0 ] = lim ð (x + ð¡x) â ð (x ) = ð·ð [x0 ](x) ð· ð¡â0 ð¡
4.11 By direct calculation â(x0ð + ð¡) â â(x0ð ) ð¡â0 ð¡ ð (x01 , x02 , . . . , x0ð + ð¡, . . . , x0ð ) â ð (x01 , x02 , . . . , x0ð , . . . , x0ð ) = lim ð¡â0 ð¡ ð (x0 + ð¡eð ) â ð (x0 ) = lim ð¡â0 ð¡ 0 â = ð·eð ð [x ]
ð·ð¥ð ð [x0 ] = lim
211
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.12 Deï¬ne the function ( ) â(ð¡) = ð (8, 8) + ð¡(1, 1) = (8 + ð¡)1/3 (8 + ð¡)2/3 =8+ð¡ The directional derivative of ð in the direction (1, 1) is â (1,1) ð (8, 8) = lim â(ð¡) â â(0) ð· ð¡â0 ð¡ =1 Generalization of this example reveals that the directional derivative of ð along any â x0 ð [x0 ] = 1 for every x0 . Economically, ray through the origin equals 1, that is ð· this means that increasing inputs in the same proportions leads to a proportionate increase in output, which is the property of constant returns to scale. We will study this property of homogeneity is some depth in Section 4.6. 4.13 Let p = âð (x0 ). Each component of p represents the action of the derivative on an element of the standard basis {e1 , e2 , . . . , eð }(see proof of Theorem 3.4) ðð = ð·ð [x0 ](eð ) Since â¥eð ⥠= 1, ð·ð [x0 ](eð ) is the directional derivative at x0 in the direction eð (Exercise 4.10) â eð (x0 ) ðð = ð·ð [x0 ](eð ) = ð· But this is simply the ð partial derivative of ð (Exercise 4.11) â eð (x0 ) = ð·ð¥ð ð (x0 ) ðð = ð·ð [x0 ](eð ) = ð· 4.14 Using the standard inner product on âð (Example 3.26) and Exercise 4.13 < âð (x0 ), x >=
ð â
ð·ð¥ð ð [x0 ]xð = ð·ð [x0 ](x)
ð=1
4.15 Since ð is diï¬erentiable ð (x1 + ð¡x) = ð (x1 ) + âð (x0 )ð ð¡x + ð(ð¡x) â¥ð¡x⥠with ð(ð¡x) â 0 as ð¡x â 0. If ð is increasing, ð (x1 + ð¡x) ⥠ð (x1 ) for every x ⥠0 and ð¡ > 0. Therefore âð (x0 )ð ð¡x + ð(ð¡x) â¥ð¡x⥠= ð¡âð (x0 )ð x + ð¡ð(ð¡x) â¥x⥠⥠0 Dividing by ð¡ and letting ð¡ â 0 âð (x0 )ð x ⥠0 for every x ⥠0 In particular, this applies for unit vectors eð . Therefore ð·ð¥ð ð (x1 ) ⥠0,
ð = 1, 2, . . . , ð
212
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
â x ð (x0 ) measures the rate of increase of ð in the di4.16 The directional derivative ð· rection x. Using Exercises 4.10, 4.14 and 3.61, assuming x has unit norm, â x ð (x0 ) = ð·ð [x0 ](x) =< âð (x0 ), x >†âð (x0 ) ð· This bound is attained when x = âð (x0 )/ âð (x0 ) since âð (x0 )2 âð (x0 ) 0 0 â ð·x ð (x ) =< âð (x ), >= = âð (x0 ) 0 0 â¥âð (x )⥠â¥âð (x )⥠The directional derivative is maximized when âð (x0 ) and x are aligned. 4.17 Using Exercise 4.14 ð» = { x â ð :< âð [x0 ], x >= 0 } 4.18 Assume each ðð is diï¬erentiable at x0 and let ð·ð [x0 ] = (ð·ð1 [x0 ], ð·ð2 [x0 ], . . . , ð·ðð [x0 ]) Then
â â â f (x0 + x) â f [x0 ] â ð·f [x0 ]x = â â
ð1 (x0 + x) â ð1 [x0 ] â ð·ð1 [x0 ]x ð2 (x0 + x) â ð2 [x0 ] â ð·ð2 [x0 ]x .. .
â â â â â
ðð (x0 + x) â ðð (x0 ) â ð·ðð [x0 ]x and ðð (x0 + x) â ðð (x0 ) â ð·ðð [x0 ]x â 0 as â¥x⥠â 0 â¥x⥠for every ð implies f (x0 + x) â f (x0 ) â ð·f [x0 ](x) â 0 as â¥x⥠â 0 â¥xâ¥
(4.43)
Therefore f is diï¬erentiable with derivative ð·f [x0 ] = ð¿ = (ð·ð1 (x0 ), ð·ð2 [x0 ], . . . , ð·ðð [x0 ]) Each ð·ðð [x0 ] is represented by the gradient âðð [x0 ] (Exercise 4.13) and therefore ð·ð [x0 ] is represented by the matrix â â â â âð1 [x0 ] ð·ð¥1 ð1 [x0 ] ð·ð¥2 ð1 [x0 ] . . . ð·ð¥ð ð1 [x0 ] â âð2 [x0 ] â â ð·ð¥1 ð2 [x0 ] ð·ð¥2 ð2 [x0 ] . . . ð·ð¥ð ð2 [x0 ] â â â â â ðœ =â â=â â .. .. .. .. .. â â â â . . . . . âðð [x0 ]
ð·ð¥1 ðð [x0 ] ð·ð¥2 ðð [x0 ] . . .
ð·ð¥ð ðð [x0 ]
Conversely, if f is diï¬erentiable, its derivative ð·f [x0 ] : âð â âð be decomposed into ð component ð·ð1 [x0 ], ð·ð2 [x0 ], . . . , ð·ðð [x0 ] functionals such that â â ð1 (x0 + x) â ð1 (x0 ) â ð·ð1 [x0 ]x â ð2 (x0 + x) â ð2 (x0 ) â ð·ð2 [x0 ]x â â â f (x0 + x) â f (x0 ) â ð·f [x0 ]x = â â .. â â . ðð (x0 + x) â ðð (x0 ) â ð·ðð [x0 ]x
213
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
(4.43) implies that ðð (x0 + x) â ðð (x0 ) â ð·ðð [x0 ]x â 0 as â¥x⥠â 0 â¥x⥠for every ð. 4.19 If ð·ð [x0 ] has full rank, then it is one-to-one (Exercise 3.25) and onto (Exercise 3.16). Therefore ð·ð [x0 ] is nonsingular. The Jacobian ðœð (x0 ) represents ð·ð [x0 ], which is therefore nonsingular if and only if det ðœð (x0 ) â= 0. 4.20 When ð is a functional, rank ð ⥠ððððð = 1. If ð·ð [x0 ] has full rank (1), then ð·ð [x0 ] maps ð onto â (Exercise 3.16), which requires that âð (x0 ) â= 0. 4.21 4.23 If ð : ð à ð â ð is bilinear ð (x0 + x, y0 + y) = ð (x0 , y0 ) + ð (x0 , y) + ð (x, y0 ) + ð (x, y) Deï¬ning ð·ð [x0 , y0 ](x, y) = ð (x0 , y) + ð (x, y0 ) ð (x0 + x, y0 + y) = ð (x0 , y0 ) + ð·ð [x0 , y0 ](x, y) + ð (x, y) Since ð is continuous, there exists ð such that ð (x, y) †ð â¥x⥠â¥yâ¥
for every x â ð and y â ð
and therefore NOTE This is not quite right. See Spivak p. 23. Avez (Tilburg) has ( )2 â¥ð (x, y)⥠†ð â¥x⥠â¥y⥠†ð â¥x⥠+ â¥y⥠†ð â¥(x, y)â¥2 which implies that â¥ð (x, y)⥠â 0 as (x, y) â 0 â¥(x, y)â¥
lim
x1 ,x2 â0
ð (x1 , x2 ) =0 â¥x1 ⥠â¥x2 â¥
Therefore ð is diï¬erentiable with derivative ð·ð [x0 , y0 ] = ð (x0 , y) + ð (x, y0 ) 4.24 Deï¬ne ð : â2 â â by ð(ð§1 , ð§2 ) = ð§1 ð§2 Then ð is bilinear (Example 3.23) and continuous (Exercise 2.79) and therefore diï¬erentiable (Exercise 4.23) with derivative ð·ð[ð§1 , z2 ] = ð(z1 , â
) + ð(â
, z2 ) 214
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
The function ð ð is the composition of ð with ð and ð, ð ð(x, y) = ð(ð (x), ð(y)) By the chain rule, the derivative of ð ð is ( ) ð·ð ð[x, y] = ð·ð[ð§1 , z2 ] ð·ð [x], ð·ð[x] = ð(z1 , ð·ð[y]) + ð(ð·ð [x], z2 ) = ð [x]ð·ð[y]) + ð(y)ð·ð [x] where z1 = ð (x) and z2 = ð(y). 4.25 For ð = 1, ð (ð¥) = ð¥ is linear and therefore (Exercise 4.6) ð·ð [ð¥] = 1 (ð·ð [ð¥](ð¥) = ð¥). For ð = 2, let ð(ð¥) = ð¥ so that ð (ð¥) = ð¥2 = ð(ð¥)ð(ð¥). Using the product rule ð·ð [x] = ð(ð¥)ð·ð(ð¥) + ð(ð¥)ð·ð(ð¥) = 2ð¥ Now assume it is true for ð â 1 and let ð(ð¥) = ð¥ðâ1 , so that ð (x) = ð¥ð(ð¥). By the product rule ð·ð [x] = ð¥ð·ð[ð¥] + ð(ð¥)1 By assumption ð·ð[ð¥] = (ð â 1)ð¥ðâ2 and therefore ð·ð [x] = ð¥ð·ð[ð¥] + ð(ð¥)1 = ð¥(ð â 1)ð¥ðâ2 + ð¥ðâ1 = ðð¥ðâ1 4.26 Using the product rule (Exercise 4.24) ð·ð¥ ð
(ð¥0 ) = ð (ð¥0 )ð·ð¥ ð¥ + ð¥0 ð·ð¥ ð (ð¥0 ) = ð0 + ð¥0 ð·ð¥ ð (ð¥0 ) where ð0 = ð (ð¥0 ). Marginal revenue equals one unit at the current price minus the reduction in revenue caused by reducing the price on existing sales. ( )â1 4.27 Fix some x0 and let ð = ð·ð [x0 ] . Let y0 = ð (x0 ). For any y, let x = â1 0 â1 0 0 ð (y + y) â ð (y ) so that ð(y) = ð (x + x) â ð (x) and â1 0 ( ) ð (y + y) â ð â1 (y0 ) â ð(y) = (x â ð ð (x0 + x) â ð (x0 )) Since ð is diï¬erentiable at x0 with ð·ð [x0 ] = ð â1 ð (x0 + x) â ð (x0 ) = ð â1 (x) + ð(x) â¥x⥠Substituting ( ) â1 0 ð (y + y) â ð â1 (y0 ) â ð(y) = x â ð ð â1 (x) + ð(x) â¥x⥠( ) = ð ð(x) â¥x⥠( ) = â¥x⥠ð ð(x) ( ) with ð(x) â 0ð as x â 0ð . Since ð â1 and ð are continuous, ð ð(x) â 0ð as y â 0. )â1 ( . We conclude that ð â1 is diï¬erentiable with derivative ð = ð·ð [x0 ]
215
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 4.28 log ð (ð¥) = ð¥ log ð and therefore
( ) ð (ð¥) = exp log ð (ð¥) = ðð¥ log ð
By the Chain Rule, ð is diï¬erentiable with derivative ð·ð¥ ð (ð¥) = ðð¥ log ð log ð = ðð¥ log ð 4.29 By Exercise 4.15, the function ð : â â â deï¬ned by ð(ðŠ) = tiable with derivative 1 ð·ðŠ ð[ðŠ] = âðŠ â2 = â 2 ðŠ
1 ðŠ
= ðŠ â1 is diï¬eren-
Applying the Chain Rule, 1/ð = ð â ð is diï¬erentiable with derivative 1 ð·ð [x] ð· [x] = ð·ð[ð (x)]ð·ð [x] = â ( )2 ð ð (x) 4.30 Applying the Product Rule to ð à (1/ð) 1 1 ð ð·ð [x] ð· [x, y] = ð (x)ð· [y] + ð ð ð(y) ð·ð[y] 1 = âð (x) ( ð·ð [x] )2 + ð(y) ð(y) ð(y)ð·ð [x] â ð (x)ð·ð[y] = ( )2 ð(y) 4.31 In the particular case where 1/3 2/3
ð (x1 , x2 ) = x1 x2 the partial derivatives at the point (8, 8) are ð·ð¥1 ð [(8, 8)] =
2 1 and ð·ð¥2 ð¹ [(8, 8)] = 3 3
4.32 The partial derivatives of ð (x) are from Table 4.4 ð·ð¥ð ð [x] = ð¥ð1 1 ð¥ð2 2 . . . ðð ð¥ððð â1 . . . ð¥ððð = ðð so that the gradient is
( âð (x) =
ð (x) ð¥ð ð1 ð2 ðð , ,..., ð¥1 ð¥2 ð¥ð
) ð (x)
4.33 Applying the chain rule (Exercise 4.22) to general power function (Example 4.15), the partial derivatives of the CES function are ð·ð¥ð ð [x] =
1 1 â1 (ð1 ð¥ð1 + ð2 ð¥ð2 + â
â
â
+ ðð ð¥ðð ) ð ðð ðð¥ðâ1 ð ð
= ðð ð¥ðâ1 (ð1 ð¥ð1 + ð2 ð¥ð2 + â
â
â
+ ðð ð¥ðð ) ð ( )1âð ð (x) = ðð ð¥ð 216
1âð ð
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.34 Deï¬ne â(ð¥) = ð (ð¥) â
ð (ð) â ð (ð) (ð¥ â ð) ðâð
Then â is continuous on [ð, ð] and diï¬erentiable on (ð, ð) with â(ð) = ð (ð) â
ð (ð) â ð (ð) (ð â ð)ð (ð) = â(ð) ðâð
By Rolleâs theorem (Exercise 5.8), there exists ð¥ â (ð, ð) such that ââ² (ð¥) = ð â² (ð¥) â
ð (ð) â ð (ð) =0 ðâð
4.35 Assume âð (x) ⥠0 for every x â ð. By the mean value theorem, for any x2 ⥠x1 ¯ â (x1 , x2 ) such that in ð, there exists x ð (x2 ) = ð (x1 ) + ð·ð [¯ x](x2 â x1 ) Using (4.6) ð (x2 ) = ð (x1 ) +
ð â
ð·ð¥ð ð (¯ x)(ð¥2ð â ð¥1ð )
(4.44)
ð=1
âð (¯ x) ⥠0 and x2 ⥠x1 implies that ð â
ð·ð¥ð ð (¯ x)(ð¥2ð â ð¥1ð ) ⥠0
ð=1
and therefore ð (x2 ) ⥠ð (x1 ). ð is increasing. The converse was established in Exercise 4.15 4.36 âð (¯ x) > 0 and x2 ⥠x1 implies that ð â
ð·ð¥ð ð (¯ x)(ð¥2ð â ð¥1ð ) > 0
ð=1
Substituting in (4.44) ð (x2 ) = ð (x1 ) +
ð â
ð·ð¥ð ð (¯ x)(ð¥2ð â ð¥1ð ) > ð (x1 )
ð=1
ð is strictly increasing. 4.37 Diï¬erentiability implies the existence of the gradient and hence the partial derivatives of ð (Exercise 4.13). Continuity of ð·ð [x] implies the continuity of the partial derivatives. To prove the converse, choose some x0 â ð and deï¬ne for the partial functions âð (ð¡) = ð (ð¥01 , ð¥02 , . . . , ð¥0ðâ1 , ð¡, ð¥0ð+1 + ð¥ð+1 , . . . , ð¥0ð + ð¥ð )
