Skip to main content
Log in

Credit risk in general equilibrium

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

Abstract

This paper contributes to the literature on default in general equilibrium. Borrowing and lending takes place via a clearing house (bank) that monitors agents and enforces contracts. Our model develops a concept of bankruptcy equilibrium that is a direct generalization of the standard general equilibrium model with financial markets. Borrowers may default in equilibrium and returns on loans are determined endogenously. Restricted to a special form of mean variance preferences, we derive a version of the capital asset pricing model with bankruptcy. In this case, we can characterize equilibrium prices and allocations and discuss implications for credit risk modeling.

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. There is a second important class of general equilibrium models allowing for default. Rather than modeling default consequences by penalties in this literature, all financial promises have to be secured by collateral in the form of assets or durable consumption goods. When a default occurs, collateral is transferred from the borrower to the lender. The key reference in this literature is Geanakoplos and Zame (2014).

  2. In addition to providing new results, the paper by Dana et al. (1999) contains an overview of this literature and clarifies all the different arbitrage notions used there. Instead of enumerating the rather long list of papers on arbitrage and general equilibrium, we refer the interested reader to this paper.

  3. We use the notation \(\vee \) and \(\wedge \) as the maximum and minimum operator. Applied to vectors, the operators give the component-wise maximum of minimum.

  4. Assume that there are \(S+1\) states of the world, each state occurring with probability \(\rho _{s}>0\), this would be the function \(u^{i}(x^{i})=-\frac{1 }{2} \sum _{s=0}^{S}\rho _{s}(\alpha ^{i}-x_{s}^{i})^{2}\). If we had a penalty function that gives a huge disutility to zero consumption, such as the logarithmic Bernoulli utility function, we would be back in the standard “no-bankruptcy” model.

  5. To obtain a bounded budget set, additional assumptions are necessary. As explained in the next section, one can either introduce a lower bound on bond trades through some kind of short-selling restriction as in Radner (1972) or appeal to a stricter notion of no arbitrage as in Werner (1987) and Dana et al. (1999).

  6. This condition is equivalent to \(\tau _{0}^{1}+\tau _{0}^{2}=0\).

  7. This is analogous to the general equilibrium, multigood financial market model, where agents have to correctly anticipate equilibrium goods prices at \(t=1\) when making their plans today (see Radner 1972).

  8. Strictly increasing means that for any \(x,x^{\prime }\in \mathbb {R}_{+}^{n}\) with \(x\ge x^{\prime }\) and \(x\ne x^{\prime }\), \(u^{i}(x)>u^{i}(x^{\prime })\).

  9. There is a literature starting from Werner (1987) and analyzed in depth in Dana et al. (1999) where stronger no-arbitrage notions together with additional assumptions on utility functions endogenously bound the choice set, so that the absence of arbitrage is necessary and sufficient for equilibrium existence when the choice set is unbounded. Since we want to focus attention on the bankruptcy clearing mechanism and the implications for the endogenous return on bonds, we choose the short-selling constraint approach.

  10. Such a formalization has been used in the literature in different versions by Modica et al. (1998), Sabarwal (2003), Araujo and Pascoa (2002). It is also close to the framework of Eisenberg and Noe (2001), which shows how one can extend our bankruptcy rule to many loan instruments and nonanonymous bankruptcy in a pure balance sheet mechanics framework. A bankruptcy occurs if agents cannot repay their due liabilities. In contrast to Zame (1993) and Dubey et al. (2005), we do not allow agents to strategically default on their loans. Agents will repay their debts as long as the value of their endowments and equity allows it. If liabilities exceed, this value a bankruptcy occurs.

  11. Define for any two vectors \(x,y\in \mathbb {R}^{n}\) the lattice operations \( x\wedge y:=(min(x_{1},y_{1}),\ldots ,min(x_{n},y_{n}))\) and \(x\vee y:=(max(x_{1},y_{1}),\ldots ,max(x_{n},y_{n}))\). By \(Y_{s}\), we denote the s-th row of the matrix \(Y\).

  12. Since the recovery rate is only defined when there is some trade in the bond, we define the recovery rate in cases where there is no bond trade as 1 by a continuous extension.

  13. Araujo and Pascoa (2002) have no utility penalties but short-selling constraints on the debt instruments, Sabarwal (2003) has \(T\) periods and no penalties but short-selling constraints on the debt instruments. Modica et al. (1998) study a model where agents can become bankrupt without penalty in states of the world of which they are ex ante unaware of. Obviously, all these models are closely related.

  14. Unlike in the general case discussed before, these preferences have a satiation point and thus the utility function is not strictly increasing on its whole domain. It is well known that one can obtain monotonicity on the relevant compact and convex set of state-contingent consumption by choosing for instance for all consumers satiation points outside the set of feasible allocations.

  15. Since we are using the probability inner product, the definition of \(\nabla u^i(\bar{x}^i)\) is

    $$\begin{aligned} \nabla u^i(\bar{x}^i) = \left( \frac{\partial u^i}{\partial x_0}(\bar{x} ^i_0), \left( \frac{1}{\rho _s} \frac{\partial u^i}{\partial x_s} (\bar{x}^i_s) \right) _{s \in S} \right) . \end{aligned}$$
  16. We would like to thank an anonymous referee of this journal for suggesting the greatly simplified proof of existence for an equilibrium in period \(t=0.\)

  17. Notice the change of normalization for the budget in \(t=0.\)

References

  • Alvarez, F., Jerman, U.: Efficiency, equilibrium and asset pricing with risk of default. Econometrica 64(4), 663–672 (2001)

    Google Scholar 

  • Araujo, A., Pascoa, M.: Bankruptcy in a model of unsecured claims. Econ. Theory 20, 455–481 (2002)

    Article  Google Scholar 

  • Araujo, A., Monteiro, P., Pascoa, M.: Infinite horizon, incomplete markets with a continuum of states. Math. Financ. 6, 119–132 (1996)

    Article  Google Scholar 

  • Araujo, A., Monteiro, P., Pascoa, M.: Incomplete markets, continuum of states and default. Econ. Theory 20(1), 205–213 (1998)

