Skip to main content
Log in

Bayesian general equilibrium

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

I introduce a general equilibrium model of non-optimizing agents that respond to aggregate variables (prices and the average demand profile of agent types) by putting a “prior” on their demand. An interim equilibrium is defined by the posterior demand distribution of agent types conditional on market clearing. A Bayesian general equilibrium (BGE) is an interim equilibrium such that aggregate variables are correctly anticipated. Under weak conditions, I prove the existence and the informational efficiency of BGE. I discuss the conditions under which the set of Bayesian and Walrasian equilibria coincide and show that the Walrasian equilibrium arises from a large class of non-optimizing behavior.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. The support of a Borel measure \(\mu \) on a second countable topological space \(X\), denoted by \(S=\mathrm{supp }\mu \), consists of the points that do not have a neighborhood with measure zero. Let \(\fancyscript{U}\) be the countable base of the topology (for the case of \(X=\mathbb {R}^C\), it suffices to take the family of all open balls with rational radii and centers with rational coordinates) and \(S=X\backslash \bigcup _{U\in \fancyscript{U}:\mu (U)=0}U\). Since \(\fancyscript{U}\) is the countable base of the topology, it follows that \(S\) is closed, \(\mu (S)=\mu (X)\), and \(\mu (U\cap S)>0\) whenever \(U\) is open and \(U\cap S\ne \emptyset \). Hence, \(S=\mathrm{supp }\mu \).

  2. I follow Foley (1994) in calling \(X_i(p,x)\) the offer set, but a more appropriate term might be offer correspondence, since sets are parametrized by the price vector \(p\) and the average demand profile \(x\).

  3. This definition can be understood as follows. If we take the prior measure \(\mu _i(p,x)\) as the reference measure \(\mu \) in (7.2) and noting that \(f_i\) corresponds to the posterior density, we get \(q=1\) and hence the Kullback–Leibler information of a single agent of type \(i\) is

    $$\begin{aligned} H_i=H(f_i;1)=\int f_i\log f_i\mathrm {d}\mu _i(p,x). \end{aligned}$$

    Now suppose that there are \(N_i\) agents of type \(i\), and let the total number of agents be \(N=\sum \nolimits _i N_i\) and the proportion be \(n_i=N_i/N\). In general, the entropy of the joint distribution of two independent random variables is the sum of the entropy of each variable. (See, for example, the excellent introductory textbook of Cover and Thomas 2006). The additivity of the entropy carries over to the Kullback–Leibler information. Therefore, the economy-wide information is

    $$\begin{aligned} H=\sum _{i=1}^I N_iH_i=\sum _{i=1}^I N_i\int f_i\log f_i\mathrm {d}\mu _i(p,x). \end{aligned}$$

    Dividing this expression by \(N\), we obtain the per capita information (2.2).

  4. \(\mathrm{cl }A\) and \(\mathrm{co }A\) denote the closure and the convex hull of \(A\), respectively.

  5. Here, I am using the term ‘arbitrage’ informally to refer to a situation that agents execute trades that yielded higher values than they had expected. Thus, ‘arbitrage’ here is synonymous to “getting a good deal by luck.”

  6. See, for example, Equation (6.3.3) in Ljungqvist and Sargent (2004), p. 144.

  7. In the literature this quantity is also known as the relative entropy, cross-entropy, information gain, \(I\)-divergence, etc.

  8. The solution of the more general entropy-like minimization problem can be found in Borwein and Lewis (1991, 1992).

  9. \(\mathrm{dom }f=\left\{ x\in X|f(x)<\infty \right\} \) is the domain of \(f\).

  10. For a subset \(A\) of a vector space, \(\dim A\) denotes the dimension of the smallest affine space that contains \(A\).

  11. For example, \(f_n\rightarrow _c f\) if \(f_n\rightarrow f\) uniformly on compact sets and \(f\) is continuous. To see this, let \(K\) a compact neighborhood of \(x\) and take \(N\) such that \(n>N\) implies \(x_n\in K\). Then

    $$\begin{aligned} \left|f_n(x_n)-f(x)\right|\le \left|f_n(x_n)-f(x_n)\right|+\left|f(x_n)-f(x)\right|\rightarrow 0. \end{aligned}$$

References

  • Becker, G.S.: Irrational behavior and economic theory. J. Polit. Econ. 70(1), 1–13 (1962)

    Article  Google Scholar 

  • Becker, R.A., Chakrabarti, S.K.: Satisficing behavior, Brouwer’s fixed point theorem and Nash equilibrium. Econ. Theory 26(1), 63–83 (2005). doi:10.1007/s00199-004-0519-z