ð = 1, 2, . . . , ð
so that ââ²ð (ð¡) = ð·ð¥ð ð (xð ) where xð = (ð¥01 , ð¥02 , . . . , ð¥0ð , ð¡, ð¥0ð+1 + ð¥ð+1 , . . . , ð¥0ð + ð¥ð ). Further, â1 (ð¥01 + ð¥1 ) = ð (x0 + x), âð (ð¥0ð ) = ð (x0 ), and âð (ð¥0ð + ð¥ð ) = âðâ1 (ð¥0ð ) so that ð (x0 + x) â ð (x0 ) =
ð â ( ) âð (ð¥0ð + ð¥ð ) â âð (ð¥0ð ) ð=1
217
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
By the mean value theorem, there exists, for each ð, ð¡Â¯ð between ð¥0ð + ð¥ð and ð¥ð such that âð (ð¥0ð + ð¥ð ) â âð (ð¥ð ) = ð·ð¥ð ð (¯ xð )ð¥ð ¯ ð = (ð¥01 , ð¥02 , . . . , ð¥0ð , ð¡Â¯, ð¥0ð+1 + ð¥ð+1 , . . . , ð¥0ð + ð¥ð ). Therefore where x ð (x0 + x) â ð (x0 ) =
ð â
ð·ð¥ð ð (¯ xð )ð¥ð
ð=1
Deï¬ne the linear functional ð(x) =
ð â
ð·ð¥ð ð (x0 )ð¥ð
ð=1
Then ð (x0 + x) â ð (x0 ) â ð(x) =
ð ( ) â ð·ð¥ð ð (¯ xð ) â ð·ð¥ð ð (x0 ) ð¥ð ð=1
and ð â ð (x0 + x) â ð (x0 ) â ð(x) †(ð·ð¥ð ð (¯ xð ) â ð·ð¥ð ð (x0 ) â£ð¥ð ⣠ð=1
so that
ð ð (x0 + x) â ð (x0 ) â ð(x) â â£ð¥ð ⣠(ð·ð¥ð ð (¯ †lim xð ) â ð·ð¥ð ð (x0 ) xâ0 â¥x⥠â¥x⥠ð=1 â€
ð â (ð·ð¥ð ð (¯ xð ) â ð·ð¥ð ð (x0 ) ð=1
=0 since the partial derivatives ð·ð¥ð ð (x) are continuous. Therefore ð is diï¬erentiable with derivative ð(x) =
ð â
ð·ð¥ð ð [x0 ]ð¥ð
ð=1
4.38 For every x1 , x2 â ð â¥ð (x1 ) â ð (x2 )⥠â€
sup
xâ[x1 ,x2 ]
â¥ð·ð (x)⥠â¥x1 â x2 â¥
by Corollary 4.1.1. If ð·ð [x] = 0 for every x â ð, then â¥ð (x1 ) â ð (x2 )⥠= 0 which implies that ð (x1 ) = ð (x2 ). We conclude that ð is constant on ð. The converse was established in Exercise 4.7. 4.39 For any x0 â ð, let ðµ â ð be an open ball of radius of radius ð centered on x0 . Applying the mean value inequality (Corollary 4.1.1) to ðð â ðð we have ( ) ðð (x) â ðð (x) â ðð (x0 ) â ðð (x0 ) †sup â¥ð·ðð [¯ x] â ð·ðð [¯ x]⥠â¥x â x0 ⥠¯ âðµ x
†ð sup â¥ð·ðð [¯ x] â ð·ðð [¯ x]⥠¯ âðµ x
218
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
for every x â ðµ. Given ð > 0, there exists ð such that for every ð, ð > ð â¥ð·ðð â ð·ðð ⥠< ð/ð and â¥ð·ðð â ð⥠< ð Letting ð â â
( ) ðð (x) â ð (x) â ðð (x0 ) â ð (x0 ) †ð â¥x â x0 â¥
(4.45)
for ð ⥠ð and x â ðµ. Applying the mean value inequality to ðð , there exists ð¿ such that â¥ðð (x) â ðð (x0 )⥠†ð â¥x â x0 â¥
(4.46)
Using (4.45) and (4.46) and the fact that â¥ð·ðð â ð⥠< ð we deduce that â¥ð (x) â ð (x0 ) â ð(x0 )⥠†3ð â¥x â x0 ⥠ð is diï¬erentiable with derivative ð. 4.40 Deï¬ne ð (ð¥) =
ðð¥+ðŠ ððŠ
By the chain rule (Exercise 4.22) ð â² (ð¥) =
ðð¥+ðŠ = ð (ð¥) ððŠ
which implies (Example 4.21) that ð (ð¥) =
ðð¥+ðŠ = ðŽðð¥ for some ðŽ â â ððŠ
Evaluating at ð¥ = 0 using ð0 = 1 gives ð (0) =
ððŠ = ðŽ for some ðŽ â â ððŠ
so that ð (ð¥) =
ððŠ ð¥ ðð¥+ðŠ = ð ððŠ ððŠ
which implies that ðð¥+ðŠ = ðð¥ ððŠ 4.41 If ð = ðŽð¥ð , ð â² (ð¥) = ððŽð¥ðâ1 and ðž(ð¥) = ð¥
ððŽð¥ðâ1 =ð ðŽð¥ð
To show that this is the only function with constant elasticity, deï¬ne ð(ð¥) =
ð (ð¥) ð¥ð
ð is diï¬erentiable (Exercise 4.30) with derivative ð â² (ð¥) =
ð¥ð ð â² (ð¥) â ð (ð¥)ðð¥ðâ1 ð¥ð â² (ð¥) â ðð (ð¥) = ð¥2ð ð¥ð+1 219
(4.47)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics If ðž(ð¥) = ð¥
ð â² (ð¥) =ð ð (ð¥)
then ð¥ð â² (ð¥) = ðð (ð¥) Substituting in (4.47) ð â² (ð¥) =
ð¥ð â² (ð¥) â ðð (ð¥) = 0 for every ð¥ â â ð¥ð+1
Therefore, ð is a constant function (Exercise 4.38). That is, there exists ðŽ â â such that ð(ð¥) =
ð (ð¥) = ðŽ or ð (ð¥) = ðŽð¥ð ð¥ð
4.42 Deï¬ne ð : ð â ð by ð(x) = ð (x) â ð·ð [x0 ](x) ð is diï¬erentiable with ð·ð[x] = ð·ð [x] â ð·ð [x0 ] Applying Corollary 4.1.1 to ð, â¥ð(x1 ) â ð(x2 )⥠â€
sup
xâ[x1 ,x2 ]
â¥ð·ð[x]⥠â¥x1 â x2 â¥
for every x1 , x2 â ð. Substituting for ð and ð·ð â¥ð (x1 ) â ð·ð [x0 ](x1 ) â ð (x2 ) + ð·ð [x0 ](x2 )⥠= â¥ð (x1 ) â ð (x2 ) â ð·ð [x0 ](x1 â x2 )⥠â€
sup
xâ[x1 ,x2 ]
â¥ð·ð [x] â ð·ð [x0 ]⥠â¥x1 â x2 â¥
4.43 Since ð·ð is continuous, there exists a neighborhood ð of x0 such that â¥ð·ð [x] â ð·ð [x0 ]⥠< ð for every x â ð and therefore for every x1 , x2 â ð sup
xâ[x1 ,x2 ]
â¥ð·ð [x] â ð·ð [x0 ]⥠< ð
By the previous exercise (Exercise 4.42) â¥ð (x1 ) â ð (x2 ) â ð·ð [x0 ](x1 â x2 )⥠†ð â¥x1 â x2 ⥠4.44 By the previous exercise (Exercise 4.43), there exists a neighborhood such that â¥ð (x1 ) â ð (x2 ) â ð·ð [x0 ](x1 â x2 )⥠†ð â¥x1 â x2 ⥠The Triangle Inequality (Exercise 1.200) implies â¥ð (x1 ) â ð (x2 )⥠â â¥ð·ð [x0 ](x1 â x2 )⥠†â¥ð (x1 ) â ð (x2 ) â ð·ð [x0 ](x1 â x2 )⥠†ð â¥x1 â x2 ⥠and therefore â¥ð (x1 ) â ð (x2 )⥠†â¥ð·ð [x0 ](x1 â x2 )⥠+ ð â¥x1 â x2 ⥠†â¥ð·ð [x0 ] + ð⥠â¥x1 â x2 ⥠220
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 4.45 Assume not. That is, assume that y = ð (x1 ) â ð (x2 ) ââ conv ðŽ
Then by the (strong) separating hyperplane theorem (Proposition 3.14) there exists a linear functional ð on ð such that ð(y) > ð(a)
for every a â ðŽ
(4.48)
where ð(ðŠ) = ð(ð (x1 ) â ð (x2 )) = ð(ð (x1 )) â ð(ð (x2 )) ðð is a functional on ð. By the mean value theorem (Theorem 4.1), there exists some ¯ â [x1 , x2 ] such that x x](x1 â x2 ) = ð â ð·ð [¯ x](x â x2 ) = ð(ð) ð â ð (x1 ) â ð â ð (x2 )) = ð·(ð â ð )[¯ for some ð â ðŽ contradicting (4.44). 4.46 Deï¬ne â : [ð, ð] â â by
( ) ( ) â(ð¥) = ð (ð) â ð (ð) ð(ð¥) â ð(ð) â ð(ð) ð (ð¥)
â â ð¶[ð, ð] and is diï¬erentiable on ð, ð) with ( ) ( ) â(ð) = ð (ð) â ð (ð) ð(ð) â ð(ð) â ð(ð) ð (ð) = ð (ð)ð(ð) â ð (ð)ð(ð) = â(ð) By Rolleâs theorem (Exercise 5.8), there exists ð¥ â (ð, ð) such that ( ) ( ) ââ² (ð¥) = ð (ð) â ð (ð) ð â² (ð¥) â ð(ð) â ð(ð) ð â² (ð¥) = 0 4.47 The hypothesis that limð¥âð ð·ð (ð¥)/ð·ð(ð¥) exists contains two implicit assumptions, namely â ð and ð are diï¬erentiable on a neighborhood ð of ð (except perhaps at ð) â ð â² (ð¥) â= 0 in this neighborhood (except perhaps at ð). Applying the Cauchy mean value theorem, for every ð¥ â ð, there exists some ðŠð¥ â (ð, ð¥) such that ð â² (ðŠð¥ ) ð (ð¥) â ð (ð) ð (ð¥) = = ð â² (ðŠð¥ ) ð(ð¥) â ð(ð) ð(ð¥) and therefore ð (ð¥) ð â² (ðŠð¥ ) ð â² (ð¥) = lim â² = lim â² ð¥âð ð(ð¥) ð¥âð ð (ðŠð¥ ) ð¥âð ð (ð¥) lim
4.48 Let ðŽ = ð1 + ð2 + â
â
â
+ ðð â= 1. Then from (4.12) ð1 log ð¥1 + ð2 log ð¥2 + . . . ðð log ð¥ð ðŽ ð2 ðð ð1 log ð¥1 + log ð¥2 + . . . log ð¥ð = ðŽ ðŽ ðŽ
lim ð(ð) =
ðâ0
and therefore lim log ð (ð, x) =
ðâ0
ð1 ð2 ðð log ð¥1 + log ð¥2 + . . . log ð¥ð ðŽ ðŽ ðŽ
so that ð1
ð1
ð1
lim ð (ð, x) = ð¥1ðŽ ð¥2ðŽ . . . ð¥ððŽ
ðâ0
which is homogeneous of degree one. 221
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.49 Average cost is given by ð(ðŠ)/ðŠ which is undeï¬ned at ðŠ = 0. We seek limðŠâ0 ð(ðŠ)/ðŠ. By LâHË opitalâs rule ð(ðŠ) ðâ² (ðŠ) = lim ðŠâ0 ðŠ ðŠâ0 1 = ðâ² (0) lim
which is marginal cost at zero output. 4.50
1. Since limð¥ââ ð â² (ð¥)/ð â² (ð¥) = k, for every ð > 0 there exists ð such that â² ð (¯ ð¥) â ð < ð/2 for every 𥠯>ð (4.49) ð â² (¯ ð¥) For every ð¥ > ð, there exists (Exercise 4.46) 𥠯 â (ð, ð¥) such that ð (ð¥) â ð (ð) ð â² (¯ ð¥) = â² ð(ð¥) â ð(ð) ð (¯ ð¥) and therefore by (4.49) ð (ð¥) â ð (ð) < ð/2 for every ð¥ > ð â ð ð(ð¥) â ð(ð)
2. ð (ð¥) ð (ð¥) â ð (ð) ð (ð¥) ð(ð¥) â ð(ð) = à à ð(ð¥) ð(ð¥) â ð(ð) ð (ð¥) â ð (ð) ð(ð¥) ð (ð¥) â ð (ð) 1 â à = ð(ð¥) â ð(ð) 1â
ð(ð) ð(ð¥) ð (ð) ð (ð¥)
For ï¬xed ð lim
1â
ð¥ââ
1â
ð(ð) ð(ð¥) ð (ð) ð (ð¥)
=1
and therefore there exists ð2 such that 1â 1â
ð(ð) ð(ð¥) ð (ð) ð (ð¥)
< 2 for every ð¥ > ð2
which implies that ð ð (ð¥) ð(ð¥) â ð < 2 à 2 for every ð¥ > ð = max{ð1 , ð2 } 4.51 We know that the result holds for ð = 1 (Exercise 4.22). Assume that the result holds for ð â 1. By the chain rule ð·(ð â ð )[x] = ð·ð[ð (x)] â ð·ð [x] If ð, ð â ð¶ ð , the ð·ð, ð·ð â ð¶ ðâ1 and therefore (by assumption) ð·(ð â ð ) â ð¶ ðâ1 , which implies that ð â ð â ð¶ ð . 222
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 4.52 The partial derivatives of the quadratic function are ð·1 ð = 2ðð¥1 + 2ðð¥2 ð·2 ð = 2ðð¥1 + 2ðð¥2 The second-order partial derivatives are ð·11 ð = 2ð
ð·21 ð = 2ð
ð·12 ð = 2ð
ð·22 ð = 2ð
4.53 Apply Exercise 4.37 to each partial derivative ð·ð ð [x]. 4.54 ð»(x0 ) =
(
ð·11 ð ð [x0 ] ð·12 ð ð [x0 ] ð·21 ð ð [x0 ] ð·22 ð ð [x0 ]
)
( =2
ð ð
ð ð
)
4.55 4.56 For any ð¥1 â ð, deï¬ne ð : ð â â by ð(ð¡) = ð (ð¡) + ð â² [ð¡](ð¥1 â ð¡) + ð2 (ð¥1 â ð¡)2 ð is diï¬erentiable on ð with ðâ² (ð¡) = ð â² [ð¡] â ð â² [ð¡] + ð â²â² [ð¡](ð¥1 â ð¡) â 2ð2 (ð¥1 â ð¡) = ð â²â² [ð¡](ð¥1 â ð¡) â 2ð2 (ð¥1 â ð¡) Note that ð(ð¥1 ) = ð (ð¥1 ) and ð(ð¥0 ) = ð (ð¥0 ) + ð â² (ð¥0 )(ð¥1 â ð¥0 ) + ð2 (ð¥1 â ð¥0 )2
(4.50)
is a quadratic approximation for ð near ð¥0 . If we require that this be exact at ð¥1 â= ð¥0 , then ð(ð¥0 ) = ð (ð¥1 ) = ð(ð¥1 ). By the mean value theorem (Theorem 4.1), there exists some 𥠯 between ð¥0 and ð¥1 such that ð(ð¥1 ) â ð(ð¥0 ) = ðâ² (¯ ð¥)(ð¥1 â ð¥0 ) = ð â²â² (¯ ð¥)(ð¥1 â ð¥0 ) â 2ð2 (ð¥1 â ð¡) = 0 which implies that ð2 =
1 â²â² ð (¯ ð¥) 2
Setting ð¥ = ð¥1 â ð¥0 in (4.50) gives the required result. 4.57 For any ð¥1 â ð, deï¬ne ð : ð â â by 1 1 ð(ð¡) = ð (ð¡) + ð â² [ð¡](ð¥1 â ð¡) + ð â²â² [ð¡](ð¥1 â ð¡)2 + ð (3) [ð¡](ð¥1 â ð¡)3 + . . . 2 3! 1 (ð) ð ð+1 + ð [ð¡](ð¥1 â ð¡) + ðð+1 (ð¥1 â ð¡) ð! ð is diï¬erentiable on ð with 1 1 ð â² (ð¡) = ð â² [ð¡] â ð â² [ð¡] + ð â²â² [ð¡](ð¥1 â ð¡) â ð â²â² [ð¡](ð¥1 â ð¡) + ð (3) [ð¡](ð¥1 â ð¡)2 â ð (3) [ð¡](ð¥1 â ð¥0 )2 + . . . 2 2 1 1 (ð+1) (ð) ðâ1 ð ð [ð¡](ð¥1 â ð¡) + + ð [ð¡](ð¥1 â ð¡) â (ð + 1)ðð+1 (ð¥1 â ð¡)ð (ð â 1)! ð! All but the last two terms cancel, so that 1 ð (ð¡) = ð (ð+1) [ð¡](ð¥1 â ð¡)ð â (ð + 1)ðð+1 (ð¥1 â ð¡)ð = ð! â²
223
(
) 1 (ð+1) ð [ð¡] â (ð + 1)ðð+1 (ð¥1 â ð¡)ð ð!