    Article  Google Scholar 

  • Azariadis, C., Kaas, L.: Endogenous credit limits with small default costs. J. Econ. Theory 148(2), 806–824 (2013)

    Article  Google Scholar 

  • Bloise, G., Reichlin, P.: Asset prices, debt constraints and inefficiency. J. Econ. Theory 146(4), 1520–1546 (2011)

    Article  Google Scholar 

  • Dana, R.A., Van, C.L., Magnien, F.: On the different notions of arbitrage and existence of equilibrium. J. Econ. Theory 87, 169–193 (1999)

    Article  Google Scholar 

  • Debreu, G.: Theory of Value. Cowles Foundation Monograph. Yale University Press, New Haven (1956)

    Google Scholar 

  • Diamond, D.: Financial intermediation and delegated monitoring. Rev. Econ. Stud. 51, 393–414 (1984)

    Article  Google Scholar 

  • Dubey, P., Geanakopols, J., Shubik, M.: Default and punishment in general equilibrium. Econometrica 73, 1–37 (2005)

    Article  Google Scholar 

  • Eisenberg, L., Noe, T.: Systemic risk in financial systems. Manage. Sci. 47(2), 236–249 (2001)

    Article  Google Scholar 

  • Gale, D., Hellwig, M.: Incentive-compatible debt contracts: the one-period problem. Rev. Econ. Stud. 52, 647–663 (1985)

    Article  Google Scholar 

  • Geanakoplos, J., Zame, W.: Collateral equilibrium, 1: a basic framework. Econ. Theory (2014). doi:10.1007/s00199-014-0803-5

  • Grandmont, J.M.: Temporary general equilibrium theory. Econometrica 45, 535–572 (1977)

    Article  Google Scholar 

  • Grandmont, J.M.: Money and Value. A Reconsideration of Classical and Neoclassical Monetary Theories. Cambridge University Press, Cambridge (1985)

    Google Scholar 

  • Green, J.: Temporary general equilibrium in a sequential trading model with spot and future transactions. Econometrica 41, 1103–1123 (1973)

    Article  Google Scholar 

  • Hellwig, C., Lorenzoni, G.: Bubbles and self-enforcing debt. Econometrica 77(4), 1137–1164 (2009)

    Article  Google Scholar 

  • Kehoe, T., Levine, D.: Debt constrained asset markets. Rev. Econ. Stud. 60(4), 865–888 (1993)

    Article  Google Scholar 

  • Kocherlakota, N.: Injecting rational bubbles. J. Econ. Theory 142(1), 218–232 (2008)

    Article  Google Scholar 

  • Luttmer, E.G.: Asset pricing in economies with frictions. Econometrica 64, 1439–1467 (1996)

    Article  Google Scholar 

  • Magill, M., Quinzii, M.: Theory of Incomplete Markets, 1st edn. MIT Press, Cambridge, MA (1996)

    Google Scholar 

  • Magill, M., Quinzii, M.: Which improves welfare more: a nominal or an indexed bond? Econ. Theory 10, 1–38 (1997)

    Article  Google Scholar 

  • Magill, M., Quinzii, M.: Infinite horizon CAPM. Econ. Theory 15, 103–138 (2000)

    Article  Google Scholar 

  • McNeil, A., Frey, R., Embrechts, P.: Quantitative Risk Management. Princeton University Press, Princeton (2005)

    Google Scholar 

  • Modica, S., Rusticini, A., Tallon, J.M.: Unawareness and Bankruptcy: a general equilibrium model. Econ. Theory 12, 259–292 (1998)

    Article  Google Scholar 

  • Nielsen, L.T.: Market equilibrium with short selling. Rev. Econ. Stud. 56, 467–473 (1989)

    Article  Google Scholar 

  • Radner, R.: Existence of equilibrium of plans, prices, and price expectations in a sequence of markets. Econometrica 40, 289–303 (1972)

    Article  Google Scholar 

  • Sabarwal, T.: Competitive equilibria with incomplete markets and endogenous bankruptcy. Contrib. Theor. Econ. 3, 1–40 (2003)

  • Townsend, R.: Optimal contracts and competitive markets with costly state verification. J. Econ. Theory 22, 265–293 (1979)

    Article  Google Scholar 

  • Werner, J.: Arbitrage and the existence of competitive equilibrium. Econometrica 55(6), 1403–1418 (1987)

    Article  Google Scholar 

  • Zame, W.: Efficiency and the role of default when security markets are incomplete. Am. Econ. Rev. 83, 1142–1164 (1993)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Martin Summer.

Additional information

We thank Toni Ahnert, Thomas Breuer, Helmut Elsinger, Philipp Hartmann, and Paul Söderlind for discussions. We would also like to thank an anonymous referee for his detailed feedback and extremely helpful comments. Furthermore, we would like to thank seminar audiences at the Austrian Central Bank (OeNB), the European Meeting of the Econometric Society in Oslo, University of Heidelberg, London School of Economics, University of Vienna, Swiss National Bank, the Macroprudential Research Network of the European Central Bank (MaRS), and the Research Task Force of the Basel Committee (RTF). Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of OeNB.

Appendix

Appendix

1.1 Proof of Proposition 1

Our proof consists of three steps:Footnote 16 In Step (i), we prove some preliminary results. In particular, we show that asset returns induce future state-contingent consumption, which can be evaluated by an indirect utility function. Moreover, a consumer will not buy and sell bonds simultaneously, the lower bound on borrowing implies a bounded set of feasible asset trades, and the bankruptcy clearing mechanism implies a strictly positive return of the bond in all states. Step (ii) provides an existence proof for an asset market equilibrium with a bankruptcy clearing mechanism in period \(t=0.\) For this proof, we can adapt standard arguments of general equilibrium theory. Finally, in Step (iii), we show that the bankruptcy clearing mechanism guarantees nonnegative consumption in all states.