    Article  Google Scholar 

  • Bewley, T.F.: Why Wages Don’t Fall During a Recession? Harvard University Press, Cambridge (1999)

  • Borwein, J.M., Lewis, A.S.: Duality relationships for entropy-like minimization problems. SIAM J. Control Optim. 29(2), 325–338 (1991). doi:10.1137/0329017

    Article  Google Scholar 

  • Borwein, J.M., Lewis, A.S.: Partially finite convex programming, part I: Quasi relative interiors and duality theory. Math. Program. 57(1–3), 15–48 (1992). doi:10.1007/BF01581072

    Article  Google Scholar 

  • Cabrales, A., Gossner, O., Serrano, R.: Entropy and the value of information for investors. Am. Econ. Rev. 103(1), 360–377 (2013). doi:10.1257/aer.103.1.360

    Article  Google Scholar 

  • Caticha, l., Giffin, A.: Updating probabilities. In: Mohammad-Djafari, A. (ed.) Bayesian Inference and Maximum Entropy Methods in Science and Engineering, Volume 872 of AIP Conference Proceedings, pp. 31–42 (2006). doi:10.1063/1.2423258

  • Chakrabarti, S.K.: On the robustness of the competitive equilibrium: utility-improvements and equilibrium points. J. Math. Econ. (2014). doi:10.1016/j.jmateco.2014.08.008

  • Cover, T.M., Thomas, J.A.: Elements of Information Theory, 2nd edn. Wiley, Hoboken (2006)

    Google Scholar 

  • Csiszár, I.: Sanov property, generalized \(I\)-projection and a conditional limit theorem. Ann. Prob. 12(3), 768–793 (1984)

    Article  Google Scholar 

  • Drèze, J.H.: Existence of an exchange equilibrium under price rigidities. Int. Econ. Rev. 16(2), 301–320 (1975). doi:10.2307/2525813

    Article  Google Scholar 

  • Duffie, D., Geanakoplos, J., Mas-Collel, A., McLennan, A.: Stationary Markov equilibria. Econometrica 62(4), 745–781 (1994). doi:10.2307/2951731

    Article  Google Scholar 

  • Foley, D.K.: A statistical equilibrium theory of markets. J. Econ. Theory 62(2), 321–345 (1994). doi:10.1006/jeth.1994.1018

    Article  Google Scholar 

  • Foley, D.K.: Statistical equilibrium in a simple labor market. Metroeconomica 47(2), 125–147 (1996). doi:10.1111/j.1467-999X.1996.tb00792.x

    Article  Google Scholar 

  • Foley, D.K.: Statistical equilibrium in economics: method, interpretation, and an example. In: Fabio, P., Frank, H. (eds.) General Equilibrium: Problems and Prospects, Chapter 4. Routledge, London and New York (2003)

    Google Scholar 

  • Geanakoplos, J.: Nash and Walras equilibrium via Brouwer. Econ. Theory 21(2–3), 585–603 (2003). doi:10.1007/s001990000076

    Article  Google Scholar 

  • Herings, J.-J.P.: Static and Dynamic Aspects of General Disequilibrium Theory, Volume 13 of C: Game Theory, Mathematical Programming and Operations Research. Kluwer Academic Publishers, Dordrecht (1996)

  • Jaynes, E.T.: Information theory and statistical mechanics. Phys. Rev. 106(4), 620–630 (1957). doi:10.1103/PhysRev.106.620

    Article  Google Scholar 

  • Jaynes, E.T.: Where do we stand on maximum entropy. In: Raphael, D.L., Myron, T. (eds.) The Maximum Entropy Formalism, Chapter 1, pp. 15–118. MIT Press, Cambridge (1978)

    Google Scholar 

  • Jaynes, E.T.: Probability Theory: The Logic of Science. In: Bretthorst G.L. (ed.) Cambridge University Press, Cambridge (2003)

  • Kahneman, D., Tversky, A.: Prospect theory: an analysis of decision under risk. Econometrica 47(2), 263–292 (1979). doi:10.2307/1914185

    Article  Google Scholar 

  • Knuth, K.H., Skilling, J.: Foundations of inference. Axioms 1(1), 38–73 (2012). doi:10.3390/axioms1010038

    Article  Google Scholar 

  • Krebs, T.: Statistical equilibrium in one-step forward looking economic models. J. Econ. Theory 73(2), 365–394 (1997). doi:10.1006/jeth.1996.2231