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Note that ð(ð¥1 ) = ð (ð¥1 ) and 1 1 ð(ð¥0 ) = ð (ð¥0 ) + ð â² [ð¥0 ](ð¥1 â ð¥0 ) + ð â²â² [ð¥0 ](ð¥1 â ð¥0 )2 + ð (3) [ð¥0 ](ð¥1 â ð¥0 )3 + . . . 2 3! 1 + ð (ð+1) [ð¥0 ](ð¥1 â ð¥0 )ð + ðð+1 (ð¥1 â ð¥0 )ð+1 (4.51) ð! is a polynomial approximation for ð near ð¥0 . If we require that ðð+1 be such that ð(ð¥0 ) = ð (ð¥1 ) = ð(ð¥1 ), there exists (Theorem 4.1) some 𥠯 between ð¥0 and ð¥1 such that ð¥)(ð¥1 â ð¥0 ) = 0 ð(ð¥1 ) â ð(ð¥0 ) = ð â² (¯ which for ð¥1 â= ð¥0 implies that ð â² (¯ ð¥) =
1 ð+1 [¯ ð¥] â (ð + 1)ðð+1 = 0 ð ð!
or ðð+1 =
1 ð ð+1 [¯ ð¥] (ð + 1)!
Setting ð¥ = ð¥1 â ð¥0 in (4.51) gives the required result. 4.58 By Taylorâs theorem (Exercise 4.57), for every ð¥ â ð â ð¥0 , there exists 𥠯 between 0 and ð¥ such that 1 ð (ð¥0 + ð¥) = ð (ð¥0 ) + ð â² [ð¥0 ]ð¥ + ð â²â² [ð¥0 ]ð¥2 + ð(ð¥) 2 where ð(ð¥) =
1 (3) ð [¯ ð¥]ð¥3 3!
and 1 ð(ð¥) = ð (3) [¯ ð¥](ð¥) ð¥2 3! ð¥] is bounded on [0, ð¥] and therefore Since ð â ð¶ 3 , ð (3) [¯ lim â£
ð¥â0
ð(ð¥) 1 ⣠= lim â£ð (3) [¯ ð¥](ð¥)⣠= 0 ð¥â0 3! ð¥2
4.59 The function ð : â â ð deï¬ned by ð(ð¡) = ð¡x0 + (1 â ð¡)x ð is ð¶ â with ð·ð[ð¡] = x and ð·ð ð(ð¡) = 0 for ð = 2, 3, . . . . By Exercise 4.51, the composite function â = ð â ð is ð¶ ð+1 . By the Chain rule ââ² (ð¡) = ð·ð [ð(ð¡)] â ð·ð[ð¡] = ð·ð [ð(ð¡)](x) Similarly
( ) ââ²â² (ð¡) = ð· ð·ð [ð(ð¡)](x) = ð·2 ð [ð(ð¡)] â ð·ð[ð¡](x â x0 ) = ð·2 ð [ð(ð¡)](x)(2) 224
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics and for all 1 †ð †ð + 1
) ( â(ð) (ð¡) = ð· ð·(ðâ1) ð [ð(ð¡)](x)(ðâ1) = ð·ð ð [ð(ð¡)] â ð·ð[ð¡](x â x0 )(ðâ1) = ð·ð ð [ð(ð¡)](x)(ð)
4.60 From Exercise 4.54, the Hessian of ð is ( ð ð»(x) = 2 ð
ð ð
)
and the gradient of ð is
( ) âð (x) = (2ðð¥1 , 2ðð¥2 ) with âð (0, 0) = 0
so that the second order Taylor series at (0, 0) is 1 ð (x) = ð (0, 0) + âð (0, 0)x + 2xð 2
(
ð ð ð ð
) x
= ðð¥21 + 2ðð¥1 ð¥2 + ðð¥22 Not surprisingly, we conclude that the best quadratic approximation of a quadratic function is the function itself. 4.61
1. Since ð·ð [x0 ] is continuous and one-to-one (Exercise 3.36), there exists a constant ð such that ð â¥x1 â x2 ⥠†â¥ð·ð [x0 ](x1 â x2 )â¥
(4.52)
Let ð = ð/2. By Exercise 4.43, there exists a neighborhood ð such that â¥ð·ð [x0 ](x1 â x2 ) â (ð (x1 ) â ð (x2 ))⥠= â¥ð (x1 ) â ð (x2 ) â ð·ð [x0 ](x1 â x2 )⥠†ð â¥x1 â x2 ⥠for every x1 , x2 â ð. The Triangle Inequality (Exercise 1.200) implies â¥ð·ð [x0 ](x1 â x2 )⥠â â¥(ð (x1 ) â ð (x2 ))⥠†ð â¥x1 â x2 ⥠Substituting (4.52) 2ð â¥x1 â x2 ⥠â â¥(ð (x1 ) â ð (x2 ))⥠†ð â¥x1 â x2 ⥠That is ð â¥x1 â x2 ⥠†â¥(ð (x1 ) â ð (x2 ))â¥
(4.53)
and therefore ð (x1 ) = ð (x2 ) =â x1 = x2 2. Let ð = ð (ð). Since the restriction of ð to ð is one-to-one and onto, and therefore there exists an inverse ð â1 : ð â ð. For any y1 , y2 â ð , let x1 = ð â1 (y1 ) and x2 = ð â1 (y2 ). Substituting in (4.53) ð ð â1 (y1 ) â ð â1 (y2 ) †â¥y1 â y2 ⥠so that â1 ð (y1 ) â ð â1 (y2 ) †1 â¥y1 â y2 ⥠ð ð â1 is continuous. 225
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
3. Since ð is open, ð = ð â1 (ð) is open. Therefore, ð = ð (ð) is a neighborhood of ð (x0 ). Therefore, ð is locally onto. 4.62 Assume to the contrary that there exists x0 â= x1 â ð with ð (x0 ) = ð (x1 ). Let x = x1 â x0 . Deï¬ne ð : [0, 1] â ð by ð(ð¡) = (1 â ð¡)x0 + ð¡x1 = x0 + ð¡x. Then ð(0) = x0 Deï¬ne
ð(1) = x1
ð â² (ð¡) = x
( ( ) ) â(ð¡) = xð ð ð(ð¡) â ð (x0 )
Then â(0) = 0 = â(1) By the mean value theorem (Mean value theorem), there exists 0 < ðŒ < 1 such that ð(ðŒ) â ð and ââ² (ðŒ) = xð ð·ð [ð(ðŒ)]x = xð ðœð (ð(ðŒ))x = 0 which contradicts the deï¬niteness of ðœð . 4.63 Substituting the linear functions in (4.35) and (4.35), the IS-LM model can be expressed as (1 â ð¶ðŠ )ðŠ â ðŒð ð = ð¶0 + ðŒ0 + ðº â ð¶ðŠ ð ð¿ðŠ ðŠ + ð¿ð ð = ð/ð which can be rewritten in matrix form as )( ) ( ) ( ðŠ ð â ð¶ðŠ ð 1 â ð¶ðŠ ðŒð = ð¿ðŠ ð¿ð ð ð/ð where ð = ð¶0 + ðŒ0 + ðº. Provided the system is nonsingular, that is 1 â ð¶ðŠ ðŒð â= 0 ð·= ð¿ðŠ ð¿ð the system can be solved using Cramerâs rule (Exercise 3.103) to yield (1 â ð¶ðŠ )ð/ð â ð¿ðŠ (ð â ð¶ðŠ ð ) ð· ð¿ð (ð â ð¶ðŠ )ð â ðŒð ð/ð ðŠ= ð· ð=
4.64 The kernel kernel ð·ð¹ [(x0 , ðœ0 )] = { (x, ðœ) : ð·ð¹ [(x0 , ðœ0 )](x, ðœ) = 0 } is the set of solutions to the equation ( ) ( ) ( ) x ð·x ð (x0 , ðœ0 )x + ð·ðœ ð (x0 , ðœ0 )ðœ 0 = ð·ð¹ [x0 , ðœ0 ] = ðœ ðœ 0 Only ðœ = 0 satisï¬es this equation. Substituting ðœ = 0, the equation reduces to ð·x ð (x0 , ðœ0 )x = 0 which has a unique solution x = 0 since ð·x ð [x0 , ðœ0 ] is nonsingular. Therefore the kernel of ð·ð¹ [x0 , ðœ0 ] consists of the single point (0, 0) which implies that ð·ð¹ [x0 , ðœ 0 ] is nonsingular (Exercise 3.19). 226
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.65 The IS curve is horizontal if its slope is zero, that is ð·ðŠ ð = â
1 â ð·ðŠ ð¶ âð·ð ðŒ
This requires either 1. unit marginal propensity to consume (ð·ðŠ ð¶ = 1) 2. inï¬nite interest elasticity of investment (ð·ð ðŒ = â) 4.66 The LM curve ð = â(ðŠ) is implicitly deï¬ned by the equation ð (ð, ðŠ; ðº, ð, ð ) = ð¿(ðŠ, ð) â ð/ð = 0 the slope of which is given by ð·ðŠ ð ð·ð ð ð·ðŠ ð¿ =â ð·ð ð¿
ð·ðŠ â = â
Economic considerations dictate that the numerator (ð·ðŠ ð ) is positive while the denominator (ð·ð ð¿) is negative. Preceded by a negative sign, the slope of the LM curve is positive. The LM curve would be vertical (inï¬nite slope) if the interest elasticity of the demand for money was zero (ð·ð ð¿ = 0). 4.67 Suppose ð is convex. For any x, x0 â ð let ) ( â(ð¡) = ð ð¡x + (1 â ð¡)x0 †ð¡ð (x) + (1 â ð¡)ð (x0 ) for 0 < ð¡ < 1. Subtracting â(0) = ð (x0 ) â(ð¡) â â(0) †ð¡ð (x) â ð¡ð (x0 ) and therefore ð (x) â ð (x0 ) â¥
â(ð¡) â â(0) ð¡
Using Exercise 4.10 ð (x) â ð (x0 ) ⥠lim
ð¡â0
â(ð¡) â â(0) â x ð [x0 ] = ð·ð [x0 ](x â x0 ) =ð· ð¡
Conversely, let x0 = ðŒx1 + (1 â ðŒ)x2 for any x1 , x2 â ð. If ð satisï¬es (4.29) on ð, then ð (x1 ) ⥠ð (x0 ) + ð·ð [x0 ](x1 â x0 ) ð (x2 ) ⥠ð (x0 ) + ð·ð [x0 ](x2 â x0 ) and therefore for any 0 †ðŒ †1 ðŒð (x1 ) ⥠ðŒð (x0 + ðŒð·ð [x0 ](x1 â x0 ) (1 â ðŒ)ð (x2 ) ⥠(1 â ðŒ)ð (x0 + (1 â ðŒ)ð·ð [x0 ](x2 â x0 ) Adding and using the linearity of ð·ð (Exercise 4.21) ðŒð (x1 ) + (1 â ðŒ)ð (x2 ) ⥠ð (x0 ) + ð·ð [x0 ](ðŒx1 + (1 â ðŒ)x2 â x0 ) = ð (x0 ) = ð (ðŒx1 + (1 â ðŒ)x2 ) That is, ð is convex. If (4.29) is strict, so is (4.54). 227
(4.54)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
4.68 Since â is convex, it has a subgradient ð â ð â (Exercise 3.181) such that â(x) ⥠â(x0 ) + ð(x â x0 ) for every x â ð (4.31) implies that ð is also a subgradient of ð on ð ð (x) ⥠ð (ð¥0 ) + ð(x â x0 ) for every x â ð Since ð is diï¬erentiable, this implies that ð is unique (Remark 4.14) and equal to the derivative of ð . Hence â is diï¬erentiable at x0 with ð·â[x0 ] = ð·ð [x0 ]. 4.69 Assume ð is convex. For every x, x0 â ð, Exercise 4.67 implies ( ) ð (x) ⥠ð (x0 ) + âð (x0 )ð x â x0 ) ( ð (x0 ) ⥠ð (x) + âð (x)ð x0 â x Adding ) ( ( ) ð (x) + ð (x0 ) ⥠ð (x) + ð (x0 ) + âð (x)ð x0 â x + âð (x0 )ð x â x0 or ( ) ( ) âð (x)ð x â x0 ⥠âð (x0 )ð x â x0 and therefore âð (x) â âð (x0 )ð x â x0 ⥠0 When ð is strictly convex, the inequalities are strict. Conversely, assume (4.32). By the mean value theorem (Theorem 4.1), there exists ¯ â (x, x0 ) such that x x)ð x â x0 ð (x) â ð (x0 ) = âð (¯ By assumption ¯ â x0 ⥠0 âð (¯ x) â âð (x0 )ð x But ¯ â x0 = ðŒx0 + (1 â ðŒ)x â x0 = (1 â ðŒ)(x â x0 ) x and therefore (1 â ðŒ)âð (¯ x) â âð (x0 )ð x â x0 ⥠0 so that âð (¯ x)ð x â x0 ⥠âð (x0 )ð x â x0 ⥠0 and therefore ð (x) â ð (x0 ) = âð (¯ x)ð x â x0 ⥠âð (x0 )ð x â x0 Therefore ð is convex by Exercise 4.67. 228
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 4.70 For ð â â, âð (ð¥) = ð â² (ð¥) and (4.32) becomes (ð â² (ð¥2 ) â ð â² (ð¥1 )(ð¥2 â ð¥1 ) ⥠0 for every ð¥1 , ð¥2 â ð. This is equivalent to ð â² (ð¥2 )(ð¥2 â ð¥1 ) ⥠ð â² (ð¥1 )(ð¥2 â ð¥1 )0 or ð¥2 > ð¥1 =â ð â² (ð¥2 ) ⥠ð â² (ð¥1 ) ð is strictly convex if and only if the inequalities are strict.
