Skip to main content
Log in

Private ownership economies with externalities and existence of competitive equilibria: a differentiable approach

  • Published:
Journal of Economics Aims and scope Submit manuscript

Abstract

We consider a general equilibrium model of a private ownership economy with consumption and production externalities. Utility functions and production technologies may be affected by the consumption and production activities of all other agents in the economy. We use homotopy techniques to show that the set of competitive equilibria is non-empty and compact. Fixing the externalities, the assumptions on utility functions and production technologies are standard in a differentiable framework. Competitive equilibria are written in terms of first order conditions associated with agents’ behavior and market clearing conditions, following the seminal paper of Smale (J Math Econ 1:1–14, 1974). The work of adapting the homotopy approach to economies with externalities on the production side is non-trivial and it requires some ingenious adjustments, because the production technologies are not required to be convex with respect to the consumption and production activities of all agents.

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.

Similar content being viewed by others

Notes

  1. McKenzie has focused on a production economy with linear technologies where individual preferences are affected by the price system, the consumption of the others and the inputs of the firms.

  2. See Section 2 of Chapter 6 in Arrow and Hahn (1971).

  3. See page 134 in Section 2 of Chapter 6 of Arrow and Hahn (1971).

  4. For a detailed analysis of this assumption and the one made by Arrow and Hahn, see Assumption 3, Remark 4 and the interpretation of Lemma 5 in Sect. 2.

  5. The transformation function is a convenient way to represent a production set using an inequality on a function, see for instance Mas-Colell et al. (1995).

  6. For technical details see Assumptions 7 and 8 in Sect. 2.

  7. See for instance, Villanacci and Zenginobuz (2005), del Mercato (2006), Mandel (2008), Bonnisseau and del Mercato (2008), Kung (2008), and Ericson and Kung (2015).

  8. In Chapter 9 of Villanacci et al. (2002), the reader can find a homotopy proof for a standard private ownership economies without externalities. Our proof is simpler than the latter one, because we do not homotopize the shares.

  9. See for instance, Chapter 4 in Milnor (1965), and Villanacci et al. (2002).

  10. See for instance, Chapter 5 in Milnor (1965), and Mas-Colell (1985).

  11. See pages 53–55 in Ericson and Kung (2015).

  12. Some basic assumptions for the existence of an equilibrium are missing, i.e. the possibility of inaction and the compactness of the set of feasible allocations, or any related assumptions. See for instance, Assumptions 1(2) and 3 in our paper, or Assumptions P(2) and UB in Bonnisseau and Médecin (2001).

  13. Let v and \(v^{\prime }\) be two vectors in \({\mathbb {R}}^{n}, v \cdot v^{\prime }\) denotes the scalar product of v and \(v^{\prime }\). Let A be a real matrix with m rows and n columns, and B be a real matrix with n rows and l columns, AB denotes the matrix product of A and B. Without loss of generality, vectors are treated as row matrices and A denotes both the matrix and the following linear mapping \(A: v \in {\mathbb {R}}^{n}\rightarrow A(v):= A v^{T}\in {\mathbb {R}}^{[m]}\) where \(v^{T}\) denotes the transpose of v and \({\mathbb {R}}^{[m]}:=\{w^{T}: w\in {\mathbb {R}}^{m} \}\). When \(m=1, A(v)\) coincides with the scalar product \(A \cdot v\), treating A and v as vectors in \({\mathbb {R}}^{n}\).

  14. We have already pointed out that Arrow and Hahn need the stronger condition given in Remark 4 to extend their existence proof to the case of externalities. However in Section 2 in Chapter 6 of Arrow and Hahn (1971), the authors do not enter into a detailed proof. On the other hand, in Bonnisseau and Médecin (2001), Assumption UB is needed to find the cube to compactify the economy in order to use fixed point arguments.

  15. For example, consider two commodities and the utility function \(u_h(x_{h}^{1},x_{h}^{2},x_{k}^{1}):=x_{h}^{1}x_{h}^{2}x_{k}^{1}\) where \((x_{h}^{1},x_{h}^{2}) \in {\mathbb {R}}_{++}^{2}\) and \(x_{k}^{1} \in {\mathbb {R}}_{+}\). This function does not satisfy Assumption 7, but it satisfies Assumption 8.

  16. “KKT conditions” means Karush–Kuhn–Tucker conditions.

  17. As regards the theory of the degree modulo 2, we refer to Chapter 4 in Milnor (1965), Appendix B in Geanakoplos and Shafer (1990), and Chapter 7 in Villanacci et al. (2002).