    Article  Google Scholar 

  • Kullback, S., Leibler, R.A.: On information and sufficiency. Ann. Math. Stat. 22(1), 79–86 (1951)

    Article  Google Scholar 

  • Ljungqvist, L., Sargent, T.J.: Recursive Macroeconomic Theory, 2nd edn. MIT Press, Cambridge (2004)

    Google Scholar 

  • Luenberger, D.G.: Optimization by Vector Space Methods. Wiley, New York (1969)

    Google Scholar 

  • McCall, J.J.: Economics of information and job search. Q. J. Econ. 84(1), 113–126 (1970)

    Article  Google Scholar 

  • McKelvey, R.D., Palfrey, T.R.: Quantal response equilibria for normal form games. Games Econ. Behav. 10(1), 6–38 (1995). doi:10.1006/game.1995.1023

    Article  Google Scholar 

  • Serfozo, R.: Convergence of Lebesgue integrals with varying measures. Sankhyā: Indian J. Stat. Ser. A 44(3),380–402 (1982).

  • Shannon, C.E.: A mathematical theory of communication. Bell Syst. Tech. J. 27(379–423), 623–656 (1948)

    Article  Google Scholar 

  • Shore, J.E., Johnson, R.W.: Axiomatic derivation of the principle of maximum entropy and the principle of minimum cross-entropy. IEEE Trans. Inf. Theory 26(1), 26–37 (1980). doi:10.1109/TIT.1980.1056144

    Article  Google Scholar 

  • Simon, H.A.: Theories of decision-making in economics and behavioral science. Am. Econ. Rev. 49(3), 253–283 (1959)

    Google Scholar 

  • Sims, C.A.: Implications of rational inattention. J. Monet. Econ. 50(3), 665–690 (2003). doi:10.1016/S0304-3932(03)00029-1

    Article  Google Scholar 

  • Toda, A.A.: Existence of a statistical equilibrium for an economy with endogenous offer sets. Econ. Theory 45(3), 379–415 (2010). doi:10.1007/s00199-009-0493-6

    Article  Google Scholar 

  • Uhlig, H.: A law of large numbers for large economies. Econ. Theory 8(1), 41–50 (1996). doi:10.1007/BF01212011

    Article  Google Scholar 

  • Van Campenhout, J.M., Cover, T.M.: Maximum entropy and conditional probability. IEEE Trans. Inf. Theory 27(4), 483–489 (1981). doi:10.1109/TIT.1981.1056374

    Article  Google Scholar 

  • Veldkamp, L.L.: Information Choice in Macroeconomics and Finance. Princeton University Press, Princeton (2011)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexis Akira Toda.

Additional information

This paper has been first drafted in 2009 and circulated under various titles. I thank Sylvain Barde, Andrew Barron, Truman Bewley, Gaetano Bloise, Donald Brown, Alessandro Citanna, Paul Feldman, Duncan Foley, Xavier Gabaix, John Geanakoplos, Sander Heinsalu, Johannes Hörner, Kazuya Kamiya, Mishael Milaković, Kohta Mori, Herakles Polemarchakis, Larry Samuelson, Paolo Siconolfi, and seminar participants at Yale and ESHIA 2010 for comments and feedback. I am grateful to two anonymous referees for comments and suggestions that significantly improved the paper. The financial supports from the Cowles Foundation, the Nakajima Foundation, and Yale University are greatly acknowledged.

Appendices

Appendix 1: Bayesian inference and maximum entropy

Given a multinomial distribution \(\mathbf {p}=(p_1,\dots ,p_K)\), where \(p_k\ge 0\) and \(\sum \nolimits _{k=1}^Kp_k=1\), Shannon (1948) defined its entropy by

$$\begin{aligned} H(\mathbf {p})=-\sum _{k=1}^Kp_k\log p_k. \end{aligned}$$
(7.1)

Jaynes (1957) proposed that when we want to assign probabilities \(\mathbf {p}=(p_1,\dots ,p_K)\) given some background information (such as moment constraints), we should maximize the Shannon entropy (7.1) subject to the constraints imposed by the background information. This is the original maximum entropy principle (MaxEnt). A prototypical example of such an inference problem is the die problem in Jaynes (1978), which dates back to a 1962 lecture:

Suppose a die has been tossed \(N\) times and we are told only that the average number of spots up was not 3.5 as one might expect for an “honest” die but 4.5. Given this information, and nothing else, what probability should we assign to \(i\) spots in the next toss?