4.71 ð â² is increasing if and only if ð â²â² = ð·ð Ⲡ⥠0 (Exercise 4.35). ð â² is strictly increasing if ð â²â² = ð·ð â² > 0 (Exercise 4.36). 4.72 Adapting the previous example
⧠ âš= 0 ð â²â² (ð¥) = ð(ð â 1)ð¥ð â 2 = ⥠0  â© indeterminate
if ð = 1 if ð = 2, 4, 6, ððð¡ð otherwise
Therefore, the power function is convex if ð is even, and neither convex if ð ⥠3 is odd. It is both convex and concave when ð = 1. 4.73 Assume ð is quasiconcave, and ð (x) ⥠ð (x0 ). Diï¬erentiability at x0 implies for all 0 < ð¡ < 1 ð (x0 + ð¡(x â x0 ) = ð (x0 ) + âð (x0 )ð¡(x â x0 ) + ð(ð¡) â¥ð¡(x â x0 )⥠where ð(ð¡) â 0 and ð¡ â 0. Quasiconcavity implies ð (x0 + ð¡(x â x0 ) ⥠ð (x0 ) and therefore âð (x0 )ð¡(x â x0 ) + ð(ð¡) â¥ð¡(x â x0 )⥠⥠0 Dividing by ð¡ and letting ð¡ â 0, we get âð (x0 )(x â x0 ) ⥠0 Conversely, assume ð is a diï¬erentiable functional satisfying (4.36). For any x1 , x2 â ð with ð (x1 ) ⥠ð (x2 for every x, x0 â ð), deï¬ne â : [0, 1] â â by ( ) ) ( â(ð¡) = ð (1 â ð¡)x1 + ð¡x2 = ð x1 + ð¡(x2 â x1 ) We need to show that â(ð¡) ⥠â(1) for every ð¡ â (0, 1). Suppose to the contrary that â(ð¡1 ) < â(1). Then (see below) there exists ð¡0 with â(ð¡0 ) < â(1) and ââ² (ð¡0 ) < 0. By the Chain Rule, this implies ââ² (ð¡0 ) = âð (x0 )(x2 â x1 ) < 0 critical where x0 = x1 + ð¡(x2 â x1 ). Since x2 â x0 = (1 â ð¡)(x2 â x1 ) this implies that ââ² (ð¡0 ) =
1 âð (x0 )(x2 â x0 ) 1âð¡ 229
(4.55)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
On the other hand, since ð (x0 ) ⥠ð (x2 ), (4.36) implies âð (x0 )(x2 â x0 ) ⥠0 contradicting (4.55). To show that there exists ð¡0 with â(ð¡0 ) < â(1) and ââ² (ð¡0 ) < 0: Since ð is continuous, there exists an open interval (ð, ð) with ð < ð¡1 < ð with â(ð) = â(ð) = â(1) and â(ð¡) < â(1) for every ð¡ â (ð, ð). By the Mean Value Theorem, there exist ð¡0 â (ð, ð¡1 ) such that 0 < â(ð¡1 ) â â(ð) = ââ² (ð¡0 )(ð¡1 â ð) which implies that ââ² (ð¡0 ) > 0. 4.74 Suppose to the contrary that ð (x) > ð (x0 ) and âð (x0 )(x â x0 ) †0 critical Let x1 = ââð (x0 ) â= 0. For every ð¡ â â+ âð (x0 )(x + ð¡x1 â x0 ) = âð (x0 )ð¡x1 + âð (x0 )(x â x0 ) †ð¡âð (x0 )x1 2
= âð¡ â¥âð (x0 )⥠< 0 Since ð is continuous, there exists ð¡ > 0 such that ð (x + ð¡x1 ) > ð (x0 ) and âð (x0 )(x + ð¡x1 â x0 ) < 0 contradicting the quasiconcavity of ð (4.36). 4.75 Suppose ð (x) < ð (x0 ) =â âð (x0 )(x â x0 ) < 0 This implies that âð (x) > âð (x0 ) =â â â ð (x0 )(x â x0 ) > 0 and âð is pseudoconcave. 4.76
1. If ð â ð¹ [ð] is concave (and diï¬erentiable) ð (x) †ð (x0 ) + âð (x0 )ð (x â x0 ) for every x, x0 â ð(equation 4.30). Therefore ð (x) > ð (x0 ) =â âð (x0 )ð (x â x0 ) > 0 ð is pseudoconcave.
2. Assume to the contrary that ð is pseudoconcave but not quasiconcave. Then, ¯ = ðŒx1 + (1 â ðŒ)x2 , x1 , x2 â ð such that there exists x ð (¯ x) < min{ð (x1 ), ð (x2 )} Assume without loss of generality that ð (¯ x) < ð (x1 ) †ð (x2 ) 230
(4.56)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Pseudoconcavity (4.38) implies ¯) > 0 âð (¯ x)(x2 â x
(4.57)
x â (1 â ðŒ)x2 )/ðŒ Since x1 = (¯ ¯= x1 â x
) 1( 1âðŒ ¯ â (1 â ðŒ)x2 â ðŒÂ¯ ¯) x (x2 â x x =â ðŒ ðŒ
Substituting in (4.57) gives ¯) < 0 âð (¯ x)(x1 â x which by pseudoconcavity implies ð (x1 ) †ð (¯ x) contradicting our assumption (4.56) . 3. Exercise 4.74. 4.77 The CES function is quasiconcave provided ð †1 (Exercise 3.58). Since ð·ð¥ð ð (x) > 0 for all x â âð+ +, the CES function with ð †1 is pseudoconcave on âð++ . 4.78 Assume that ð : ð â â is homogeneous of degree ð, so that for every x â ð ð (ð¡x) = ð¡ð ð (x) for every ð¡ > 0 Diï¬erentiating both sides of this identity with respect to ð¥ð ð·ð¥ð ð (ð¡x)ð¡ = ð¡ð ð·ð¥ð ð (x) and dividing by ð¡ > 0 ð·ð¥ð ð (ð¡x) = ð¡ðâ1 ð·ð¥ð ð (x) 4.79 If ð is homogeneous of degree ð â x ð (x) = lim ð (x + ð¡x) â ð (x) ð· ð¡â0 ð¡ ð ((1 + ð¡)x) â ð (x) = lim ð¡â0 ð¡ (1 + ð¡)ð ð (x) â ð (x) = lim ð¡â0 ð¡ (1 + ð¡)ð â 1 = lim ð (x) ð¡â0 ð¡ Applying LâHËopitalâs Rule (Exercise 4.47) (1 + ð¡)ðâ1 ð(1 + ð¡)ðâ1 ð (x) = lim =ð ð¡â0 ð¡â0 ð¡ 1 lim
and therefore â x ð (x) = ðð (x) ð·
(4.58)
4.80 For ï¬xed x, deï¬ne â(ð¡) = ð (ð¡x) By the Chain Rule ââ² (ð¡) = ð¡ð·ð [ð¡x](x) = ð¡ðð (ð¡x) = ð¡ðâ(ð¡) 231
(4.59)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
using(4.40). Diï¬erentiating the product â(ð¡)ð¡âð ( ) ( ) ð·ð¡ â(ð¡)ð¡âð = âðâ(ð¡)ð¡âðâ1 + ð¡âð ââ² (ð¡) = ð¡âð ââ² (ð¡) â ðð¡â(ð¡) = 0 from (4.59). Since this holds for every ð¡, â(ð¡)ð¡âð must be constant (Exercise 4.38), that is there exists ð â â such that â(ð¡)ð¡âð = ð =â â(ð¡) = ðð¡ð Evaluating at ð¡ = 1, â(1) = ð and therefore â(ð¡) = ð¡ð â(1) Since â(ð¡) = ð (ð¡x) and â(1) = ð (x), this implies ð (ð¡x) = ð¡ð ð (x) for every x and ð¡ > 0 ð is homogeneous of degree ð. 4.81 If ð is linearly homogeneous and quasiconcave, then ð is concave (Proposition 3.12). Therefore, its Hessian is nonpositive deï¬nite (Proposition 4.1). and its diagonal elements ð·ð¥2 ð ð¥ð ð (x) are nonpositive (Exercise 3.95). By Wicksellâs law, ð·ð¥2 ð ð¥ð ð (x) is nonnegative. 4.82 Assume ð is homogeneous of degree ð, that is ð (ð¡x) = ð¡ð ð (x) for every x â ð and ð¡ > 0 By Eulerâs theorem ð·ð¡ ð [ð¡x](ð¡x) = ðð (ð¡x) and therefore the elasticity of scale is
ð¡ ð¡ ð·ð¡ ð (ð¡x) ðð (ð¡x) = ð = ðž(x) = ð (ð¡x) ð (ð¡x) ð¡=1
Conversely, assume that
ð¡ ðž(x) = ð·ð¡ ð (ð¡x) =ð ð (ð¡x) ð¡=1
that is ð·ð¡ ð (ð¡x) = ðð (ð¡x) By Eulerâs theorem, ð is homogeneous of degree ð. 4.83 Assume ð â ð¹ (ð) is diï¬erentiable and homogeneous of degree ð â= 0. By Eulerâs theorem ð·ð [x](x) = ðð (x) â= 0 for every x â ð such that ð (x) â= 0. 4.84 ð satisï¬es Eulerâs theorem ðð (x) =
ð â
ð·ð ð (x)ð¥ð
ð=1
232
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Diï¬erentiating with respect to ð¥ð ðð·ð ð (x) =
ð â
ð·ðð ð (x)ð¥ð + ð·ð ð (x)
ð=1
or (ð â 1)ð·ð ð (x) =
ð â
ð·ðð ð (x)ð¥ð
ð = 1, 2, . . . , ð
ð=1
Multiplying each equation by ð¥ð and summing (ð â 1)
ð â
ð·ð ð (x)ð¥ð =
ð=1
ð â ð â
ð·ðð ð (x)ð¥ð ð¥ð = xâ² ð»x
ð=1 ð=1
By Eulerâs theorem, the left hand side is (ð â 1)ðð (x) = xâ² ð»x 4.85 If ð is homothetic, there exists strictly increasing ð and linearly homogeneous â such that ð = ð â â (Exercise 3.175). Using the Chain Rule and Exercise 4.78 ð·ð¥ð ð (ð¡x) = ð â² (ð (ð¡x))ð·ð¥ð â(ð¡x) = ð¡ð â² (ð (ð¡x))ð·ð¥ð â(x) and therefore ð·ð¥ð ð (ð¡x) ð¡ð â² (ð (ð¡x))ð·ð¥ð â(x) = ð·ð¥ð ð (ð¡x) ð·ð¥ð ð¡ð â² (ð (ð¡x))ð·ð¥ð â(x) ð·ð¥ð â(x) = ð·ð¥ð â(x) ð·ð¥ð ð â² (ð (x)ð·ð¥ð â(x) = ð·ð¥ð ð â² (ð (x))ð·ð¥ð â(x) ð·ð¥ð ð (x) = ð·ð¥ð ð (x) ð·ð¥ð
233
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 5: Optimization 5.1 As stated, this problem has no optimal solution. Revenue ð (ð¥) increases without bound as the rate of exploitation ð¥ gets smaller and smaller. Given any positive exploitation rate ð¥0 , a smaller rate will increase total revenue. Nonexistence arises from inadequacy in modeling the island leadersâ problem. For example, the model ignores any costs of extraction and sale. Realistically, we would expect per-unit costs to decrease with volume (increasing returns to scale) at least over lower outputs. Extraction and transaction costs should make vanishingly small rates of output prohibitively expensive and encourage faster utilization. Secondly, even if the government weights future generations equally with the current generation, it would be rational to value current revenue more highly than future revenue and discount future returns. Discounting is appropriate for two reasons â Current revenues can be invested to provide a future return. There is an opportunity cost (the interest foregone) to delaying extraction and sale. â Innovation may create substitutes which reduce the future demand for the fertilizer. If the government is risk averse, it has an incentive to accelerate exploitation, trading-oï¬ of lower total return against reduced risk. 5.2 Suppose that xâ is a local optimum which is not a global optimum. That is, there exists a neighborhood ð of xâ such that ð (xâ , ðœ) ⥠ð (x, ðœ) for every x â ð â© ðº(ðœ) and also another point xââ â ðº(ðœ) such that ð (xââ , ðœ) > ð (xâ , ðœ) Since ðº(ðœ) is convex, there exists ðŒ â (0, 1) such that ðŒxâ + (1 â ðŒ)xââ â ð â© ðº(ðœ) By concavity of ð ð (ðŒxâ + (1 â ðŒ)xââ , ðœ) ⥠ðŒð (xâ , ðœ) + (1 â ðŒ)ð (xââ , ðœ) > ð (xâ , ðœ) contradicting the assumption that xâ is a local optimum. 5.3 Suppose that xâ is a local optimum which is not a global optimum. That is, there exists a neighborhood ð of xâ such that ð (xâ , ðœ) ⥠ð (x, ðœ) for every x â ð â© ðº(ðœ) and also another point xââ â ðº(ðœ) such that ð (xââ , ðœ) > ð (xâ , ðœ) Since ðº(ðœ) is convex, there exists ðŒ â (0, 1) such that ðŒxâ + (1 â ðŒ)xââ â ð â© ðº(ðœ)
234
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
By strict quasiconcavity of ð ð (ðŒxâ + (1 â ðŒ)xââ , ðœ) > min{ ð (xâ , ðœ), ð (xââ , ðœ) } > ð (xâ , ðœ) contradicting the assumption that xâ is a local optimum. Therefore, if xâ is local optimum, it must be a global optimum. Now suppose that xâ is a weak global optimum, that is ð (xâ , ðœ) ⥠ð (x, ðœ) for every x â ð but there another point xââ â ð such that ð (xââ , ðœ) = ð (xâ , ðœ) Since ðº(ðœ) is convex, there exists ðŒ â (0, 1) such that ðŒxâ + (1 â ðŒ)xââ â ð â© ðº(ðœ) By strict quasiconcavity of ð ð (ðŒxâ + (1 â ðŒ)xââ , ðœ) > min{ ð (xâ , ðœ), ð (xââ , ðœ) } = ð (xâ , ðœ) contradicting the assumption that xâ is a global optimum. We conclude that every optimum is a strict global optimum and hence unique. 5.4 Suppose that xâ is a local optimum of (5.3) in ð, so that ð (xâ ) ⥠ð (x)
(5.80)
for every x in a neighborhood ð of xâ . If ð is diï¬erentiable, ð (x) = ð (xâ ) + ð·ð [xâ ](x â xâ ) + ð(x) â¥x â xâ ⥠where ð(x) â 0 as x â xâ . (5.80) implies that there exists a ball ðµð (xâ ) such that ð·ð [xâ ](x â xâ ) + ð(x) â¥x â xâ ⥠†0 for every x â ðµð (xâ ). Letting x â xâ , we conclude that ð·ð [xâ ](x â xâ ) †0 for every x â ðµð (xâ ). Suppose there exists x â ðµð (xâ ) such that ð·ð [xâ ](x â xâ ) = ðŠ < 0 Let dx = x â xâ so that x = xâ + dx. Then xâ â dx â ðµð (xâ ). Since ð·ð [xâ ] is linear, ð·ð [xâ ](âdx) = âð·ð [xâ ](dx) = âðŠ > 0 contradicting (5.80). Therefore ð·ð [xâ ](x â xâ ) = 0 for every x â ðµð (xâ ).