  18. Notice that if, in addition, E is a regular economy, i.e. 0 is a regular value of F, then the computation of the degree modulo 2 implies that the number of equilibria of E is finite and odd. However, this paper does not address any regularity issue. In the presence of consumption and production externalities, the analysis of regular economies is quite sensitive. For the case without externalities on the production side, see Bonnisseau (2003), Kung (2008), and Bonnisseau and del Mercato (2010). In the presence of production externalities, the analysis of regular economies deserves a separate analysis, see del Mercato et al. (2017).

  19. Namely, the homotopy \(\Phi \) defined in (12) of Sect. 5.2.

  20. Using Debreu’s vocabulary, the Pareto optimal allocation \(({\widetilde{x}},{\widetilde{y}})\) is an equilibrium relative to some price system, see Section 6.4 in Debreu (1959).

  21. For example, the redistribution \({\widehat{e}}_{h} :={\widehat{s}}_{jh} \sum _{h\in {\mathcal {H}}} e_{h}\) with \({\widehat{s}}_{jh} := \frac{{\widetilde{p}}\cdot {\widetilde{x}}_{h}}{{\widetilde{p}}\cdot \sum _{h\in {\mathcal {H}}} {\widetilde{x}}_{h}}\).

  22. Even in the absence of externalities, homotopizing the shares complicates the homotopy proof, see for instance Section 2.2 in Chapter 9 of Villanacci et al. (2002).

  23. Indeed, at the economy E, the equilibrium individual wealth is strictly positive because of the possibility of inaction (Point 2 of Assumption 1) and standard arguments from profit maximization.

  24. Indeed, \(t_{j}(0,y_{-j},x)=0\) because of the possibility of inaction, and \(t_{j}({\widetilde{y}}_{j},{\overline{y}}_{-j},{\overline{x}} )=0\) because \(G ( {\widetilde{\xi }})= 0\). If \(t_{j}\) is quasi-convex with respect to all the variables, then \(t_{j}((1-\tau ){\widetilde{y}}_{j},\tau y_{-j}+(1-\tau ){\overline{y}}_{-j},\tau x+(1-\tau ){\overline{x}})\le 0\).

  25. Indeed, \(t_{j}(0,{\overline{y}}_{-j},{\overline{x}})=0\) and \(t_{j}({\widetilde{y}}_{j},{\overline{y}}_{-j},{\overline{x}} )=0\) imply that \(t_{j}((1-\tau ){\widetilde{y}}_{j},{\overline{y}}_{-j},{\overline{x}})\le 0\), because \(t_{j}\) is quasi-convex with respect to \(y_{j}\).

  26. The proof of these claims is standard because \({\overline{E}}\) is a standard production economy without externalities.

References

  • Arrow KJ, Hahn F (1971) General competitive analysis. Holden-Day, San Francisco

    Google Scholar 

  • Balasko Y (2015) Wealth concerns and equilibrium. J Math Econ 59:92–101

    Article  Google Scholar 

  • Bonnisseau J-M (2003) Regular economies with non-ordered preferences. J Math Econ 39:153–174

    Article  Google Scholar 

  • Bonnisseau J-M, del Mercato EL (2008) General consumption constraints and regular economies. J Math Econ 44:1286–1301

    Article  Google Scholar 

  • Bonnisseau J-M, del Mercato EL (2010) Externalities, consumption constraints and regular economies. Econ Theory 44:123–147

    Article  Google Scholar 

  • Bonnisseau J-M, Médecin J-P (2001) Existence of marginal pricing equilibria in economies with externalities and non-convexities. J Math Econ 36:271–294

    Article  Google Scholar 

  • Debreu G (1959) Theory of value. Yale University Press, New Haven

    Google Scholar 

  • Ericson RE, Kung F-C (2015) Fundamental non-convexities and externalities: a differentiable approach. B.E. J Theor Econ 15(1):49–62

    Article  Google Scholar 

  • del Mercato EL (2006) Existence of competitive equilibria with externalities: a differential viewpoint. J Math Econ 42:525–543

    Article  Google Scholar 

  • del Mercato EL, Platino V (2017) On the regularity of smooth production economies with externalities: competitive equilibrium à la Nash. Econ Theory. doi:10.1007/s00199-016-1029-5 (forthcoming)

  • Geanakoplos J, Shafer W (1990) Solving systems of simultaneous equations in economics. J Math Econ 19:69–93

    Article  Google Scholar 

  • Kung F-C (2008) Voluntary contributions to multiple public goods in a production economy with widespread externalities. J Math Econ 44:1364–1378

    Article  Google Scholar 

  • Laffont J-J (1976) Decentralization with externalities. Eur Econ Rev 7:359–375

    Article  Google Scholar 

  • Laffont J-J (1988) Fundamentals of public economics. The MIT Press, Cambridge

    Google Scholar 

  • Laffont J-J, Laroque G (1972) Effets externes et théorie de l’équilibre général. In: Cahiers du Séminaire d’Econométrie, vol 14. CNRS, Paris, pp 1–48