One drawback of the Shannon entropy is that it is not clear how to define it for distributions on a continuous space (say, Euclidean space). To circumvent this difficulty, Kullback and Leibler (1951) introduced the information measure

$$\begin{aligned} H(p;q)=\int p(x)\log \frac{p(x)}{q(x)}\mathrm {d}x, \end{aligned}$$

where \(q(x)\) is the “prior” and \(p(x)\) is the “posterior” density. Here we are implicitly using the Lebesgue measure \(\mathrm {d}x\) and the density functions with respect to the Lebesgue measure, but it does not need to be so. A big advantage of the Kullback–Leibler information is that it is invariant to the choice of the reference measure: if measures \(P,Q,\mu _1,\mu _2\) are mutually absolutely continuous and \(p_i=\mathrm {d}P/\mathrm {d}\mu _i, q_i=\mathrm {d}Q/\mathrm {d}\mu _i\) are Radon–Nikodym derivatives (which correspond to density functions of probability distributions), then

$$\begin{aligned} H(p_1;q_1)=\int p_1\log \frac{p_1}{q_1}\mathrm {d}\mu _1=\int p_2\log \frac{p_2}{q_2}\mathrm {d}\mu _2=H(p_2,;q_2), \end{aligned}$$

so the choice of \(\mu _1,\mu _2\) is irrelevant. Thus, given any “prior” measure \(Q\) and “posterior” measure \(P\), we can define the Kullback–Leibler informationFootnote 7 of \(P\) with respect to \(Q\) by

$$\begin{aligned} H(P;Q)=\int \frac{\mathrm {d}P}{\mathrm {d}Q}\log \frac{\mathrm {d}P}{\mathrm {d}Q}\mathrm {d}Q=\int p\log \frac{p}{q}\mathrm {d}\mu , \end{aligned}$$
(7.2)

where \(\mu \) is any reference measure and \(p=\mathrm {d}P/\mathrm {d}\mu \), \(q=\mathrm {d}Q/\mathrm {d}\mu \) are Radon–Nikodym derivatives (density functions).

The Shannon entropy (7.1) corresponds to the Kullback–Leibler information (7.2) with respect to the uniform prior on a discrete set modulo the sign and an additive constant. Thus, the maximum entropy principle of Jaynes (1957) can be generalized to what I refer to as the minimum information principle, which prescribes to minimize the Kullback–Leibler information (7.2) subject to the given constraints. Axiomatizations of the minimum information principle have been obtained by Shore and Johnson (1980), Caticha and Giffin (2006), and Knuth and Skilling (2012).

Returning to Bayesian inference, Van Campenhout and Cover (1981) showed that Bayes’s theorem implies the minimum information principle in the following sense: the conditional distribution of a random variable \(X_n\) given the empirical observation

$$\begin{aligned} \frac{1}{N}\sum _{n=1}^NT(X_n)=\bar{T}, \end{aligned}$$

where \(X_n\)’s are i.i.d. with prior density \(g\), converges to \(f_\lambda (x)=\mathrm {e}^{\lambda 'T(x)}g(x)\) (suitably normalized) as \(N\rightarrow \infty \), where \(\lambda \) is chosen to satisfy the population moment constraint

$$\begin{aligned} \int T(x)f_\lambda (x)\mathrm {d}x=\bar{T}. \end{aligned}$$

This \(f_\lambda (x)\) turns out to be the solution to

$$\begin{aligned} \min _f H(f;g)~\text {subject to}~\int T(x)f(x)\mathrm {d}x=\bar{T}, \end{aligned}$$

i.e., the minimum information problem, and that \(\lambda \) is the corresponding Lagrange multiplier.Footnote 8 Although Van Campenhout and Cover (1981) proved the statement only for the case \(T\) is a real function, Csiszár (1984) showed that the same conclusion holds even if \(T\) is vector-valued and the sample moment constraints \(\frac{1}{N}\sum \nolimits _{n=1}^NT(X_n)=\bar{T}\) are replaced by the condition that the sample moments belong to a specified convex set, in particular with inequality constraints. Thus, computing the posterior distribution (in the Bayesian sense) reduces to solving the minimum Kullback–Leibler information problem, at least in the large sample limit.