235
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
5.5 We apply the reasoning of Example 5.5 to each component. Formally, for each ð, let ðËð be the projection of ð along the ðð¡â axis ðËð (ð¥ð ) = ð (ð¥â1 , ð¥â2 , . . . , ð¥âðâ1 , ð¥ð , ð¥âð+1 , . . . , ð¥âð ) ð¥âð maximizes ðËð (ð¥ð ) over â+ , for which it is necessary that ð·ð¥ð ðËð (ð¥âð ) †0
ð¥âð ⥠0
ð¥âð ð·ð¥ð ðËð (ð¥âð ) = 0
Substituting ð·ð¥ð ðËð (ð¥âð ) = ð·ð¥ð ð [xâ ] yields ð·ð¥ð ð [xâ ] †0
ð¥âð ⥠0
ð¥âð ð·ð¥ð ð [xâ ] = 0
5.6 By Taylorâs Theorem (Example 4.33) 1 2 ð (xâ + dx) = ð (xâ ) + âð (xâ )dx + dxð ð»ð (xâ )dx + ð(dx) â¥dx⥠2 with ð(dx) â 0 as dx â 0. Given 1. âð (xâ ) = 0 and 2. ð»ð (xâ ) is negative deï¬nite and letting dx â 0, we conclude that ð (xâ + dx) < ð (xâ ) for small dx. xâ is a strict local maximum. 5.7 If xâ is a local minimum of ð (x), it is necessary that ð (xâ ) †ð (x) for every x in a neighborhood ð of xâ . Assuming that ð is ð¶ 2 , ð (x) can be approximated by 1 ð (x) â ð (xâ ) + âð (xâ )dx + dxð ð»ð (xâ )dx 2 where dx = x â xâ . If xâ is a local minimum, then there exists a ball ðµð (xâ ) such that 1 ð (xâ ) †ð (xâ ) + âð (xâ )dx + dxð ð»ð (xâ )dx 2 or 1 âð (xâ )dx + dxð ð»ð (xâ )dx ⥠0 2 for every dx â ðµð (xâ ). To satisfy this inequality for all small dx requires that the ï¬rst term be zero and the second term nonnegative. In other words, for a point xâ to be a local minimum of a function ð , it is necessary that the gradient be zero and the Hessian be nonnegative deï¬nite at xâ . Furthermore, by Taylorâs Theorem 1 2 ð (xâ + dx) = ð (xâ ) + âð (xâ )dx + dxð ð»ð (xâ )dx + ð(dx) â¥dx⥠2 with ð(dx) â 0 as dx â 0. Given 236
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
(1,2,3)
3 2 ð¥ 2
0
1
1 ð¥1
2
Figure 5.1: The strictly concave function ð (ð¥1 , ð¥2 ) = ð¥1 ð¥2 + 3ð¥2 â ð¥21 â ð¥22 has a unique global maximum. 1. âð (xâ ) = 0 and 2. ð»ð (xâ ) is positive deï¬nite and letting dx â 0, we conclude that ð (xâ + dx) > ð (xâ ) for small dx. xâ is a strict local minimum. 5.8 By the Weierstrass theorem (Theorem 2.2), ð has a maximum ð¥â and a minimum ð¥â on [ð, ð]. Either â ð¥â â (ð, ð) and ð â² (ð¥â ) = 0 (Theorem 5.1) or â ð¥â â (ð, ð) and ð â² (ð¥â ) = 0 (Exercise 5.7) or â Both maxima and minima are boundary points, that is ð¥â , ð¥â â {ð, ð} which implies that ð is constant on [ð, ð] and therefore ð â² (ð¥) = 0 for every ð¥ â (ð, ð) (Exercise 4.7). 5.9 The ï¬rst-order conditions for a maximum are ð·ð¥1 ð (ð¥1 , ð¥2 ) = ð¥2 â 2ð¥1 = 0 ð·ð¥2 ð (ð¥1 , ð¥2 ) = ð¥1 + 3 â 2ð¥2 = 0 which have the unique solution ð¥â1 = 1, ð¥â2 = 2. (1, 2) is the only stationary point of ð and hence the only possible candidate for a maximum. To verify that (1, 2) satisï¬es the second-order condition for a maximum, we compute the Hessian of ð ) ( â2 1 ð»(x) = 1 â2 which is negative deï¬nite everywhere. Therefore (1, 2) is a strict local maximum of ð . Further, since ð is strictly concave (Proposition 4.1), we conclude that (1, 2) is a strict global maximum of ð (Exercise 5.2), where it attains its maximum value ð (1, 2) = 3 (Figure 5.1). 237
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5.10 The ï¬rst-order conditions for a maximum (or minimum) are ð·1 ð (ð¥) = 2ð¥1 = 0 ð·2 ð (ð¥) = 2ð¥2 = 0 which have a unique solution ð¥1 = ð¥2 = 0. This is the only stationary point of ð . Since the Hessian of ð ( ) 2 0 ð»= 0 2 is positive deï¬nite, we deduce (0, 0) is a strict global minimum of ð (Proposition 4.1, Exercise 5.2). 5.11 The average ï¬rmâs proï¬t function is 1 1 Î (ð, ð) = ðŠ â ð â ð â 2 6 and the ï¬rmâs proï¬t maximization problem is 1 1 max Î (ð, ð) = ð 1/6 ð1/3 â ð â ð â ð,ð 2 6 A necessary condition for a proï¬t maximum is that the proï¬t function be stationary, that is 1 â5/6 1/3 1 ð ð â =0 6 2 1 1/6 â2/3 â1=0 ð·ð Î (ð, ð) = ð ð 3
ð·ð Î (ð, ð) =
which can be solved to yield ð=ð=
1 9
The ï¬rmâs output is ðŠ=
1 1/6 1 1/3 1 = 9 9 3
and its proï¬t is 1 11 1 1 1 1 â â =0 Î ( , ) = â 3 3 3 29 9 6 5.12 By the Chain Rule ð·x (â â ð )[xâ ] = ð·â â ð·x ð [xâ ] = 0 Since ð·â > 0 ð·x (â â ð )[ð¥â ] = 0 ââ ð·x ð [xâ ] = 0 â â ð has the same stationary points as ð .
238
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5.13 Since the log function is monotonic, ï¬nding the maximum likelihood estimators is equivalent to solving the maximization problem ( Exercise 5.12) max log ð¿(ð, ð) = â ð,ð
ð 1 â ð log 2ð â ð log ð â 2 (ð¥ð¡ â ð)2 2 2ð ð¡=1
For (Ë ð, ð Ë 2 ) to solve this problem, it is necessary that log ð¿ be stationary at (Ë ð, ð Ë 2 ), that is ð, ð Ë2) = ð·ð log ð¿(Ë
ð 1 â (ð¥ð¡ â ð Ë) = 0 ð Ë 2 ð¡=1
ð·ð log ð¿(Ë ð, ð Ë2) = â
ð 1 â ð + 3 (ð¥ð¡ â ð Ë)2 = 0 ð Ë ð Ë ð¡=1
which can be solved to yield ð Ë=𥠯= ð Ë2 =
ð 1 â ð¥ð¡ ð ð¡=1
ð 1â (ð¥ð¡ â 𥠯)2 ð ð¡=1
5.14 The gradient of the objective function is ) ( â2(ð¥1 â 1) âð (x) = â2(ð¥2 â 1) while that of the constraint is
( âð(x) =
2ð¥1 2ð¥2
)
A necessary condition for the optimal solution is that these be proportional that is ) ( ( ) 2ð¥1 â2(ð¥1 â 1) =ð âð (ð¥) = = âð(x) â2(ð¥2 â 1) 2ð¥2 which can be solved to yield ð¥1 = ð¥2 =
1 1+ð
which includes an unknown constant of proportionality ð. However, any solution must also satisfy the constraint ( )2 1 ð(ð¥1 , ð¥2 ) = 2 =1 1+ð This can be solved for ð ð=
â 2â1
and substituted into (5.80) 1 ð¥1 = ð¥2 = â 2 239
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 5.15 The consumerâs problem is max ð¢(x) = ð¥1 + ð log ð¥2 xâ¥0
subject to ð(x) = ð¥1 + ð2 ð¥2 â ð = 0 The ï¬rst-order conditions for a (local) optimum are ð·ð¥1 ð¢(xâ ) = 1 †ð = ð·ð¥1 ð(ð¥â )
ð¥1 ⥠0
ð †ðð2 = ð·ð¥2 ð(ð¥â ) ð·ð¥2 ð¢(xâ ) = ð¥2
ð¥2 ⥠0
( ð¥2
ð¥1 (1 â ð) = 0 ) =0
ð â ðð2 ð¥2
(5.81) (5.82)
We can distinguish two cases: Case 1 ð¥1 = 0 in which case the budget constraint implies that ð¥2 = ð/ð2 . Case 2 ð¥1 > 0 In this case, (5.81) implies that ð = 1. Consequently, the ï¬rst inequality of (5.82) implies that ð¥2 > 0 and therefore the last equation implies ð¥2 = ð/ð2 with ð¥1 = ð â ð. We deduce that the consumer ï¬rst spends portion ð of her income on good 2 and the remainder on good 1. 5.16 Suppose without loss of generality that the ï¬rst ð components of yâ are strictly positive while the remaining components are zero. That is ðŠðâ > 0 ðŠðâ = 0
ð = 1, 2, . . . , ð ð = ð + 1, ð + 2, . . . , ð
(xâ , yâ ) solves the problem max ð (x) subject to g(x) = 0 ð = ð + 1, ð + 2, . . . , ð ðŠð = 0 By Theorem 5.2, there exist multipliers ð1 , ð2 , . . . , ðð and ðð+1 , ðð+2 , . . . , ðð such that ð·x ð [xâ , yâ ] = ð·y ð [xâ , yâ ] =
ð â ð=1 ð â
ðð ð·x ðð [xâ , yâ ] ðð ð·y ðð [xâ , yâ ] +
ð=1
ð â
ðð ðŠ ð
ð=ð+1
Furthermore, ðð ⥠0 for every ð so that ð·y ð [xâ , yâ ] â€
ð â
ðð ð·y ðð [xâ , yâ ]
ð=1
with ð·ðŠð ð [xâ , yâ ] =
ð â
ðð ð·ðŠð ðð [xâ , yâ ] if ðŠð > 0
ð=1
5.17 Assume that xâ = (ð¥â1 , ð¥â2 ) solves max ð (ð¥1 , ð¥2 )
ð¥1 ,ð¥2
240
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
subject to ð(ð¥1 , ð¥2 ) = 0 By the implicit function theorem, there exists a function â : â â â such that ð¥1 = â(ð¥2 )
(5.83)
and ð(â(ð¥2 ), ð¥2 ) = 0 for ð¥2 in a neighborhood of ð¥â2 . Furthermore ð·â[ð¥â2 ] = â
ð·ð¥1 ð[xâ ] ð·ð¥2 ð[xâ ]
(5.84)
Using (5.41), we can convert the original problem into the unconstrained maximization of a function of a single variable max ð (â(ð¥2 ), ð¥2 ) ð¥2
If ð¥â2 maximizes this function, it must satisfy the ï¬rst-order condition (applying the Chain Rule) ð·ð¥1 ð [ð¥â] â ð·â[ð¥â2 ] + ð·ð¥2 ð [xâ ] = 0 Substituting (5.42) yields ( ) ð·ð¥1 ð[xâ ] ð·ð¥1 ð [ð¥â] â + ð·ð¥2 ð [xâ ] = 0 ð·ð¥2 ð[xâ ] or ð·ð¥1 ð [ð¥â] ð·ð¥1 ð[xâ ] = ð·ð¥2 ð [xâ ] ð·ð¥2 ð[xâ ] 5.18 The consumerâs problem is max ð¢(x) xâð
subject to pð x = ð Solving for ð¥1 from the budget constraint yields âð ð â ð=2 ðð ð¥ð ð¥1 = ð1 Substituting this in the utility function, the aï¬ordable utility levels are â ( ) ð â ðð=2 ðð ð¥ð , ð¥2 , ð¥3 , . . . , ð¥ð ð¢ Ë(ð¥2 , ð¥3 , . . . , ð¥ð ) = ð¢ ð1
(5.85)
and the consumerâs problem is to choose (ð¥2 , ð¥3 , . . . , ð¥ð ) to maximize (5.85). The ï¬rst-order conditions are that ð¢ Ë(ð¥2 , ð¥3 , . . . , ð¥ð ) be stationary, that is for every good ð = 2, 3, . . . , ð âð ( ) ð â ð=2 ðð ð¥ð â ð·ð¥ð ð¢ Ë(ð¥2 , ð¥3 , . . . , ð¥ð ) = ð·ð¥1 ð¢(x )ð·ð¥ð + ð·ð¥ð ð¢(xâ ) = 0 ð1 241
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
which reduces to ð·ð¥1 ð¢(xâ )(â
ð1 ) + ð·ð¥ð ð¢(xâ ) = 0 ðð
or ð1 ð·ð¥1 ð¢(xâ ) = â ð·ð¥ð ð¢(x ) ðð
ð = 2, 3, . . . , ð
This is the familiar equality between the marginal rate of substitution and the price ratio (Example 5.15). Since our selection of ð¥1 was arbitrary, this applies between any two goods. 5.19 Adapt Exercise 5.6. 5.20 Corollary 5.1.2 implies that xâ is a global maximum of ð¿(x, ð), that is ð¿(xâ , ð) ⥠ð¿(x, ð) for every x â ð which implies ð (xâ ) â
â
ðð ðð (xâ ) ⥠ð (x) â
â
ðð ðð (x) for every x â ð