1.1.1 Preliminary results

The indirect utility function Consider the price simplex in \(\mathbb {R}_{+}^{J+2},\)

$$\begin{aligned} Q:=\left\{ (p_{0},q_{b},q_{e})\in \mathbb {R}_{+}^{J+2}|\ p_{0}+q_{b}+q_{e} \mathbf {1}_{J}=1\right\} \!. \end{aligned}$$

For given \(r_{{\varvec{1}}}\in \) \((0,1]^{S}\)and \(\left( p_{0},q_{b},q_{e}\right) \in Q,\) each consumer \(i\in I\) maximizes \( u^{i}(x_{0}^{i},x_{\mathbf {1}}^{i})\) by choosing asset trades \( (z_{b+}^{i},z_{b-}^{i},z_{e}^{i})\) and a consumption allocation \( (x_{0}^{i},x_{{\varvec{1}}}^{i})\in X^{i}\) subject toFootnote 17

$$\begin{aligned} \begin{array}{llll} (a) &{} p_{0}\left( x_{0}^{i}-\omega _{0}^{i}\right) &{} \le &{} -q_{b}z_{b+}^{i}+q_{b}z_{b-}^{i}-q_{e}z_{e}^{i}, \\ (b) &{} x_{{\varvec{1}}}^{i} - \omega _{{\varvec{1}}}^{i}-Y\delta ^{i} &{} \le &{} r_{{\varvec{1}}}z_{b+}^{i} - \mathbb {1}z_{b-}^{i}+Yz_{e}^{i},\\ (c) &{} z_{b+}^i&{} \ge &{} 0 \\ (d) &{} z_{b-}^{i} &{} \ge &{} 0, \\ (e) &{} z_{b-}^{i} &{} \le &{} \kappa , \\ (f) &{} z_{e}^{i} &{} \ge &{} -\delta ^{i}, \end{array} \end{aligned}$$

where the constraint \((e)\) follows from A5 in Proposition 1.

Since \(u^{i}\) is strictly increasing in its arguments (Proposition 1, A2), condition \((b)\) is binding for an optimal choice \((x_{0}^{i},x_{{\varvec{1}}}^{i},z_{b+}^{i}, z_{b-}^{i}, z_{e}^{i}),\) i.e.,

$$\begin{aligned} x_{{\varvec{1}}}^{i}=\omega _{{\varvec{1}}}^{i} + Y\left( \delta ^{i}+z_{e}^{i}\right) +r_{{\varvec{1}}}z_{b+}^{i} - \mathbb {1}z_{b-}^{i}. \end{aligned}$$
(8)

Moreover, for \(r_{{\varvec{1}}}\ne 1\!\!1,\) the consumer will never take a short position \(z_{b-}^{i}>0\) and a long position \(z_{b+}^{i}>0\) simultaneously. These observations allow us to restrict attention to net trades in bonds \(z_{b}^{i}:= \left( z_{b+}^{i}-z_{b-}^{i}\right) =\left( z_{b}^{i}\vee 0\right) +\left( z_{b}^{i}\wedge 0\right) \in \mathbb {R}.\) Hence, we can write \(z^{i}=(z_{b}^{i},z_{e}^{i})\) for the net asset trades of a consumer \(i\in I.\)

For each consumer \(i\in I,\) consider the consumption set in period 0, \( X_{0}^{i}:=\mathbb {R}_{+},\) and the set of feasible net asset trades,\(\ Z^{i}:=[-\kappa ,\infty )\times \left( \mathbb {R}_{+}^{J} -\left\{ \delta ^{i}\right\} \right) \).

For any \(r_{1}\in (0,1]^{S}\), Eq. (8) allows us to define the following indirect utility function \(v^{i}:X_{0}^{i}\times Z^{i}\rightarrow \mathbb {R}\)

$$\begin{aligned} v^{i}(x_{0}^{i},z_{b}^{i},z_{e}^{i}; r_{{\varvec{1}}}):=u^{i}\left( x_{0}^{i}, \omega _{{\varvec{1}}}^{i}+ Y\left( \delta ^{i}+z_{e}^{i}\right) +r_{{\varvec{1}}}\left( z_{b}^{i}\vee 0\right) +\mathbb {1}\left( z_{b}^{i}\wedge 0\right) \right) \end{aligned}$$

By Assumption A2, the indirect utility function \(v^{i}\) is a continuous function. Moreover, \(v^{i}\) is strictly increasing and concave in \(\left( x_{0}^{i},z_{b}^{i},z_{e}^{i}\right) \).

Feasible allocations An allocation for a consumer \(i\in I\ \)in period \(t=0\) is a vector of period 0 consumption and of asset trades \((x_{0}^{i},z^{i}) \in X_{0}^{i}\times Z^{i}.\) Denote by \(A:=\prod _{i\in I}\left( X_{0}^{i}\times Z^{i}\right) \subset \mathbb {R}^{J+2}\) the set of allocations for the economy. Denote \((x_{0}^{i},z^{i})_{i\in I}\) by \((x_0,z)\). An allocation \( (x_0,z) \in A\) is weakly feasible if \(\sum _{i\in I}\left( x_{0}^{i}-\omega _{0}^{i},z^{i}\right) \le \left( 0,0\right) \) holds. Since \(X_{0}^{i}\times Z^{i}\) is bounded below for all \(i\in I\), the set of weakly feasible allocations

$$\begin{aligned} F:=\left\{ (x_{0}^{i},z^{i})_{i\in I}\in A\ |\ \sum _{i\in I}\left( x_{0}^{i}-\omega _{0}^{i},z^{i}\right) \le \left( 0,0\right) \right\} \end{aligned}$$

is a compact set.