Appendix 2: Proof of equilibrium existence

Proof of Step 1

That \(H^*(\xi ;p,x)\) is \(C^1\) in \(\xi \) and is differentiable under the integral sign follow by Assumption 1 and Lebesgue’s dominated convergence theorem. I only show that \(H^*(\xi ;p,x)\) is continuous in \((p,x,\xi )\) since the case for its first derivatives is similar. Let \((p_n,x_n,\xi _n)\rightarrow (p,x,\xi )\) as \(n\rightarrow \infty \), \(f_n(y)=\mathrm {e}^{-\xi _n'y}\), and \(f(y)=\mathrm {e}^{-\xi 'y}\) for \(y\in \mathbb {R}_+^C\). Then for any sequence such that \(y_n\rightarrow y\), we have \(f_n(y_n)\rightarrow f(y)\). (This property is referred to as “\(f_n\) continuously converges to \(f\).”) Since \(f_n\le 1, \left\{ f_n\right\} \) are uniformly \(\mu _i(p_n,x_n)\)-integrable, i.e.,

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\sup _n\int _{f_n>\alpha }f_n(y)\mu _i(\mathrm {d}y;p_n,x_n)=0. \end{aligned}$$

(Just take \(\alpha \ge 1\).) Therefore, by Theorem 9.2, we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\int f_n(y)\mu _i(\mathrm {d}y;p_n,x_n)=\int f(y)\mu _i(\mathrm {d}y;p,x). \end{aligned}$$

Hence, \(\int \mathrm {e}^{-\xi 'y}\mu _i(\mathrm {d}y;p,x)\) is continuous in \((p,x,\xi )\), and so is \(H^*(\xi ;p,x)\). \(\square \)

Proof of Proposition 3.2

\(\fancyscript{E}\) has a Bayesian general equilibrium because (3.14) is stronger than Assumption 2. Suppose that \(\fancyscript{E}\) has a non-degenerate equilibrium \((p,x,(f_i))\). Then

$$\begin{aligned} x_i=\int yf_i(y)\mu _i(\mathrm {d}y;p,x)\in \mathrm{cl }\mathrm{co }X_i(p,x). \end{aligned}$$

By market clearing and the nature of Lagrange multipliers, we have

$$\begin{aligned} \sum _{i=1}^In_i(x_i-e_i)=\bar{x}[f;\mu (p,x)]-\bar{e}\le 0 \end{aligned}$$

and \(p\cdot \sum _{i=1}^In_i(x_i-e_i)=0\). Then for any \(y=(y_i)\) with \(y_i\in X_i(p,x)\), by (3.14) we get

$$\begin{aligned} p\cdot \sum _{i=1}^In_i(y_i-e_i)\ge 0=p\cdot \sum _{i=1}^In_i(x_i-e_i), \end{aligned}$$

so \(p\cdot y\ge p\cdot x_i\) for all \(i\) and \(y\in X_i(p,x)\). Hence, by Definition 2.4, \((p,x)\) is a degenerate equilibrium.

Let \(\fancyscript{E}^n=\left\{ I,\left\{ n_i\right\} ,\left\{ e_i\right\} ,\left\{ \mu _i^n\right\} \right\} \) be an economy such that

$$\begin{aligned} \mu _i^n(B;p,x)=\mu _i\left( (1-1/n)^{-1}B;p,x\right) \end{aligned}$$

for any Borel set \(B\), that is, the prior \(\mu _i^n(p,x)\) is obtained by shrinking the domain of \(\mu _i(p,x)\) by \(1-1/n\) about the origin. Then the offer set is \(X_i^n(p,x)=\mathrm{supp }\mu _i^n(p,x)=(1-1/n)X_i(p,x)\), so

$$\begin{aligned} \sum _{i=1}^I n_i\inf \left\{ p\cdot y|y\in X_i^n(p,x)\right\}&=\left( 1-\frac{1}{n}\right) \sum _{i=1}^I n_i\inf \left\{ p\cdot y|y\in X_i(p,x)\right\} \\&<\sum _{i=1}^I n_i\inf \left\{ p\cdot y|y\in X_i(p,x)\right\} . \end{aligned}$$

Hence, by (3.14) we get

$$\begin{aligned} \sum _{i=1}^I n_i\inf \left\{ p\cdot (y-e_i)|y\in X_i^n(p,x)\right\} <\sum _{i=1}^I n_i\inf \left\{ p\cdot (y-e_i)|y\in X_i(p,x)\right\} =0. \end{aligned}$$

By Theorem 3.2, the economy \(\fancyscript{E}^n\) has a non-degenerate equilibrium \((p^n,(x_i^n),(f_i^n))\), where

$$\begin{aligned} x_i^n=\int yf_i^n(y)\mathrm {d}\mu _i(p^n,x^n)\in \mathrm{cl }\mathrm{co }X_i(p^n,x^n). \end{aligned}$$