Since g(xâ ) = 0 this implies ð (xâ ) ⥠ð (x) â
â
ðð ðð (x) for every x â ð
A fortiori ð (xâ ) ⥠ð (x) for every x â ðº = { x â ð : g(x) = 0 } 5.21 Suppose that xâ is a local maximum of ð on ðº. That is, there exists a neighborhood ð such that ð (xâ ) ⥠ð (x) for every x â ð â© ðº But for every x â ðº, ðð (x) = 0 for every ð and â ð¿(x) = ð (x) + ðð ðð (x) = ð (x) and therefore ð¿(xâ ) ⥠ð¿(x) for every x â ð â© ðº 5.22 The area of the base is Base = ð€2 = ðŽ/3 and the four sides Sides = 4ð€â â â ðŽ ðŽ =4 3 12 4ðŽ = 16 2ðŽ = 3 242
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5.23 Let the dimensions of the vat be ð€ à ð à â. We wish to min Surface area = ðŽ = ð€ à ð + 2ð€â + 2ðâ
ð€,ð,â
subject to ð€ à ð à â = 32 The Lagrangean is ð¿(ð€, ð, â, ð) = ð€ð + 2ð€â + 2ðâ â ðð€ðâ. The ï¬rst-order conditions for a maximum are ð·ð€ ð¿ = ð + 2â â ððâ = 0
(5.86)
ð·ð ð¿ = ð€ + 2â â ðð€â = 0
(5.87)
ð·â ð¿ = 2ð€ + 2ð â ðð€ð = 0 ð€ðâ = 32
(5.88)
Subtracting (5.45) from (5.44) ð â ð€ = ð(ð â ð€)â This equation has two possible solutions. Either ð=
1 or ð = ð€ â
But if ð = 1/â, (5.44) implies that ð = 0 and the volume is zero. Therefore, we conclude that ð€ = ð. Substituting ð€ = ð in (5.45) and (5.46) gives ð€ + 2â = ðð€â 4ð€ = ðð€2 from which we deduce that ð=
4 ð€
Substituting in (5.45) ð€ + 2â =
4 ð€â = 4â ð€
which implies that ð€ = 2â or â =
1 ð€ 2
To achieve the required volume of 32 cubic metres requires that 1 ð€ à ð à â = ð€ à ð€ à ð€ = 32 2 so that the dimensions of the vat are ð€=4
ð=4
â=2
The area of sheet metal required is ðŽ = ð€ð + 2ð€â + 2ðâ = 48 243
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics 5.24 The Lagrangean for this problem is
ð¿(x, ð) = ð¥21 + ð¥22 + ð¥23 â ð(2ð¥1 â 3ð¥2 + 5ð¥3 â 19) A necessary condition for xâ to solve the problem is that the Lagrangean be stationary at xâ , that is ð·ð¥1 ð¿ = 2ð¥â1 â 2ð = 0 ð·ð¥2 ð¿ = 2ð¥â2 + 3ð = 0 ð·ð¥3 ð¿ = 2ð¥â3 â 5ð = 0
which implies 3 ð¥â2 = â ð 2
ð¥â1 = ð
ð¥â2 =
5 ð 2
(5.89)
It is also necessary that the solution satisfy the constraint, that is 2ð¥â1 â 3ð¥â2 + 5ð¥â3 = 19 Substituting (5.89) into the constraint we get 9 25 2ð + ð + ð = 19ð = 19 2 2 which implies ð = 1. Substituting in (5.89), the solution is xâ = (1, â 32 , 52 ). Since the constraint is aï¬ne and the objective (âð ) is concave, stationarity of the Lagrangean is also suï¬cient for global optimum (Corollary 5.2.4). 5.25 The Lagrangean is 1âðŒ ð¿(ð¥1 , ð¥2 , ð) = ð¥ðŒ â ð(ð1 ð¥1 + ð2 ð¥2 â ð) 1 ð¥2
The Lagrangean is stationary where ð·ð¥1 ð¿ = ðŒð¥ðŒâ1 ð¥1âðŒ â ðð1 = 0 1 2 1âðŒâ1 ð·ð¥2 ð¿ = 1 â ðŒð¥ðŒ â ðð2 = 0 1 ð¥2
Therefore the ï¬rst-order conditions for a maximum are 1â
ðŒð¥ðŒâ1 ð¥1âðŒ = ðð1 1 2
(5.90)
1âðŒâ1 ðŒð¥ðŒ 1 ð¥2
(5.91)
= ðð2 ð1 ð¥1 + ð2 ð¥2 â ð
Dividing (5.48) by (5.49) gives ðŒð¥ðŒâ1 ð¥1âðŒ 1 2
(1âðŒ)â1
1 â ðŒð¥ðŒ 1 ð¥2
= ð1 ð2
which simpliï¬es to ð1 ðŒð¥2 = (1 â ðŒ)ð¥1 ð2
or
ð2 ð¥2 =
(1 â ðŒ) ð1 ð¥1 ðŒ
Substituting in the budget constraint (5.50) (1 â ðŒ) ð1 ð¥1 = ð ðŒ ðŒ + (1 â ðŒ) ð1 ð¥1 = ð ðŒ
ð1 ð¥1 +
244
(5.92)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics so that ð¥â1 =
ð ðŒ ðŒ + (1 â ðŒ) ð1
From the budget constraint (5.92) ð¥â2 =
(1 â ðŒ) ð ðŒ + (1 â ðŒ) ð2
5.26 The Lagrangean is ðŒð 1 ðŒ2 ð¿(x, ð) = ð¥ðŒ 1 ð¥2 . . . ð¥ð â ð(ð1 ð¥1 + ð2 ð¥2 + ... + ðð ð¥ð )
The ï¬rst-order conditions for a maximum are ðŒð â1 1 ðŒ2 ð ð·ð¥ð ð¿ = ðŒð ð¥ðŒ . . . ð¥ðŒ ð â ððð = 1 ð¥2 . . . ð¥ð
ðŒð ð¢(ð¥) â ððð = 0 ð¥ð
or ðŒð ð¢(ð¥) = ðð ð¥ð ð
ð = 1, 2, . . . , ð
Summing over all goods and using the budget constraint ð â ðŒð ð¢(ð¥) ð=1
Letting
âð
ð=1
ð
ð
=
ð
â ð¢(ð¥) â ðŒð = ðð ð¥ð = ð ð ð=1 ð=1
ðŒð = ðŒ, this implies ð ð¢(x) = ð ðŒ
Substituting in (5.93) ðð ð¥ð =
ðŒð ð ðŒ
or ð¥âð =
ðŒð ð ðŒ ðð
ð = 1, 2, . . . , ð
5.27 The Lagrangean is ð¿(x, ð) = ð€1 ð¥1 + ð€2 ð¥2 â ð(ð¥ð1 + ð¥ð2 â ðŠ ð ). The necessary conditions for stationarity are ð·ð¥1 ð¿(x, ð) = ð€1 â ððð¥ðâ1 =0 1 ð·ð¥2 ð¿(x, ð) = ð€2 â ððð¥ðâ1 =0 2 or ð€1 = ððð¥ðâ1 1 ð€2 = ððð¥ðâ1 2
245
(5.93)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics which reduce to ð¥ðâ1 ð€1 = 1ðâ1 ð€2 ð¥ 2 ð€2 ðâ1 ðâ1 ð¥2 = ð¥ ð€1 1 ð ( ) ðâ1 ð€2 ð¥ð2 = ð¥ð1 ð€1 Substituting in the production constraint (
) ð ð€2 ðâ1 ð + ð¥1 = ðŠ ð ð€1 ( ð ) ( ) ðâ1 ð€2 1+ ð¥ð1 = ðŠ ð ð€1 ð¥ð1
we can solve ð¥1 ( ð¥ð1
=
(
1+ (
ð¥1 =
(
1+
ð€2 ð€1 ð€2 ð€1
ð )â1 ) ðâ1
ðŠð
ð )â1/ð ) ðâ1
ðŠ
Similarly ( ð¥2 =
( 1+
ð€1 ð€2
ð )â1/ð ) ðâ1
ðŠ
5.28 Example 5.27 is ï¬awed. The optimum of the constrained maximization problem (â = ð€/2) is in fact a saddle point of the Lagrangean. It maximizes the Lagrangean in the feasible set, but not globally. The net beneï¬t approach to the Lagrange multiplier method is really only applicable when the Lagrangean (net beneï¬t function) is concave, so that every stationary point is a global maximum. This requirement is satisï¬ed in many standard examples, such as the consumerâs problem (Example 5.21) and cost minimization (Example 5.28). It is also met in Example 5.29. The requirement of concavity is not recognized in the text, and Section 5.3.6 should be amended accordingly. 5.29 The Lagrangean ð¿(x, ð) =
ð â
( ðð (ð¥ð ) + ð ð· â
ð=1
ð â
) ð¥ð
(5.94)
ð=1
can be rewritten as ð¿(x, ð) = â
ð â ) ( ðð¥ð â ðð (ð¥ð ) + ðð·
(5.95)
ð=1
The ðth term in the sum is the net proï¬t of plant ð if its output is valued at ð. Therefore, if the company undertakes to buy electricity from its plants at the price ð and instructs each plant manager to produce so as to maximize the plantâs net proï¬t, each manager 246
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
will be induced to choose an output level which maximizes the proï¬t of the company as a whole. This is the case whether the price ð is the market price at which the company can buy electricity from external suppliers or the shadow price determined by the need to satisfy the total demand ð·. In this way, the shadow price ð can be used to decentralize the production decision. 5.30 The Lagrangean for this problem is ð¿(ð¥1 , ð¥2 , ð1 , ð2 ) = ð¥1 ð¥2 â ð1 (ð¥21 + 2ð¥22 â 3) â ð2 (2ð¥21 + ð¥22 â 3) The ï¬rst-order conditions for stationarity ð·ð¥1 ð¿ = ð¥2 â 2ð1 ð¥1 â 4ð2 ð¥1 = 0 ð·ð¥2 ð¿ = ð¥1 â 4ð1 ð¥2 â 2ð2 ð¥2 = 0 can be written as ð¥2 = 2(ð1 + 2ð2 )ð¥1
(5.96)
ð¥1 = 2(2ð1 + ð2 )ð¥2
(5.97)
which must be satisï¬ed along with the complementary slackness conditions ð¥21 + 2ð¥22 â 3 †0
ð1 ⥠0
ð1 (ð¥21 + 2ð¥22 â 3) = 0
2ð¥21 + ð¥22 â 3 †0
ð2 ⥠0
ð2 (2ð¥21 + ð¥22 â 3) = 0
First suppose that both constraints are slack so that ð1 = ð2 = 0. Then the ï¬rst-order conditions (5.96) and (5.97) imply that ð¥1 = ð¥2 = 0. (0, 0) satisï¬es the Kuhn-Tucker conditions. Next suppose that the ï¬rst constraint is binding while the second constraint have two solutions, is slack â (ð2 = 0).â The ï¬rst-order â and (5.97) â â conditions (5.96) â ð¥1 = 3/2, ð¥2 = 3/2, ð = 1/(2 2) and ð¥1 = â 3/2, ð¥2 = â 3/2, ð = 1/(2 2), but these violate the second constraint. Similarly, there is no solution in which the ï¬rst constraint is slack and the second constraint binding. Finally, assume that the both constraints are binding. This implies that ð¥1 = ð¥2 = 1 or ð¥1 = ð¥2 = â1, which points satisfy the ï¬rst-order conditions (5.96) and (5.97) with ð1 = ð2 = 1/6. We conclude that three points satisfy the Kuhn-Tucker conditions, namely (0, 0), (1, 1) and (â1, â1). Noting the objective function, we observe that (0, 0) in fact minimizes the objective. We conclude that there are two local maxima, (1, 1) and (â1, â1), both of which achieve the same level of the objective function. 5.31 Dividing the ï¬rst-order conditions, we obtain ð ð(ð â ð) ð·ð ð
(ð, ð) = â ð·ð ð
(ð, ð) ð€ (1 â ð)ð€ Using the revenue function ð
(ð, ð) = ð(ð (ð, ð))ð (ð, ð) the marginal revenue products of capital and labor are ð·ð ð
(ð, ð) = ð·ðŠ ð(ðŠ)ð·ð ð (ð, ð) ð·ð ð
(ð, ð) = ð·ðŠ ð(ðŠ)ð·ð ð (ð, ð) so that their ratio is equal to the ratio of the marginal products ð·ð ð (ð, ð) ð·ð ð
(ð, ð) = ð·ð ð
(ð, ð) ð·ð ð (ð, ð 247
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
The necessary condition for optimality can be expressed as ð ð ð âð ð·ð ð (ð, ð) = â ð·ð ð (ð, ð) ð€ 1âð ð€ whereas the necessary condition for cost minimization is (Example 5.16) ð·ð ð (ð, ð) ð = ð·ð ð (ð, ð) ð€ The regulated ï¬rm does not use the cost-minimizing combination of inputs. 5.32 The general constrained optimization problem max ð (x) x
subject to g(x) †0 can be transformed into an equivalent equality constrained problem max ð (x) x,s
subject to g(x) + s = 0 and s ⥠0 Ë (x, s) = g(x) + s, the through the addition of nonnegative slack variables s. Letting g ï¬rst-order conditions a local optimum are (Exercise 5.16) â â ð·x ð (xâ ) = ðð ð·x ðËð (xâ , sâ ) = ðð ð·x ðð (xâ ) â ðð ð·s ðËð (x, s) = ð (5.98) 0 = ð·s ð (xâ ) †ðð s = 0
sâ¥0
(5.99)