  • Mandel A (2008) An index formula for production economies with externalities. J Math Econ 44:1385–1397

    Article  Google Scholar 

  • Mas-Colell A (1985) The theory of general economic equilibrium. A differentiable approach. Cambridge University Press, Cambridge

    Google Scholar 

  • Mas-Colell A, Whinston MD, Green JR (1995) Microeconomic theory. Oxford University Press, Oxford

    Google Scholar 

  • McKenzie LW (1955) Competitive equilibrium with dependent consumer preferences. In: Proceedings of the the Second symposium of linear programming, Washington, DC, pp 277–294

  • Milnor JW (1965) Topology from the differentiable viewpoint. The University Press of Virginia, Charlottesville

    Google Scholar 

  • Smale S (1974) Global analysis and economics. II A: extension of a theorem of Debreu. J Math Econ 1:1–14

    Article  Google Scholar 

  • Villanacci A, Carosi L, Benevieri P, Battinelli A (2002) Differential topology and general equilibrium with complete and incomplete markets. Kluwer, Dordrecht

    Book  Google Scholar 

  • Villanacci A, Zenginobuz U (2005) Existence and regularity of equilibria in a general equilibrium model with private provision of a public good. J Math Econ 41:617–636

    Article  Google Scholar 

Download references

Acknowledgements

This paper dates back to 2011 and it has been available on the website: http://www.parisschoolofeconomics.eu/fr/del-mercato-elena/publications. We are delighted that this paper has been quoted by Balasko (2015). Furthermore, we have recently found a similar contribution in Ericson and Kung (2015). The 2011’s version has been presented at the Public Economic Theory (PET 10) and Public Goods, Public Projects, Externalities (PGPPE) Closing Conference, Bogazici University, 2010, and the Fifth Economic Behavior and Interaction Models (EBIM) Doctoral Workshop on Economic Theory, Bielefeld University, 2010. We thank the participants of these conferences for useful comments. We are grateful to two anonymous referees of this journal for their detailed comments and suggestions. Finally, we acknowledge the support of the Alliance Joint Project “Asymmetric Information, Externalities and Restricted Participation”, Columbia University and Université Paris 1 Panthéon–Sorbonne, 2016–2017.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincenzo Platino.

Appendix

Appendix

Proof of Lemma 5

Let \((x^{\prime },y^{\prime }) \in A(x,y;r)\). Since \(\sum \nolimits _{h\in {\mathcal {H}}} x_{h}^{\prime } \gg 0, y^{\prime }\) belongs to the bounded set C(r) given by Assumption 3. Furthermore, for every \(h\in {\mathcal {H}}, 0\ll x_{h}^{\prime }\ll \sum \nolimits _{h\in {\mathcal {H}}} x_{h}^{\prime } \le \sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\prime } + r\). Thus, there exists a bounded set \(K(r) \subseteq {\mathbb {R}}^{\textit{CH}}_{++}\times {\mathbb {R}}^{\textit{CJ}} \) such that for every \((x,y) \in {\mathbb {R}}^{\textit{CH}}_{++}\times {\mathbb {R}}^{\textit{CJ}}, A(x,y;r) \subseteq K(r)\). \(\square \)

Proof of Proposition 13

Let \({\overline{E}}\) be the production economy defined in Section 5.1. We remind that \(A({\overline{x}},{\overline{y}};r):=\{(x^{\prime },y^{\prime })\in {\mathbb {R}}_{++}^{CH}\times {\mathbb {R}}^{\textit{CJ}} : {\overline{t}}_{j}(y_{j}^{\prime })\le 0, \forall j \in {\mathcal {J}}\,\text { and}\,\sum \nolimits _{h\in {\mathcal {H}}} x_{h}^{\prime }-\sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\prime }\le r \}\). Consider the set \(U(r):= \{( u^{\prime }_{h})_{h\in {\mathcal {H}}} \in { \prod \nolimits _{h\in {\mathcal {H}}} } \mathrm{Im}\,{\overline{u}}_{h} \vert \exists (x^{\prime },y^{\prime })\in A({\overline{x}},{\overline{y}};r) :{\overline{u}}_{h}(x^{\prime }_{h}) \ge u^{\prime }_{h}, \forall h\in {\mathcal {H}}\} \). By Point 2 of Assumption 1, the set \(U_{r}\) is non-empty. Fix \(( u^{\prime }_{h})_{h\in {\mathcal {H}}} \in U(r)\) and consider the maximization problem

$$\begin{aligned} \begin{array}{ll} {\max \limits _{(x,y) \in {\mathbb {R}}_{++}^{\textit{CH}}\times {\mathbb {R}}^{\textit{CJ}}}} &{} {\overline{u}}_{1}(x_{1}) \\ \text {subject to } &{} \left\{ \begin{array}{l} {\overline{t}}_{j}(y_{j})\le 0, \quad \forall j \in {\mathcal {J}} \\ {\overline{u}}_{h}(x_{h}) \ge u^{\prime }_{h}, \quad \forall h \in {\mathcal {H}} \\ \sum _{h\in {\mathcal {H}}}x_{h}-\sum \limits _{j\in {\mathcal {J}}}y_{j} \le r \end{array} \right. \end{array} \end{aligned}$$