Then by market clearing and the nature of Lagrange multipliers, after some algebra we obtain

$$\begin{aligned} \sum _{i=1}^In_i(x_i^n-e_i)\le 0~\text {and}~ p^n\cdot \sum _{i=1}^In_i(x_i^n-e_i)=0, \end{aligned}$$
(8.1)

where \(\bar{e}=\sum _{i=1}^I n_ie_i\) is the aggregate endowment. Since \(x_i^n\in \mathrm{cl }\mathrm{co }X_i(p^n,x^n)\subset X\subset \mathbb {R}_+^C\) and \(\left\{ x_i^n\right\} \) is bounded (\(\because \) market clearing), by taking a subsequence if necessary, we may assume \(x_i^n\) converges to some \(x_i\) as \(n\rightarrow \infty \). Since \(\varDelta ^{C-1}\) is compact, we may also assume \(p^n\rightarrow p\). Letting \(n\rightarrow \infty \) in (8.1), we get \(\sum _{i=1}^In_i(x_i-e_i)\le 0\) and \(p\cdot \sum _{i=1}^In_i(x_i-e_i)=0\). By Assumption 4, we have \(x_i\in \mathrm{cl }\mathrm{co }X_i(p,x)\), where \(x=(x_i)\). By the same argument as above, \((p,x)\) is a degenerate equilibrium. \(\square \)

Appendix 3: Mathematical results

Lemma 9.1

(Chebyshev’s inequality) If \(f,g:\mathbb {R}\rightarrow \mathbb {R}\) are increasing (decreasing) functions and \(X\) is a random variable, then

$$\begin{aligned} \mathrm{E }[f(X)g(X)]\ge \mathrm{E }[f(X)]\mathrm{E }[g(X)]. \end{aligned}$$

Proof

Let \(X'\) be an i.i.d. copy of \(X\). Since \(f,g\) are monotone, we have

$$\begin{aligned} (f(X)-f(X'))(g(X)-g(X'))\ge 0. \end{aligned}$$

Taking expectations of both sides, noting that \(X,X'\) are i.i.d., and rearranging terms, we obtain

$$\begin{aligned} \mathrm{E }[f(X)g(X)]\ge \mathrm{E }[f(X)]\mathrm{E }[g(X)]. \end{aligned}$$

\(\square \)

Clearly if one of \(f,g\) is increasing and the other is decreasing, the reverse inequality holds.

Theorem 9.1

(Generalized Kuhn–Tucker) Let \(X\) be a linear vector space, \(Z_1,Z_2\) normed spaces, \(\Omega \) a convex subset of \(X\), and \(P\) the positive cone in \(Z_1\). Assume that \(P\) contains an interior point.

Let \(f\) be a real-valued convex functional on \(\Omega \), \(G_1:\Omega \rightarrow Z_1\) a convex mapping, and \(G_2:X\rightarrow Z_2\) an affine mapping. Assume the existence of a point \(x_1\in \Omega \) for which \(G_1(x_1)<0\) (i.e., \(G_1(x_1 )\) is an interior point of \(N=-P\)) and \(G_2(x_1)=0\), and that 0 is an interior point of \(G_2(\Omega )\). Let

$$\begin{aligned} \mu _0=\inf f(x)~\text {subject to}~x\in \Omega ,~G_1(x)\le 0,~G_2(x)=0 \end{aligned}$$
(9.1)

and assume \(\mu _0\) is finite. Then there exist \(z_1^*\ge 0\) in \(Z_1^*\) and \(z_2^*\in Z_2^*\) such that

$$\begin{aligned} \mu _0=\inf _{x\in \Omega }\left[ f(x)+\left\langle G_1(x),z_1^*\right\rangle +\left\langle G_2(x),z_2^*\right\rangle \right] . \end{aligned}$$
(9.2)

Furthermore, if the infimum is achieved in (9.1) by \(x_0\in \Omega \), it is achieved by \(x_0\) in (9.2) and \(\left\langle G_2(x_0),z_2^*\right\rangle =0\).