Condition (5.98) implies that ðð ⥠0 for every ð. Furthermore, rewriting the constraint as s = âg(x) the complementary slackness condition (5.99) becomes ðð g(x) = 0
g(x) †0
This establishes the necessary conditions of Theorem 5.3. 5.33 The equality constrained maximization problem max ð (x) x
subject to g(x) = 0 is equivalent to the problem max ð (x) x
subject to g(x) †0 âg(x) †â0 By the Kuhn-Tucker theorem (Theorem 5.3), there exists nonnegative multipliers + â â + â ð+ 1 , ð2 , . . . , ðð and ð1 , ð2 , . . . , ðð such that â â â â ð·ð (xâ ) = ð+ ðâ (5.100) ð ð·ðð [x ] â ð ð·ðð [x ] = 0 248
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics with â ð+ ð ðð (x) = 0 and ðð ðð (x) = 0
ð = 1, 2, . . . , ð
â Deï¬ning ðð = ð+ ð â ðð , (5.100) can be written as
ð·ð (xâ ) =
â
ðð ð·ðð [xâ ]
which is the ï¬rst-order condition for an equality constrained problem. Furthermore, if xâ satisï¬es the inequality constraints ð(xâ ) †0 and ð(xâ ) ⥠0 it satisï¬es the equality ð(xâ ) = 0 5.34 Suppose that xâ solves the problem max cð x subject to ðŽx †0 x
with Lagrangean ð¿ = cð x â ðð ðŽx Then there exists ð ⥠0 such that ð·x ð¿ = cð â ðð ðŽ = 0 that is, ðŽð ð = c. Conversely, if there is no solution, there exists x such that ðŽx †0 and cð x > cð 0 = 0 5.35 There are two binding constraints at (4, 0), namely ð(ð¥1 , ð¥2 ) = ð¥1 + ð¥2 †4 â(ð¥1 , ð¥2 ) = âð¥2 †0 with gradients âð(4, 0) = (1, 1) ââ(4, 0) = (0, 1) which are linearly independent. Therefore the binding constraints are regular at (0, 4). 5.36 The Lagrangean for this problem is ð¿(x, ð) = ð¢(x) â ð(pð x â ð) and the ï¬rst-order (Kuhn-Tucker) conditions are (Corollary 5.3.2) ð·ð¥ð ð¿[xâ , ð] = ð·ð¥ð ð¢[xâ ] â ððð †0 ð
â
p x â€ð
xâð ⥠0 ðâ¥0
ð¥âð (ð·ð¥ð ð¢[xâ ] â ððð ) = 0 ð
â
ð(p x â ð) = 0
for every good ð = 1, 2, . . . , ð. Two cases must be distinguished. 249
(5.101) (5.102)
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Case 1 ð > 0 This implies that pð x = ð, the consumer spends all her income. Condition (5.101) implies ð·ð¥ð ð¢[xâ ] †ððð for every ð with ð·ð¥ð ð¢[xâ ] = ððð for every ð for which ð¥ð > 0 This case was analyzed in Example 5.17. Case 2 ð = 0 This allows the possibility that the consumer does not spend all her income. Substituting ð = 0 in (5.101) we have ð·ð¥ð ð¢[xâ ] = 0 for every ð. At the optimal consumption bundle xâ , the marginal utility of every good is zero. The consumer is satiated, that is no additional consumption can increase satisfaction. This case was analyzed in Example 5.31. In summary, at the optimal consumption bundle xâ , either â the consumer is satiated (ð·ð¥ð ð¢[xâ ] = 0 for every ð) or â the consumer consumes only those goods whose marginal utility exceeds the threshold ð·ð¥ð ð¢[xâ ] ⥠ððð and adjusts consumption so that the marginal utility is proportional to price for all consumed goods. 5.37 Assume x â ð·(xâ ). Then there exists ðŒ ¯ â â such that xâ + ðŒx â ð for every 0 †ðŒ †ðŒ. ¯ Deï¬ne ð â ð¹ ([0, ðŒ ¯ ]) by ð(ðŒ) = ð (xâ + ðŒx). If xâ is a local maximum, ð has a local maximum at 0, and therefore ð â² (0) †0 (Theorem 5.1). By the chain rule (Exercise 4.22), this implies ð â² (0) = ð·ð [xâ ](x) †0 and therefore x â / ð» + (xâ ). 5.38 If x is a tangent vector, so is ðœx for any nonnegative ðœ (replace 1/ðŒð by ðœ/ðŒð in the preceding deï¬nition. Also, trivially, x = 0 is a tangent vector (with xð = xâ and ðŒð = 1 for all ð). The set ð of all vectors tangent to ð at xâ is therefore a nonempty cone, which is called the cone of tangents to ð at xâ . To show that ð is closed, let xð be a sequence in ð converging to some x â âð . We need to show that x â ð . Since xð â ð , there exist feasible points xðð â ð and ðŒðð such that (xðð â xâ )/ðŒðð â xð as ð â â For any ð choose ð such that â¥xð â x⥠â€
1 ð 2
and then choose ð such that â¥xðð â xâ ⥠†ð and â¥(xðð â xâ )/ðŒðð â xð ⥠â€
1 ð 2
Relabeling xðð as xð and ðŒðð as ðŒð we have we have constructed a sequence xð in S such that ð x â xâ †ð and
ð (x â xâ )/ðŒð â x †(xð â xâ )/ðŒð â xð + â¥xð â x⥠†1 ð 1
Letting ð â â, xð converges to xâ and (xð â xâ )/ðŒð converges to x, which proves that x â ð as required. 250
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
5.39 Assume x â ð·(xâ ). That is, there exists ðŒ ¯ such that xâ + ðŒx â ð for every ðŒ â [0, ðŒ ¯ ]. For ð = 1, 2, . . . , let ðŒð = ðŒ ¯ /ð. Then xð = xâ + ðŒð x â ð, xð â xâ and ð â â â (x â x )/ðŒð = (x + ðŒð x â x )/ðŒð = x. Therefore, x â ð (xâ ). 5.40 Let dx â ð (xâ ). Then there exists a feasible sequence {xð } converging to xâ and a sequence {ðŒð } of nonnegative scalars such that the sequence {(xð â xâ )/ðŒð } converges to dx. For any ð â ðµ(xâ ), ðð (xâ ) = 0 and ðð (xð ) = ð·ðð [xâ ](xð â xâ ) + ðð xð â xâ where ðð â 0 as k â â. This implies 1 1 ðð (xð ) = ð ð·ðð [xâ ](xð â xâ ) + ðð (xð â xâ )/ðŒð ð ðŒ ðŒ Since xð is feasible 1 ðð (xð ) †0 ðŒð and therefore ð·ðð [xâ ]((xð â xâ )/ðŒð ) + ð ð (xð â xâ )/ðŒð †0 Letting ð â â we conclude that ð·ðð [xâ ](dx) †0 That is, dx â ð¿. 5.41 ð¿0 â ð¿1 by deï¬nition. Assume dx â ð¿1 . That is ð·ðð [xâ ](dx) < 0 â
ð·ðð [x ](dx) †0
for every ð â ðµ ð (xâ ) ð¶
â
for every ð â ðµ (x )
(5.103) (5.104)
where ðµ ð¶ (xâ ) = ðµ(xâ ) â ðµ ð (xâ ) is the set of concave binding constraints at xâ . By concavity (Exercise 4.67), (5.104) implies that ðð (xâ + ðŒdx) †ðð (xâ ) = 0 for every ðŒ ⥠0 and ð â ðµ ð¶ (xâ ) From (5.103) there exists some ðŒð such that ðð (xâ + ðŒdx) < 0 for every ðŒ â [0, ðŒð ] and ð â ðµ ð (xâ ) Furthermore, since ðð (xâ ) < 0 for all ð â ð(xâ ), there exists some ðŒð > 0 such that ðð (xâ + ðŒdx) < 0 for every ðŒ â [0, ðŒð ] and ð â ð(xâ ) Setting ðŒ ¯ = min{ðŒð , ðŒð } we have ðð (xâ + ðŒdx) †0 for every ðŒ â [0, ðŒ ¯ ] and ð = 1, 2, . . . , ð or xâ + ðŒdx â ðº = { x : ðð (x) †0, ð = 1, 2, . . . , ð } for every ðŒ â [0, ðŒ ¯] Therefore dx â ð·. We have previously shown (Exercises 5.39 and 5.40) that ð· â ð â ð¿. 251
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
5.42 Assume that g satisï¬es the Quasiconvex CQ condition at xâ . That is, for every Ë such that ðð (Ë ð â ðµ(xâ ), ðð is quasiconvex, âðð (xâ ) â= 0 and there exists x x) < 0. Ë â xâ . Quasiconvexity and regularity implies that Consider the perturbation dx = x for every binding constraint ð â ðµ(xâ ) (Exercises 4.74 and 4.75) ðð (Ë x) < ðð (xâ ) =â âðð (xâ )ð (Ë x â xâ ) = âðð (xâ )ð dx < 0 That is ð·ðð [xâ ](dx) < 0 Therefore, dx â ð¿0 (xâ ) â= â
and g satisï¬es the Cottle constraint qualiï¬cation condition. 5.43 If the binding constraints ðµ(xâ ) are regular at xâ , their gradients are linearly independent. That is, there exists no ðð â= 0, ð â ðµ(xâ ) such that â ðð âðð [xâ ] = 0 ðâðµ(xâ )
By Gordanâs theorem (Exercise 3.239), there exists dx â âð such that âðð [xâ ]ð dx < 0 for every ð â ðµ(xâ ) Therefore dx â ð¿0 (xâ ) â= â
. 5.44 If ðð concave, ðµ ð (xâ ) = â
, and AHUCQ is trivially satisï¬ed (with dx = 0 â ð¿1 ). For every ð, let ðð = { dx : ð·ðð [xâ ](dx) < 0 } Then
â ð¿1 (xâ ) = â
â
â©
ðð â
â©
â
â©
â
ðâðµ ð (xâ )
â ðð â
ðâðµ ð¶ (xâ )
where ðµ ð¶ (xâ ) and ðµ ð (xâ ) are respectively the concave and nonconcave constraints binding at xâ . If ðð satisï¬es the AHUCQ condition, ð¿1 (xâ ) â= â
and Exercise 1.219 implies that â â â â â© â© â© ð¿1 = â ðð â â ðð â ðâðµ ð (xâ )
ðâðµ ð¶ (xâ )
Now ðð = { dx : ð·ðð [xâ ](dx) †0 } and therefore
â©
ð¿1 =
ðð = ð¿
ðâðµ(xâ )
Since (Exercise 5.41) ð¿1 â ð â ð¿ and ð is closed (Exercise 5.38), we have ð¿ = ð¿1 â ð â ð¿ which implies that ð = ð¿. 252
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
5.45 For each ð = 1, 2, . . . , ð, either ðð (xâ ) < 0 which implies that ðð = 0 and therefore ðð ð·ðð [xâ ](x â xâ ) = 0 or â
ðð (x ) = 0 Since ðð is quasiconvex and ðð (x) †0 = ð(xâ ), Exercise 4.73 implies that ð·ðð [xâ ](x â xâ ) †0. Since ðð ⥠0, this implies that ðð ð·ðð [xâ ](x â xâ ) †0. We have shown that for every ð, ðð ð·ðð [xâ ](x â xâ ) †0. The ï¬rst-order condition implies that â ðð ð·ðð [xâ ](x â xâ ) †0 ð·ð [xâ ](x â xâ ) = ð
If âð (xâ ) â€
â
ðð âðð (xâ )
xâ ⥠0
( )ð âð (xâ ) â ðð âðð (xâ ) xâ = 0
The ï¬rst-order conditions imply that for every x â ðº, x ⥠0 and (
)ð âð (xâ ) â ðð âðð (xâ ) x †0
and therefore ( )ð âð (xâ ) â ðð âðð (xâ ) (x â xâ ) †0 or âð (xâ )ð (x â xâ ) â€
â
ðð âðð (xâ )ð (x â xâ ) †0
5.46 Assuming ð¥ð = ð¥ð = 0, the constraints become 2ð¥ð †30 2ð¥ð †25 ð¥ð †20 The ï¬rst and third conditions are redundant, which implies that ðð = ðð = 0. Complementary slackness requires that, if ð¥ð > 0, ð·ð¥ð ð¿ = 1 â 2ðð â 2ðð â ðð = 0 or ðð = 12 . Evaluating the Lagrangean at (0, 1/2, 0) yields )) ( ( 1 = 3ð¥ð + ð¥ð + 3ð¥ð ð¿ x, 0, , 0 2 1 â (ð¥ð + 2ð¥ð + 3ð¥ð â 25) 2 25 5 3 = + ð¥ð + ð¥ð 2 2 2 This basic feasible solution is clearly not optimal, since proï¬t would be increased by increasing either ð¥ð or ð¥ð .