The bankruptcy clearing mechanism The bankruptcy clearing mechanism (compare Eq. 4) is a function \(\rho :A\rightarrow (0,1]^{S}\) defined by its component functions

$$\begin{aligned} \rho _{s}(x_{0},z)=\left\{ \begin{array}{lll} \frac{\sum \nolimits _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i}+z_{e}^{i}))}{\sum \nolimits _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) }. &{} \hbox {for} &{} \sum \nolimits _{i=1}^{I}\left( -z_{b}^{i} \vee 0\right) >0\\ 1 &{} \hbox {for} &{} \sum \nolimits _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) =0 \end{array}\right. \! , \end{aligned}$$

\(\rho _{s}(x_{0},z)\) is a continuous function on \(A.\) Notice that, by Assumption A3, \(\omega _{s}^{i}>0\) for all \(s\in S\) and all \(i\in I.\) Hence, \(\left( -z_{b}^{i}\vee 0\right) \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i}+z_{e}^{i}))=-z_{b}^{i}\) for \(0<-z_{b}^{i}<\omega _{s}^{i}.\) This implies that \(\rho _{s}(x_{0},z)=1\) in a neighborhood of \(z_{b}=0.\) Thus, \(\rho _{s}(x_{0},z)\) is well defined and continuous at \(z_{b}=0.\) Moreover, \(\frac{\left( -z_{b}^{i}\vee 0\right) \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i}+z_{e}^{i}))}{\left( -z_{b}^{i}\vee 0\right) }\) is an increasing function of \(z_{b}^{i}\) with a minimum \(\frac{\kappa \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i} +z_{e}^{i}))}{\kappa } \ge \frac{\kappa \wedge \omega _{s}^{i}}{\kappa }\ \)at \(z_{b}^{i}=-\kappa .\) Therefore, we have

$$\begin{aligned} \rho _{s}(x_{0},z)&= \frac{\sum _{i=1}^{I}\left( -z_{b}^{i} \vee 0\right) \wedge (\omega _{s}^{i}+Y_{s} (\delta ^{i} +z_{e}^{i}))}{\sum _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) } \\&= \sum _{i=1}^{I}\left[ \frac{\left( -z_{b}^{i} \vee 0\right) \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i}+z_{e}^{i}))}{\left( -z_{b}^{i}\vee 0\right) } \right] \frac{\left( -z_{b}^{i}\vee 0\right) }{\sum _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) } \\&\ge \sum _{i=1}^{I}\left[ \frac{\kappa \wedge \omega _{s}^{i}}{\kappa } \right] \frac{\left( -z_{b}^{i}\vee 0\right) }{\sum _{i=1}^{I}\left( -z_{b}^{i}\vee 0\right) } \\&\ge \underset{i\in I}{\min } \left[ \frac{\kappa \wedge \omega _{s}^{i}}{ \kappa }\right] \underset{=1}{\underbrace{\sum _{i=1}^{I}\frac{\left( -z_{b}^{i}\vee 0\right) }{\sum _{i=1}^{I} \left( -z_{b}^{i}\vee 0\right) }}}\\&= \underset{i\in I}{\min } \left[ \frac{\kappa \wedge \omega _{s}^{i}}{ \kappa }\right] =:\underline{\rho }_{s}>0. \end{aligned}$$

Hence, for all \(s\in S,\ \ \rho _{s}(x_{0},z)\) is bounded below by a strictly positive value \(\underline{\rho }_{s}\). Denote by \(R:=\prod _{s\in S} \left[ \underline{\rho }_{s},1\right] \subset (0,1]^{S}\) the set of feasible return rates for the bond.

1.1.2 Existence of equilibrium in \(\mathbf {t=0}\)

Choose a compact and convex set \(K\subset \mathbb {R}^{J+2}\) such that \( (x_{0}^{i},z^{i})_{i\in I}\in F\) implies \((x_{0}^{i},z^{i}) \in \text {int}K \) for all \(i\in I.\)

For any \(\left( p_{0},q_{b},q_{e}\right) \in Q,\) denote by

$$\begin{aligned} \mathbb {B}_{c}^{i}(p_{0},q_{b},q_{e}):=\left\{ (x_{0}^{i},z_{b}^{i},z_{e}^{i}) \in \left( X_{0}^{i}\times Z^{i}\right) \cap K|\ p_{0}\left( x_{0}^{i}\!-\!\omega _{0}^{i}\right) \!+\!q_{b}z_{b}^{i}\!+\!q_{e}z_{e}^{i}\le 0\right\} \end{aligned}$$

the bounded budget correspondence for \(t=0\ \) restricted to the compact set \(K.\)

Lemma 1

\(\mathbb {B}_{c}^{i}: Q\rightarrow \left( X_{0}^{i}\times Z^{i}\right) \cap K,\) is a compact- , convex-valued, and continuous correspondence.

Proof

Since \(\mathbb {B}_{c}^{i}(p_{0},q_{b},q_{e}) \subset K\) for all \((p_{0},q_{b},q_{e}) \in Q,\) the budget set is compact. It is obviously convex. Since \(\omega _{0}^{i}>0,\) \(\kappa >0\) and \(\delta ^{i}\in \mathbb {R}_{++}^{J},\) continuity follows by standard arguments, e.g., Debreu (1956), p. 63. \(\square \)

For \(r_{{\varvec{1}}}\in R\) and \(\left( p_{0},q_{b},q_{e}\right) \in Q\), define the bounded demand correspondence \(f_{c}^{i}:Q\times R\rightarrow \mathbb {R}^{J+2}\) by

$$\begin{aligned} f_{c}^{i}(p_{0},q_{b},q_{e};r_{{\varvec{1}}}):=\arg \max \left\{ v^{i}(x_{0}^{i},z_{b}^{i},z_{e}^{i};r_{{\varvec{1}}})|\ \left( x_{0}^{i},z_{b}^{i},z_{e}^{i}\right) \in \mathbb {B} _{c}^{i}(p_{0},q_{b},q_{e})\right\} . \end{aligned}$$

Lemma 2

\(f_{c}^{i}(p_{0},q_{b},q_{e};r_{{\varvec{1}}})\) is a nonempty, compact- and convex-valued, u.h.c. correspondence from \(Q\times R\) into \(\mathbb {R}^{J+2}\).