Proof

Similar to Luenberger (1969), Theorem 1, p. 217. \(\square \)

Proposition 9.1

Let \((X,\fancyscript{B},\mu )\) be a measure space, where \(X\) is a topological space, \(\fancyscript{B}\) is the Borel \(\sigma \)-algebra, and \(\mu (X)>0\). Let \(T:X\rightarrow \mathbb {R}^C\) be measurable. Then,

$$\begin{aligned} f(\xi ):=\log \left( \int \mathrm {e}^{-\xi 'T(x)}\mu (\mathrm {d}x)\right) \end{aligned}$$

is convex and lower semi-continuous on \(\mathrm{dom }f\).Footnote 9 Furthermore, \(f\) is strictly convex if \(\dim T(\mathrm{supp }\mu )=C\).Footnote 10

Proof

See Proposition B.4 in Toda (2010). \(\square \)

Proposition 9.2

Let \((X,\fancyscript{B},\mu )\) be as in Proposition 9.1 and \(\phi :X\rightarrow \mathbb {R}\) be measurable. If \(\int \mathrm {e}^{t\phi (x)}\mu (\mathrm {d}x)<\infty \) for some \(t>0\), then

$$\begin{aligned} \lim _{t\rightarrow \infty }\frac{1}{t}\log \left( \int \mathrm {e}^{t\phi }\mathrm {d}\mu \right) =\mathrm{ess }\sup \phi . \end{aligned}$$
(9.3)

If \(\phi \) is upper semi-continuous, then (9.3) is equal to \(\sup \left\{ \phi (x)|x\in \mathrm{supp }\mu \right\} \).

Proof

Let \(E_n=\left\{ x\in X|\phi (x)\ge -n\right\} \). If \(\mu (E_n)=0\) for all \(n\), then

$$\begin{aligned} \mu (X)=\mu \left( \textstyle \bigcup _{n=1}^\infty E_n\right) =\lim _{n\rightarrow \infty }\mu (E_n)=0, \end{aligned}$$

which contradicts \(\mathrm{supp }\mu \ne \emptyset \). Therefore, \(\mu (E_n)>0\) for some \(n\). Letting

$$\begin{aligned} v=\mathrm{ess }\sup \phi =\sup \left\{ c|\mu (\left\{ x\in X|\phi (x)\ge c\right\} )>0\right\} , \end{aligned}$$

we have \(v>-\infty \).

Let us first prove (9.3) when \(v<\infty \). Define

$$\begin{aligned} X_+&=\left\{ x\in X|\phi (x)>v\right\} ,~X_-=\left\{ x\in X|\phi (x)\le v\right\} ,~\text {and}\\ X_n&=\left\{ x\in X\big |\phi (x)\ge v+\frac{1}{n}\right\} . \end{aligned}$$

Since \(X_+=\bigcup _{n=1}^\infty X_n\) and \(\mu (X_n)=0\) by the definition of \(v\), we have \(\mu (X_+)=0\). Obviously, \(X_\pm \) are disjoint and \(X_+\cup X_-=X\), so \(\mu (X_-)=\mu (X)>0\). Fix \(t_0>0\) such that \(\int \mathrm {e}^{t_0\phi (x)}\mathrm {d}\mu <\infty \). Then, for all \(t>0\) we obtain

$$\begin{aligned} \int \mathrm {e}^{t\phi (x)}\mathrm {d}\mu =\mathrm {e}^{tv}\int _{X_-}\mathrm {e}^{t(\phi (x)-v)}\mathrm {d}\mu . \end{aligned}$$
(9.4)

Denote the integral over \(X_-\) in (9.4) by \(I(t)\). Since \(\phi (x)\le v\) for \(x\in X_-\), for each \(x\in X_-\) the integrand \(\mathrm {e}^{t(\phi (x)-v)}\) is decreasing in \(t\), so \(I(t)\) is decreasing in \(t\). (In particular, \(0<I(t)<\infty \) for \(t\ge t_0\).) Hence, for \(t\ge t_0\) we obtain

$$\begin{aligned} \frac{1}{t}\log \left( \int \mathrm {e}^{t\phi (x)}\mathrm {d}\mu \right) =v+\frac{1}{t}\log I(t)\le v+\frac{1}{t}\log I(t_0). \end{aligned}$$
(9.5)

Letting \(t\rightarrow \infty \) in (9.5), we obtain

$$\begin{aligned} \limsup _{t\rightarrow \infty }\frac{1}{t}\log \left( \int \mathrm {e}^{t\phi (x)}\mathrm {d}\mu \right) \le v. \end{aligned}$$
(9.6)

To show the reverse inequality, take any \(\epsilon >0\) and let

$$\begin{aligned} A=\left\{ x\in X_-|\phi (x)\ge v-\epsilon \right\} . \end{aligned}$$

By assumption and the definition of \(X_\pm \), we have

$$\begin{aligned} \mu (A)=\mu (\left\{ x\in X|\phi (x)\ge v-\epsilon \right\} )>0. \end{aligned}$$