253
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Following the hint, we allow ð¥ð > 0, retaining the assumption that ð¥ð = 0. We must be alert to the possibility that ð¥ð = 0. With ð¥ð = 0, the constraints become 2ð¥ð + ð¥ð †30 2ð¥ð + 3ð¥ð †25 ð¥ð + ð¥ð †20 The ï¬rst constraint is redundant, which implies that ðð = 0. If ð¥ð > 0, complementary slackness requires that ð·ð¥ð ð¿ = 3 â 3ðð â ðð = 0 or ðð = 3(1 â ðð )
(5.105)
The requirement that ðð ⥠0 implies that ðð †1. Substituting (5.105) in the second ï¬rst-order condition ð·ð¥ð ð¿ = 1 â 2ðð â ðð = 1 â 2ðð â 3(1 â ðð ) = â2 + ðð implies that ð·ð¥ð ð¿ = â2 + ðð < 0
for every ðð †1
Complementary slackness then requires implies that ð¥ð = 0. The constraints now become ð¥ð †30 3ð¥ð †25 ð¥ð †20 The ï¬rst and third are redundant, so that ðð and ðð = 0. Equation (5.105) implies that ðð = 1. Evaluating the Lagrangean at this point (ð = 0, 1, 0), we have ð¿(ð¥, (0, 1, 0)) = 3ð¥ð + ð¥ð + 3ð¥ð â (ð¥ð + 2ð¥ð + 3ð¥ð â 25) = 25 + 2ð¥ð â ð¥ð Clearly this is not an optimal solution, An increase in ð¥ð is indicated. This leads us to the hypothesis ð¥ð > 0, ð¥ð > 0, ð¥ð = 0 which was evaluated in the text, and in fact lead to the optimal solution. 5.47 If we ignore the hint and consider solutions with ð¥ð > 0, ð¥ð ⥠0, ð¥ð = 0, the constraints become 2ð¥ð + 2ð¥ð †30 ð¥ð + 2ð¥ð †25 2ð¥ð + ð¥ð †20 These three constraints are linearly dependent, so that any one of them is redundant and can be eliminated. For example, 3/2 times the ï¬rst constraint is equal to the sum of 254
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
the second and third constraints. The feasible solution ð¥ð = 0, ð¥ð = 5, ð¥ð = 10, where the constraints are linearly dependent, is known as a degenerate solution. Degeneracy is a signiï¬cant feature of linear programming, allowing the theoretical possibility of a breakdown in the simplex algorithm. Fortunately, such breakdown seems very rare in practice. Degeneracy at the optimal solution indicates multiple optima. One way to proceed in this example is to arbitrarily designate one constraint as redundant, assuming the corresponding multiplier is zero. Arbitrarily choosing ðð = 0 and proceeding as before, complementary slackness (ð¥ð > 0) requires that ð·ð¥ð ð¿ = 3 â 2ðð â ðð = 0 or ðð = 3 â 2ðð
(5.106)
Nonnegativity of ðð implies that ðð †32 . Substituting (5.106) in the second ï¬rst-order condition yields ð·ð¥ð ð¿ = 1 â 2ðð â 2ðð = 1 â 2ðð â 2(3 â 2ðð ) = â5 + 2ðð < 0 for every ðð â€
3 2
Complementary slackness therefore implies that ð¥ð = 0, which takes us back to the starting point of the presentation in the text, where ð¥ð > 0, ð¥ð = ð¥ð = 0. 5.48 Assume that (c1 , ð§1 ) and (c2 , ð§2 ) belong to ðµ. That is ð§1 ⥠ð§ â ð§2 ⥠ð§ â
c1 †0 c2 †0 For any ðŒ â (0, 1),
ð§Â¯ = ðŒð§1 + (1 â ðŒ)ð§2 †ð§ â
¯ c = ðŒc1 + (1 â ðŒ)c2 †0
and therefore (¯ c, ð§Â¯) â ðµ. This shows that ðµ is convex. Let 1 = (1, 1, . . . , 1) â âð . Then (c â 1, ð§ + 1) â int ðµ â= â
. There ðµ has a nonempty interior. 5.49 Let (c, ð§) â int ðµ. This implies that c < 0 and ð§ > ð§ â . Since ð£ is monotone ð£(c) †ð£(0) = zâ < ð§ which implies that (c, ð§) â / ðŽ. 5.50 The linear functional ð¿ can be decomposed into separate components, so that there exists (Exercise 3.47) ð â ð â and ðŒ â â such that ð¿(c, ð§) = ðŒð§ â ð(c) Assuming ð â âð , there exists (Proposition 3.4) ð â âð such that ð(c) = ðð c and therefore ð¿(c, ð§) = ðŒð§ â ðð c The point (0, ð§ â + 1) belongs to ðµ. Therefore, by (5.75), ð¿(0, ð§ â ) †ð¿(0, ð§ â + 1) 255
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics which implies that ðŒð§ â â ðð 0 †ðŒ(ð§ â + 1) â ðð 0
or ðŒ ⥠0. Similarly, let { e1 , e2 , . . . , eð } denote the standard basis for âð (Example 1.79). For any ð = 1, 2, . . . , ð, the point (0 â eð , ð§ â ) (which corresponds to decreasing resource ð by one unit) belongs to ðµ and therefore (from (5.75)) ð§ â â ðð (0 â eð ) = ð§ â + ðð ⥠ð§ â â ðð 0 = ð§ â which implies that ðð ⥠0. 5.51 Let cË = ð(Ë x) < 0 and ð§Ë = ð (Ë x) Suppose ðŒ = 0. Then, since ð¿ is nonzero, at least one component of ð must be nonzero. That is, ð â© 0 and therefore ðð ðË < 0
(5.107)
But (Ë c, ð§Ë) â ðŽ and (5.74) implies ðŒË ð§ â ðð cË â€ ðŒð§ â â ðð 0 and therefore ðŒ = 0 implies ðð ðË â¥ 0 contradicting (5.107). Therefore, we conclude that ðŒ > 0. 5.52 The utilityâs optimization problem is max ð(ðŠ, ð ) =
ðŠ,ð â¥0
ð â« â ð=1
ðŠð 0
(ðð (ð ) â ðð )ðð â ð0 ð
subject to ðð (y, ð ) = ðŠð â ð †0
ð = 1, 2, . . . , ð
The demand independence assumption ensures that the objective function ð is concave, since its Hessian â â ð·ð1 0 . . . 0 0 â 0 ð·ð2 . . . 0 0â â ð»ð = â â 0 ... ð·ðð 0â 0 ... 0 0 is nonpositive deï¬nite (Exercise 3.96). The constraints are linear and hence convex. Moreover, there exists a point (0, 1) such that for every ð = 1, 2, . . . , ð ðð (0, 1) = 0 â 1 < 0 Therefore the problem satisï¬es the conditions of Theorem 5.6. The optimal solution (yâ , ð â ) satisï¬es the Kuhn-Tucker conditions, that is there exist multipliers ð1 , ð2 , . . . , ðð such that for every period ð = 1, 2, . . . , ð ð·ðŠð ð¿ = ðð (ðŠð ) â ðð â ðð †0
ðŠð ⥠0
ðŠð (ðð (ðŠð ) â ðð â ðð ) = 0
ðŠð †ð
ðð ⥠0
ð(ð â ðŠð ) = 0
256
(5.108)
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics and that capacity be chosen such that ð· ð ð¿ = ð0 â
ð â
( ðð †0
ð â¥0
ð
ð0 â
ð=1
ð â
) ðð
=0
(5.109)
ð=1
where ð¿ is the Lagrangean ð¿(ðŠ, ð, ð) =
ð â« â ð=1
0
ðŠð
(ðð (ð ) â ðð )ðð â ð0 ð â
ð â
ðð (ðŠð â ð )
ð=1
In oï¬-peak periods (ðŠð < ð ), complementary slackness requires that ðð = 0 and therefore from (5.108) ðð (ðŠð ) = ðð assuming ðŠð > 0. In peak periods (ðŠð = ð ) ðð (ðŠð ) = ðð + ðð We conclude that it is optimal to price at marginal cost in oï¬-peak periods and charge a premium during peak periods. Furthermore, (5.109) implies that the total premium is equal to the marginal capacity cost ð â
ðð = ð0
ð=1
Furthermore, note that ð â
ðð ðŠð =
ð=1
â Peak
=
=
ðð ðŠð
Oï¬-peak
â
ðð ðŠð +
ðŠð =ð
=
â
ðð ðŠð +
â
â
ðð ðŠð
ðð =0
ðð ð
ðŠð =ð ð â
ðð ð = ð0 ð
ð=1
Therefore, the utilityâs total revenue is ð
(ðŠ, ð ) = = = =
ð â ð=1 ð â ð=1 ð â ð=1 ð â
ðð (ðŠð )ðŠð (ðð + ðð )ðŠð ðð ðŠ ð +
ð â
ðð ðŠð
ð=1
ðð ðŠð + ð0 ð = ð(ðŠ, ð )
ð=1
Under the optimal pricing policy, revenue equals cost and the utility breaks even. 257
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Chapter 6: Comparative Statics 6.1 The Jacobian is
( ð»ð¿ ðœg
ðœ=
ðœgð 0
)
where ð»ð¿ is the Hessian of the Lagrangean. We note that â ð»ð¿ (x0 ) is negative deï¬nite in the subspace ð = { x : ðœg x = 0 } (since x0 satisï¬es the conditions for a strict local maximum) â ðœg has rank ð (since the constraints are regular). Consider the system of equations ( ð»ð¿ ðœg
ðœgð 0
)( ) ( ) x 0 = y 0
(6.28)
where x â âð and y â âð . It can be decomposed into ð»ð¿ x + ðœgð y = 0
(6.29)
ðœg x = 0
(6.30)
Suppose x solves (6.30). Multiplying (6.29) by xð gives xð ð»ð¿ x + xð ðœgð y = xð ð»ð¿ x + (ðœg x)ð y = 0 But (6.30) implies that the second term is 0 and therefore xð ð»ð¿ x = 0. Since ð»ð¿ is positive deï¬nite on ð = { x : ðœg x = 0 }, we must have x = 0. Then (6.29) reduces to ðœgð y = 0 Since ðœg has rank ð, this has only the trivial solution y = 0 (Section 3.6.1). We have shown that the system (6.38) has only the trivial solution (0, 0). This implies that the matrix ðœ is nonsingular. 6.2 The Lagrangean for this problem is
( ) ð¿ = ð (x) â ðð g(x) â c
By Corollary 6.1.1 âð£(c) = ð·c ð¿ = ð 6.3 Optimality implies ð (x1 , ðœ1 ) ⥠ð (x, ðœ 1 ) and ð (x2 , ðœ 2 ) ⥠ð (x, ðœ2 ) for every x â ð In particular ð (x1 , ðœ1 ) ⥠ð (x2 , ðœ1 ) and ð (x2 , ðœ2 ) ⥠ð (x1 , ðœ2 ) Adding these inequalities ð (x1 , ðœ1 ) + ð (x2 , ðœ2 ) ⥠ð (x2 , ðœ1 ) + ð (x1 , ðœ2 ) 258
Solutions for Foundations of Mathematical Economics
c 2001 Michael Carter â All rights reserved
Rearranging and using the bilinearity of ð gives ð (x1 â x2 , ðœ1 ) ⥠ð (x1 â x2 , ðœ2 ) and ð (x1 â x2 , ðœ 1 â ðœ 2 ) ⥠0 6.4 Let ð1 denote the proï¬t maximizing price with the cost function ð1 (ðŠ) and let ðŠ1 be the corresponding output. Similarly let ð2 and ðŠ2 be the proï¬t maximizing price and output when the costs are given by ð2 (ðŠ). With cost function ð1 , the ï¬rms proï¬t is Î = ððŠ â ð1 (ðŠ) Since this is maximised at ð1 and ðŠ1 (although the monopolist could have sold ðŠ2 at price ð2 ) ð1 ðŠ1 â ð1 (ðŠ1 ) ⥠ð2 ðŠ2 â ð1 (ðŠ2 ) Rearranging ð1 ðŠ1 â ð2 ðŠ2 ⥠ð1 (ðŠ1 ) â ð1 (ðŠ2 )
(6.31)
The increase in revenue in moving from ðŠ2 to ðŠ1 is greater than the increase in cost. Similarly ð2 ðŠ2 â ð2 (ðŠ2 ) ⥠ð1 ðŠ1 â ð2 (ðŠ1 ) which can be rearranged to yield ð2 (ðŠ1 ) â ð2 (ðŠ2 ) ⥠ð1 ðŠ1 â ð2 ðŠ2 Combining the previous inequality with (6.31) yields ð2 (ðŠ1 ) â ð2 (ðŠ2 ) ⥠ð1 (ðŠ1 ) â ð1 (ðŠ2 )
(6.32)
6.5 By Theorem 6.2 ð·w Î [w, ð] = âxâ and ð·ð Î [w, ð] = ðŠ â and therefore 2 Î (ð, w) ⥠0 ð·ð ðŠ(ð, w) = ð·ðð 2 Î (ð, w) †0 ð·ð€ð ð¥ð (ð, w) = âð·ð€ ð ð€ð 2 Î (ð, w) = ð·ð€ð ð¥ð (ð, w) ð·ð€ð ð¥ð (ð, w) = âð·ð€ ð ð€ð 2 Î (ð, w) = âð·ð€ð ðŠ(ð, w) ð·ð ð¥ð (ð, w) = âð·ð€ ðð
since Î is convex and therefore ð»Î (w, ð) is symmetric (Theorem 4.2) and nonnegative deï¬nite (Proposition 4.1). 6.6 By Shephardâs lemma (6.17) ð¥ð (ð€, ðŠ) = ð·ð€ð ð(ð€, ðŠ) 259
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics Using Youngâs theorem (Theorem 4.2), 2 ð·ðŠ ð¥ð [w, ðŠ] = ð·ð€ ð[w, ðŠ] ððŠ 2 ð[w, ðŠ] = ð·ðŠð€ ð
= ð·ð€ð ð·ðŠ ð[w, ðŠ] Therefore ð·ðŠ ð¥ð [w, ðŠ] ⥠0 ââ ð·ð€ð ð·ðŠ ð[w, ðŠ] ⥠0 6.7 The demand functions must satisfy the budget contraint identically, that is ð â
ðð ð¥ð (p, ð) = ð for every p and ð
ð=1
Diï¬erentiating with respect to m ð â
ðð ð·ð ð¥ð [p, ð] = 1
ð=1
This is the Engel aggregation condition, which simply states that any additional income be spent on some goods. Multiplying each term by ð¥ð ð/(ð¥ð ð) ð â ðð ð¥ð ð=1
ð ð·ð ð¥ð [p, ð] = 1 ð ð¥ð (p, ð)
the Engel aggregation condition can be written in elasticity form ð â
ðŒð ðð = 1
ð=1
where ðŒð = ðð ð¥ð /ð is the budget share of good ð. On average, goods must have unit income elasticities. Diï¬erentiating the budget constraint with respect to ðð ð â
ðð ð·ðð ð¥ð [p, ð] + ð¥ð (ð, ð) = 0
ð=1
This is the Cournot aggregation condition, which implies that an increase in the price of ðð is equivalent to a decrease in real income of ð¥ð ððð . Multiplying each term in the sum by ð¥ð /ð¥ð gives ð â ðð ð¥ð ð=1
ð¥ð
ð·ðð ð¥ð [p, ð] = âð¥ð
Multiplying through by ðð /ð ð â ðð ð¥ð ðð ð=1
ð ð¥ð
ð·ðð ð¥ð [p, ð] = â
ð â
ðŒð ððð = âðŒð
ð=1
260
ðð ð¥ð ð
c 2001 Michael Carter â All rights reserved
Solutions for Foundations of Mathematical Economics
6.8 Supermodularity of Î (x, ð, âw) follows from Exercises 2.50 and 2.51. To show strictly increasing diï¬erences, consider two price vectors w2 ⥠w1 Î (x, ð, âw1 ) â Î (x, ð, âw2 ) =
ð â
(âð€ð1 )ð¥ð â
ð=1
=
ð â
ð â
(âð€ð2 )ð¥ð
ð=1
(ð€ð2 â ð€ð1 )ð¥ð
ð=1
âð
Since w2 ⥠w1 , w2 â w1 ⥠0 and
2 ð=1 (ð€ð
â ð€ð1 )ð¥ð is strictly increasing in x.
6.9 For any ð2 â¥( ð1 , ðŠ 2 = ð (ð2 ) †ð (ð1 ) = ðŠ 1 and ð(ðŠ 1 , ð) â ð(ðŠ 2 , ð) is increasing in ð and therefore â ð(ð (ð2 ), ð) â ð(ð (ð1 ), ð)) is increasing in ð. 6.10 The ï¬rmâs optimization problem is max ðððŠ â ð(ðŠ)
ðŠââ+
The objective function ð (ðŠ, ð, ð) = ðððŠ â ð(ðŠ) is â supermodular in ðŠ (Exercise 2.49) â displays strictly increasing diï¬erences in (ðŠ, ð) since ( ) ð (ðŠ 2 , ð, ð) â ð (ðŠ 1 , ð, ð) = ðð(ðŠ 2 â ðŠ 1 ) â ð(ðŠ 2 ) â ð(ðŠ 1 ) is strictly increasing in ð for ðŠ 2 > ðŠ 1 . Therefore (Corollary 2.1.2), the ï¬rmâs output correspondence is strongly increasing and every selection is increasing (Exercise 2.45). Therefore, the ï¬rmâs output increases as the yield increases. It is analogous to an increase in the exogenous price. 6.11 With two factors, the Hessian is ( ð11 ð»ð = ð21
ð12 ð22
)
Therefore, its inverse is (Exercise 3.104) ð»ðâ1
1 = Î
(
ð22 âð21
âð12 ð11
)
where Î = ð11 ð22 â ð12 ð21 ⥠0 by the second-order condition. Therefore, the Jacobian of the demand functions is ( ) ) ( 1 1 â1 ð·ð€1 ð¥1 ð·ð€2 ð¥1 ð22 âð12 = ð»ð = ð·ð€1 ð¥2 ð·ð€2 ð¥2 ð11 ð ðÎ âð21 Therefore ð21 ð·ð€1 ð¥2 = â ðÎ
{
0 otherwise
View more...
Comments