Proof

The indirect utility \(v^{i}\) is a continuous function on \(\mathbb {R}_{+}\times \mathbb {R}^{J+1}\times R.\) By Lemma 1, \(\mathbb {B}_{c}^{i}(p_{0},q_{b},q_{e})\) is a compact-, convex-valued, and continuous correspondence on \(Q\). Hence, by the maximum theorem, the demand correspondence \(f_{c}^{i}(p_{0},q_{b},q_{e}; r_{{\varvec{1}}})\) is nonempty, compact-valued, and u.h.c. for all \( (p_{0},q;r_{{\varvec{1}}})\in Q\times R\). Since \(v^{i}\) is concave, it follows by standard arguments that \(f_{c}^{i}(p_{0},q_{b},q_{e};r_{{ \varvec{1}}})\) is convex-valued. \(\square \)

Lemma 3

Consider a vector of prices \((p_{0},q_{b},q_{e})\in Q\) and an allocation of optimal choices \((x_{0}^{i},z^{i})\in f_{c}^{i}(p_{0},q_{b},q_{e};r_{{\varvec{1}}})\) for all \(i\in I\) which is weakly feasible, \((x_{0}^{i},z^{i})_{i\in I}\in F.\) Then, all prices must be strictly positive, \((p_{0},q_{b},q_{e})>>0.\)

Proof

Suppose there was a price \(q_{e}^{j}=0,\) since the indirect utility is strictly increasing in \((x_{0}^{i},z^{i}),\) \(z_{e_{j}}^{i}\) must be in the boundary of \(K.\) Since \((x_{0}^{i},z^{i})_{i\in I}\in F\) implies \( (x_{0}^{i},z^{i})\in \text {int}K\) for all \(i\in I,\) this contradicts \( (x_{0}^{i},z^{i})_{i\in I}\in F.\) A similar argument holds for \(p_{0}=0\) and \(q_{b}=0.\) \(\square \)

Lemma 3 shows also that a weakly feasible allocation of optimal choices cannot be in the boundary of \(K.\) Hence, for a feasible allocation, the constraints of \(K\) will never be binding.

Define the individual excess demand correspondence \(\zeta ^{i}:Q\times R\rightarrow \mathbb {R}^{J+2}\) as

$$\begin{aligned} \zeta ^{i}(p_{0},q;r_{{{\varvec{1}}}}):= f_{c}^{i}(p_{0},q;r_{{{ \varvec{1}}}})- \left\{ (\omega _{0}^{i},0,0)\right\} . \end{aligned}$$

Obviously, \(\zeta ^{i}\) inherits all relevant properties of \(f_{c}^{i},\) in particular, \(\zeta ^{i}\) is a nonempty, compact- and convex-valued u.h.c. correspondence. Since \(f_{c}^{i}(p_{0},q;r_{{{\varvec{1}}}})\subseteq \mathbb {B}_{c}^{i}(p_{0},q),\) individual excess demand correspondences \( \zeta ^{i}\) are bounded below by \((-\omega _{0}^{i},-\kappa ,-\delta ^{i}).\) Consequently, the aggregate excess demand correspondence \(\sum _{i\in I}\zeta ^{i}(p_{0},q;r_{{{\varvec{1}}}})\) is bounded below by \(\left( -\sum _{i\in I}\omega _{0}^{i},-|I|\kappa , -\sum _{i\in I}\delta ^{i}\right) .\)

We denote by \(\zeta :Q\times R\rightarrow K^{I},\)

$$\begin{aligned} \zeta (p_{0},q;r_{{\varvec{1}}}):=\prod \limits _{i\in I}\zeta ^{i}(p_{0},q;r_{{\varvec{1}}}), \end{aligned}$$

the Cartesian product of the individual excess demand correspondences. By Lemma 2, \(\zeta (p_{0},q;r_{{\varvec{1}}})\) is a nonempty, compact-, and convex-valued u.h.c. correspondences on \(Q\times R\) as a product of correspondences with these properties.

For any excess demand vector \((x_{0},z):=(x_{0}^{i},z^{i})_{i\in I}\in K^{|I|},\) let

$$\begin{aligned} \mu (x_{0},z):=\arg \max \left\{ p_{0} \sum _{i=1}^{I}\left( x_{0}^{i}-\omega _{0}^{i}\right) +q_{b}\sum _{i=1}^{I}z_{b}^{i}+q_{e} \sum _{i=1}^{I}z_{e}^{i}|~(p_{0},q_{b},q_{e})\in Q\right\} \end{aligned}$$

be the set of prices in \(Q\) that maximize the value of the excess demand\(.\) Since \(Q\) is compact and \(p_{0} \sum _{i=1}^{I}\left( x_{0}^{i} -\omega _{0}^{i}\right) +q_{b}\sum _{i=1}^{I}z_{b}^{i}+q_{e} \sum _{i=1}^{I}z_{e}^{i}\) is a continuous function on \(Q,\ \mu (x_{0},z)\) is not empty. By the maximum theorem, \(\mu :K^{|I|}\) \(\rightarrow Q\) is a compact-, convex-valued, and u.h.c. correspondence on \(K^{|I|}.\)

For each \((p_{0},q_{b},q_{e},r_{{\varvec{1}}},(x_{0},z))\in Q\times R\times K^{|I|}\), define the correspondence

$$\begin{aligned} \varPhi (p_{0},q_{b},q_{e},r_{{\varvec{1}}}, (x_{0},z)):=\mu (x_{0},z)\times \left\{ \rho (x_{0},z)\right\} \times \zeta (p_{0},q_{b},q_{e},r_{{\varvec{1}}}). \end{aligned}$$

The correspondence \(\varPhi \) is a mapping \(Q\times R\times K^{|I|}\rightarrow \) \(Q\times R\times K^{|I|}.\) The correspondence \(\varPhi \) is nonempty, compact-valued, convex-valued, and u.h.c., as a Cartesian product of correspondences with these properties. Hence, by Kakutani’s fixed point theorem, there is \((p_{0}^{*},q_{b}^{*},q_{e}^{*},r_{{\ \varvec{1}}}^{*},\left( x_{0}^{*},z^{*}\right) ) \in \varPhi (p_{0}^{*},q_{b}^{*},q_{e}^{*},r_{{\varvec{1}}}^{*},\left( x_{0}^{*},z^{*}\right) ).\)