By taking a compact subset of \(A\) if necessary, we may assume \(0<\mu (A)<\infty \) since \(\mu \) is regular. Therefore, we obtain

$$\begin{aligned} \frac{1}{t}\log \left( \int \mathrm {e}^{t\phi (x)}\mathrm {d}\mu \right)&\ge \frac{1}{t}\log \left( \mathrm {e}^{t(v-\epsilon )}\int _A\mathrm {e}^{t(\phi (x)-v+\epsilon )}\mathrm {d}\mu \right) \nonumber \\&\ge v-\epsilon +\frac{1}{t}\log \mu (A). \end{aligned}$$
(9.7)

Letting \(t\rightarrow \infty \) in (9.7) and then \(\epsilon \rightarrow 0\), we obtain

$$\begin{aligned} \liminf _{t\rightarrow \infty }\frac{1}{t}\log \left( \int \mathrm {e}^{t\phi (x)}\mathrm {d}\mu \right) \ge v. \end{aligned}$$
(9.8)

(9.3) follows by (9.6) and (9.8).

If \(v=\infty \), let \(F_n=\left\{ x\in X|\phi (x)\ge n\right\} \). By the definition of \(v\), we have \(\mu (F_n)>0\). Then we obtain the same result as (9.7) with \(A\) replaced by \(F_n\) and \(v-\epsilon \) replaced by \(n\). Letting \(n\rightarrow \infty \) we get (9.3).

Finally let us show \(\mathrm{ess }\sup \phi =\sup \left\{ \phi (x)|x\in \mathrm{supp }\mu \right\} \) if \(\phi \) is upper semi-continuous. Let \(u=\sup \left\{ \phi (x)|x\in \mathrm{supp }\mu \right\} \). If \(u<\infty \), for all \(\epsilon >0\) there exists an \(x_0\in X\) such that \(u-\epsilon <\phi (x_0)\). Since \(\phi \) is upper semi-continuous, there exists an open neighborhood \(U\) of \(x_0\) such that \(x\in U\) implies \(\phi (x)>u-\epsilon \). Since \(\mu (U\cap X)>0\) by assumption, it follows that \(v\ge u-\epsilon \). Since \(\epsilon >0\) is arbitrary, we obtain \(v\ge u\). A similar reasoning holds for the case \(u=\infty \).

To show the reverse inequality, take any \(\epsilon >0\). By the definition of \(v\), we have \(\mu (\left\{ x\in X|\phi (x)\ge v-\epsilon \right\} )>0\). In particular, there exists an \(x_0\in \mathrm{supp }\mu \) such that \(\phi (x_0)\ge v-\epsilon \). Therefore,

$$\begin{aligned} u=\sup \left\{ \phi (x)|x\in \mathrm{supp }\mu \right\} \ge \phi (x_0)\ge v-\epsilon . \end{aligned}$$

Since \(\epsilon >0\) is arbitrary, we obtain \(u\ge v\). Therefore, \(u=v\). \(\square \)

Finally, I need a convergence theorem of Lebesgue integrals with varying measures. Let \(X\) be a locally compact second countable Hausdorff space (e.g., Euclidean space) with Borel \(\sigma \)-algebra \(\fancyscript{B}\). We say that \(f_n\) continuously converges to \(f\), denoted by \(f_n\rightarrow _c f\), if \(\lim f_n(x_n)\rightarrow f(x)\) for any \(x_n\rightarrow x\).Footnote 11

Theorem 9.2

Let \(\mu , \left\{ \mu _n\right\} \) be finite Borel measures on \(X\). Suppose that \(f_n\ge 0, f_n\rightarrow _c f, \mu _n\rightarrow \mu \) weakly, and \(\int f_n\mathrm {d}\mu _n<\infty \) for all \(n\). Then

$$\begin{aligned} \lim _{n\rightarrow \infty }\int f_n\mathrm {d}\mu _n=\int f\mathrm {d}\mu \end{aligned}$$

if and only if \(\left\{ f_n\right\} \) is uniformly \(\left\{ \mu _n\right\} \)-integrable, i.e.,

$$\begin{aligned} \lim _{\alpha \rightarrow \infty }\sup _n\int _{f_n>\alpha }f_n\mathrm {d}\mu _n=0. \end{aligned}$$

Proof

See Serfozo (1982), Theorem 3.5. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Toda, A.A. Bayesian general equilibrium. Econ Theory 58, 375–411 (2015). https://doi.org/10.1007/s00199-014-0849-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-014-0849-4

Keywords

JEL Classification

Navigation