By construction of \(\mu ,\)

$$\begin{aligned}&p_{0}^{*}\sum _{i=1}^{I} \left( x_{0}^{*i} -\omega _{0}^{i}\right) +q_{b}^{*}\sum _{i=1}^{I}z_{b}^{*i}+q_{e}^{*}\sum _{i=1}^{I}z_{e}^{*i}\ge p_{0} \sum _{i=1}^{I}\left( x_{0}^{*i}-\omega _{0}^{i}\right) \\&\quad +q_{b}\sum _{i=1}^{I}z_{b}^{*i}+q_{e}\sum _{i=1}^{I}z_{e}^{*i} \end{aligned}$$

for all \((p_{0},q_{b},q_{e})\in Q.\) Moreover, by the budget constraint,

$$\begin{aligned} p_{0}^{*}\left( x_{0}^{*i} -\omega _{0}^{i}\right) +q_{b}^{*}z_{b}^{*i}+q_{e}^{*}z_{e}^{*i}\le 0 \end{aligned}$$

for all \(i\in I.\) Hence,

$$\begin{aligned} p_{0}\sum _{i=1}^{I}\left( x_{0}^{*i} -\omega _{0}^{i}\right) +q_{b}\sum _{i=1}^{I}z_{b}^{*i}+q_{e} \sum _{i=1}^{I}z_{e}^{*i}\le 0 \end{aligned}$$

for all \((p_{0},q_{b},q_{e})\in Q.\) This implies

$$\begin{aligned} \sum _{i\in I}\left( x_{0}^{*i}-\omega _{0}^{i}\right) \le 0,\quad \sum _{i\in I}z^{*i}\le 0. \end{aligned}$$

Therefore, the allocation \((x_{0}^{*i},z^{*i})_{i\in I}\) is weakly feasible, i.e., \((x_{0}^{*},z^{*})\in F.\)

Since \((x_{0}^{*i},z^{*i})\in f_{c}^{i}(p_{0}^{*},q_{b}^{*},q_{e}^{*},r_{{\varvec{1}}}^{*})\) for all \(i\in I\) and \((x_{0}^{*i},z^{*i})_{i\in I}\) is a weakly feasible allocation, it follows from Lemma 3 that all prices must be strictly positive, \(\left( p_{0}^{*},q_{b}^{*}, q_{e}^{*}\right) >>0. \) Moreover, since the indirect utility is strictly increasing, the budget constraint must be binding for all \(i\in I,\) i.e., \(p_{0}^{*}\left( x_{0}^{^{*i}}-\omega _{0}^{i}\right) +q_{b}^{*}z_{b}^{*i}+q_{e}^{*}z_{e}^{*i}=0.\) Together with weak feasibility, this implies market clearing,

$$\begin{aligned} \sum _{i\in I} \left( x_{0}^{*i} -\omega _{0}^{i}\right) =0,\quad \sum _{i\in I}z^{*i}=0. \end{aligned}$$

1.1.3 Existence of equilibrium in \(t=1\)

It remains to show that in a bankruptcy equilibrium

  1. (i)

    \(r_{{\varvec{1}}}^{*}=\rho (x_{0}^{*},z^{*})\ \) and

  2. (ii)

    \((x_{0}^{*i},z^{*i})\in f_{c}^{i}(p_{0}^{*},q_{b}^{*},q_{e}^{*}, r_{{\varvec{1}}}^{*})\ \)for all \(i\in I \) with \(\sum _{i\in I}z^{*i}=0,\)

consumption will be nonnegative in every state \(s\in S\), i.e., condition (iii) of the definition of a bankruptcy equilibrium

$$\begin{aligned} \sum _{i=1}^{I}\left( x_{s}^{*i}\vee 0\right) = \sum _{i=1}^{I}(\omega _{s}^{i}+Y_{s}\delta ^{i}) \end{aligned}$$

for all \(s\in S\) is also satisfied.

Recall that, for an equilibrium price system \((p_{0}^{*},q_{b}^{*},q_{e}^{*},r_{{\varvec{1}}}^{*})\) and an equilibrium allocation \((x_{0}^{*i},z^{*i})\in f_{c}^{i}(p_{0}^{*},q_{b}^{*},q_{e}^{*},r_{{\varvec{1}}}^{*}),\) by Eq. 8,

$$\begin{aligned} x_{s}^{*i}=\omega _{s}^{i} +Y_{s}\left( \delta ^{i}+z_{e}^{*i}\right) +r_{s}(z_{b}^{*i}\vee 0)+(z_{b}^{*i}\wedge 0) \end{aligned}$$

and, by Eq. 4,

$$\begin{aligned} r_{{\varvec{1}}}^{*}=\rho (x_{0}^{*},z^{*}):=\frac{ \sum _{i=1}^{I}\left( -z_{b}^{*i}\vee 0\right) \wedge (\omega _{s}^{i}+Y_{s} (\delta ^{i}+z_{e}^{*i}))}{\sum _{i=1}^{I}(-z_{b}^{*i}\vee 0)} \end{aligned}$$

for all \(s\in S\) hold. These conditions imply nonnegative consumption in all states.

Lemma 4

If for all \(i\in I\ \) and all \(s\in S\) Eqs. 8 and 4 hold, then

$$\begin{aligned} \sum _{i=1}^{I}z_{b}^{*i}=0\quad \text { and } \quad \sum _{i=1}^{I}z_{e}^{ *i}=0 \end{aligned}$$

imply

$$\begin{aligned} \sum _{i=1}^{I}\left( x_{s}^{*i}\vee 0\right) = \sum _{i=1}^{I}(\omega _{s}^{i}+Y_{s}\delta ^{i}) \end{aligned}$$

for all \(s\in S.\)

Proof

For notational convenience, we drop the reference \(^{*}\) to the equilibrium values. Consider an arbitrary \(s\in S.\) By Eq. 8, summing over all consumers \(i\in I,\) one obtains

$$\begin{aligned} \sum _{i=1}^{I}\left( x_{s}^{i} -\omega _{s}^{i}- Y_{s}\delta ^{i}\right) =\sum _{i=1}^{I}\left[ r_{s}(z_{b}^{i}\vee 0)+(z_{b}^{i}\wedge 0)+Y_{s}z_{e}^{i}\right] , \end{aligned}$$

which is equivalent to

$$\begin{aligned} \sum _{i=1}^{I}\left[ (x_{s}^{i}\vee 0) - \omega _{s}^{i}-Y_{s}\delta ^{i} \right] =\sum _{i=1}^{I}\left[ r_{s}(z_{b}^{i}\vee 0)+(z_{b}^{i}\wedge 0)+Y_{s}z_{e}^{i}-(0\wedge x_{s}^{i})\right] . \end{aligned}$$

The asset market equilibrium conditions

$$\begin{aligned} \sum _{i=1}^{I}z_{b}^{i}=0 \quad \text { and } \quad \sum _{i=1}^{I}z_{e}^{i}=0 \end{aligned}$$

imply

$$\begin{aligned}&\sum _{i=1}^{I}\left[ r_{s}(z_{b}^{i}\vee 0)+(z_{b}^{i}\wedge 0)+Y_{s}z_{e}^{i}-(x_{s}^{i}\wedge 0)\right] \\&\quad =r_{s}\underset{=-\sum _{i=1}^{I}(z_{b}^{i}\wedge 0)}{\underbrace{ \sum _{i=1}^{I}(z_{b}^{i}\vee 0)}}+\sum _{i=1}^{I}(z_{b}^{i}\wedge 0)-\sum _{i=1}^{I}(x_{s}^{i}\wedge 0)+Y_{s}\underset{=0}{\underbrace{ \sum _{i=1}^{I}z_{e}^{i}}} \\&\quad =r_{s}\sum _{i=1}^{I}(-z_{b}^{i}\vee 0) +\sum _{i=1}^{I}(z_{b}^{i}\wedge 0)-\sum _{i=1}^{I}(x_{s}^{i}\wedge 0). \end{aligned}$$

Denote by \(I^{+}:=\{i\in I|\ z_{b}^{i}\ge 0\}\ \) the set of consumers who are not borrowing, by \(I_{s}^{-}:=\{i\in I|\ z_{b}^{i}<0,\ x_{s}^{i}\ge 0\}\) the set of loan takers who are solvent, and by \(I_{i}^{-}:\{i\in I|\ z_{b}^{i}<0,\ x_{s}^{i}<0\}\) consumers who are insolvent, then substituting Eq. 4 for \(r_{s}\) yields

$$\begin{aligned}&r_{s}\sum _{i=1}^{I}(-z_{b}^{i}\vee 0) +\sum _{i=1}^{I}(z_{b}^{i}\wedge 0)-\sum _{i=1}^{I}(x_{s}^{i}\wedge 0) \\&\quad =r_{s}\sum _{i\in I^{-}}(-z_{b}^{i}\vee 0)+\sum _{i\in I^{-}}(z_{b}^{i}\wedge 0)-\sum _{i\in I^{-}}(x_{s}^{i}\wedge 0) \\&\quad =\sum _{i\in I^{-}}\left[ \left( -z_{b}^{i}\right) \wedge (\omega _{s}^{i}+Y_{s}(\delta ^{i}+z_{e}^{i}))\right] -\sum _{i\in I^{-}}(-z_{b}^{i})-\sum _{i\in I^{-}}(x_{s}^{i}\wedge 0) \\&\quad =\sum _{i\in I_{i}^{-}}\left[ \omega _{s}^{i} +Y_{s}(\delta ^{i}+z_{e}^{i})) \right] -\sum _{i\in I_{i}^{-}}(-z_{b}^{i}) -\sum _{i\in I_{i}^{-}}(x_{s}^{i}) \\&\quad =\sum _{i\in I_{i}^{-}} \left[ \left( \omega _{s}^{i}+Y_{s}\left( \delta ^{i}+z_{e}^{i}\right) +z_{b}^{i}\right) -x_{s}^{i})\right] =0, \end{aligned}$$

since \(x_{s}^{i}<0\) implies \(x_{s}^{i}= \omega _{s}^{i}+Y_{s}\left( \delta ^{i}+z_{e}^{i}\right) +z_{b}^{i}<0.\ \) Hence, \(\sum _{i=1}^{I}\left( x_{s}^{i}\vee 0\right) =\sum _{i=1}^{I}(\omega _{s}^{i} +Y_{s}\delta ^{i})\). \(\square \)

1.2 Proof of Proposition 2

We partition the matrix

$$\begin{aligned} T= \left[ \begin{array}{l@{\quad }l@{\quad }l} -q_b &{} q_b &{} -q_{e}\\ r_{\varvec{1}} &{} - 1\!\!1 &{} Y \end{array}\right] \end{aligned}$$

and the portfolio vector \(z^i \in Z^i\) into long-bond, short-bond, and equity trades by \(T=(T_{b+},T_{b-},T_e)\) and \(z^i=(z^i_{b+},z^i_{b-},z^i_e)^T \), respectively. Using Lagrange multipliers \(\pi ^i\in \mathbb {R}^{S+1}\), \(\sigma ^i_{b+} \ge 0\), and \(\sigma ^i_{b-}\ge 0\), and assuming we are at an interior solution (\(\delta ^i\) are such that short sales constraints on equity are not binding at equilibrium, and parameters are such that \(x^i_0 > 0\) at equilibrium), the KKT conditions for the minimization of the Lagrange function

$$\begin{aligned} L^i(x^i, z^i, \pi ^i, \sigma ^i_{b+}, \sigma ^i_{b-})= -u^i(x^i)+ \langle \pi ^i,x^i-\omega ^i- e^i-Tz^i\rangle -\sigma ^i_{b+} z^i_{b+} -\sigma ^i_{b-} z^i_{b-} \end{aligned}$$

imply that

$$\begin{aligned} \langle \nabla u^i(\bar{x}^i),T_e \rangle&= (0,\ldots ,0),\\ \langle \nabla u^i(\bar{x}^i),T_{b+} \rangle&= -\sigma ^i_{b+}\le 0,\text {and}\\ \langle \nabla u^i(\bar{x}^i),T_{b-} \rangle&= -\sigma ^i_{b-}\le 0, \end{aligned}$$

Since the optimization problem is convex, the KKT conditions are also sufficient for an optimal solution to the agent’s problem.

The gradient of the linear quadratic utility function fulfills in the equilibrium allocation \(\bar{x}^i\)

$$\begin{aligned} \langle \nabla u^i(\bar{x}^i),\tau \rangle \le 0 \quad \forall \tau \in \mathcal {C }. \end{aligned}$$

Summing up all agent’s equilibrium gradients, we define the vector

$$\begin{aligned} \bar{\gamma }:= \sum _{i=1}^I \nabla u^i(\bar{x}^i)= (\alpha _0, \alpha _1 1\!\!1-((\omega _{\varvec{1}}+Y\delta )-d_{\varvec{1}}))^T, \end{aligned}$$

where \(d_{\varvec{1}}=\sum _{i=1}^I d_{\varvec{1}}^i := \sum _{i=1}^I (1\!\!1 - r_{{ \varvec{1}}})z^i_{b-}\) is the aggregate shortfall from promises on the bond. Still we have \(\langle \bar{\gamma },\tau \rangle \le 0 \quad \forall \tau \in \mathcal {C}.\) Since any trade \(\tau \in \mathcal {C}\) can be decomposed as \(\tau =(-c(m),m)\), we get for \(\frac{1}{\alpha _0}{\bar{\gamma }}\) that \( \frac{\bar{\gamma }_{\varvec{1}}}{\alpha _0}= \frac{\alpha _1}{\alpha _0}1\!\!1- \frac{1}{\alpha _0}\tilde{ \omega }_{\varvec{1}}\) and \(c(m)\ge \left\langle \frac{ \bar{\gamma }_{\varvec{1}}}{\alpha _0}, m \right\rangle \), which proves the proposition. \(\square \)

1.3 Proof of Proposition 3

Suppose agent \(i\) goes long in the bond. Define the matrix by \(T_\text {long} = \begin{pmatrix} -q_b &{}\quad q_e \\ r_{\varvec{1}} &{}\quad Y \end{pmatrix} \). We know from proposition 2 that

$$\begin{aligned} \langle T_\text {long}^T, \nabla u^i(\bar{x}^i) \rangle&= \begin{pmatrix} 0\\ 0\end{pmatrix} \text { and} \end{aligned}$$
(9)
$$\begin{aligned} \langle T_\text {long}^T, \bar{\gamma } \rangle&= \begin{pmatrix} -\sigma _{b+} \\ 0\end{pmatrix} , \end{aligned}$$
(10)

where \(\nabla u^i(x^i)= (\alpha _0^i, \alpha _1^i-x^i_{\varvec{1}})^T\) and \(\bar{ \gamma }= (\alpha _0, \alpha _11\!\!1 -\tilde{\omega }_{\varvec{1}})^T\). Divide Eq. (9) by \(\alpha _0^i\) and Eq. (10) by \(\alpha _0\) , subtract the equations and multiply the result by \(\alpha _0^i\) again. With \(\bar{\tau }_{\varvec{1}}^i:=\bar{x}_{\varvec{1}}^i -\omega _{\varvec{1}}^i - Y \delta ^i\) this gives

$$\begin{aligned} \langle Y_{b+}^T, \bar{\tau }_{\varvec{1}}^i \rangle = \left\langle Y_{b+}^T, \left( \alpha _1^i-\frac{\alpha _0^i}{\alpha _0}\alpha _1\right) 1\!\!1 -\left( \left( \omega _{\varvec{1}} ^i+ Y \delta ^i\right) - \frac{\alpha _0^i}{\alpha _0} \tilde{\omega }_{\varvec{1}}\right) \right\rangle - \frac{\alpha _0^i}{\alpha _0} \begin{pmatrix} \sigma _{b+}\\ 0 \end{pmatrix} . \end{aligned}$$

We can write

$$\begin{aligned} \begin{pmatrix} \sigma _{b+} \\ 0 \end{pmatrix} = \left\langle Y_{b+}^T, \sigma _{b+}\frac{r_{\varvec{1}}- P_Y(r_{\varvec{1}})}{||r_{\varvec{1}} -P_Y(r_{\varvec{1}})||^2} \right\rangle = \langle Y_{b+}^T, \sigma _{b+} r_{{\varvec{1} }e} \rangle . \end{aligned}$$

Now, since \(\langle Y_{b+}^T, v_{\varvec{1}} \rangle =\langle Y_{b+}^T, P_{Y_{b+}}(v_{\varvec{1}}) \rangle \) for any vector \(v_{\varvec{1}}\in \mathbb {R}^{S}\), it follows that

$$\begin{aligned} \bar{\tau }_{\varvec{1}}^i = P_{Y_{b+}} \left( \left( \alpha _1^i-\frac{\alpha _0^i}{\alpha _0} \alpha _1\right) 1\!\!1 -\left( \left( \omega _{\varvec{1}}^i + Y \delta ^i\right) -\frac{\alpha _0^i}{ \alpha _0}\tilde{\omega }_{\varvec{1}}\right) - \sigma _{b+} \frac{\alpha _0^i}{\alpha _0} r_{ {\varvec{1}}e} \right) . \end{aligned}$$

Finally, since \(r_{\varvec{1}}-P_Y(r_{\varvec{1}})\in \text {span}(Y_{b+})\), the result follows.

The results for agents going short and for agents that do not trade in the bond are proved similarly. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eichberger, J., Rheinberger, K. & Summer, M. Credit risk in general equilibrium. Econ Theory 57, 407–435 (2014). https://doi.org/10.1007/s00199-014-0822-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-014-0822-2

Keywords

JEL Classification

Navigation