This problem has at least one solution \(({\widetilde{x}},{\widetilde{y}}) \in {\mathbb {R}}_{++}^{\textit{CH}}\times {\mathbb {R}}^{\textit{CJ}}\). Then, \(({\widetilde{x}},{\widetilde{y}})\) solves the maximization problem given below and it is a Pareto optimal allocation of the economy \({\overline{E}}\).Footnote 26

$$\begin{aligned} \begin{array}{ll} {\max \limits _{(x,y)\in {\mathbb {R}}_{++}^{\textit{CH}}\times {\mathbb {R}}^{\textit{CJ}}}} &{} {\overline{u}}_{1}(x_{1}) \\ \text {subject to} &{} \left\{ \begin{array}{l} - {\overline{t}}_{j}(y_{j})\ge 0, \quad \forall j \in {\mathcal {J}} \\ {\overline{u}}_{h}(x_{h}) - {\overline{u}}_{h}\left( {\widetilde{x}}_{h}\right) \ge 0, \quad \forall h\ne 1 \\ r - \sum \limits _{h\in {\mathcal {H}}}x_{h} + \sum _{j\in {\mathcal {J}}}y_{j} \ge 0 \end{array} \right. \end{array} \end{aligned}$$
(14)

We now claim that there exists \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }}):=(({\widetilde{\beta }}_{j})_{j\in {\mathcal {J}}},( {\widetilde{\theta }} _{h}) _{h \ne 1},{\widetilde{\gamma }}) \in {\mathbb {R}}_{++}^{J}\times {\mathbb {R}}_{++}^{H-1}\times {\mathbb {R}}_{++}^{C}\) such that \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\) is the unique solution to system (6). We first prove the existence of \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\), afterwards we show the uniqueness of \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\).

Existence of \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }}) \gg 0\). We know that \(({\widetilde{x}},{\widetilde{y}})\) solves problem (14). The KKT conditions associated with problem (14) are given by

$$\begin{aligned} \left\{ \begin{array}{l} D_{x_1} {\overline{u}}_{1}( x_{1}) = \gamma , \quad \forall h \ne 1: \theta _{h}D_{x_h}{\overline{u}}_{h}( x_{h}) = \gamma \quad \text {and}\quad \theta _{h} \left( {\overline{u}}_{h}( x_{h}) - {\overline{u}}_{h}\left( {\widetilde{x}}_{h}\right) \right) =0, \\ \forall j \in {\mathcal {J}} : \gamma = \beta _{j}D_{y_{j}}{\overline{t}}_{j}(y_{j})\quad \text {and}\quad \beta _{j}\left( - {\overline{t}}_{j}(y_{j})\right) =0, \quad \forall c \in {\mathcal {C}} : \gamma ^{c}\left( r^{c} - {\sum \limits _{h\in {\mathcal {H}}}}x_{h}^{c}+\sum \limits _{j\in {\mathcal {J}}}y_{j}^{c}\right) =0\\ \end{array} \right. \nonumber \\ \end{aligned}$$
(15)

where \((\beta ,\theta , \gamma ):=((\beta _{j})_{j\in {\mathcal {J}}},( \theta _{h}) _{h\ne 1},(\gamma ^{c})_{c \in {\mathcal {C}}}) \in {\mathbb {R}}_{+}^{J}\times {\mathbb {R}}_{+}^{H-1}\times {\mathbb {R}}^{C}_{+}\) are the Lagrange multipliers associated with the constraint functions of problem (14). KKT conditions are obviously necessary conditions to solve problem (14). Indeed, it is easy to verify that the Jacobian matrix associated with the constraint functions of problem (14) has full row rank equal to \(J+(H-1)+C\). This property is standard because the externalities are fixed. Therefore, there exists \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }}) \ge 0\) such that \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\) solves system (15). Furthermore, Point 4 of Assumption 1 and Point 2 of Assumption 6 imply that all the Lagrange multipliers \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\) must be strictly positive. Consequently, all the constraints of problem (14) are binding, and then \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\) is a solution to system (6).

Uniqueness of \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }}, {\widetilde{\theta }},{\widetilde{\gamma }})\). Define \({\widetilde{\theta }}_{1}:=1\), by equations (1) and (2) of system (6), for all h one gets \(D_{x_h} {\overline{u}}_{h}({\widetilde{x}}_{h}) = \frac{{\widetilde{\gamma }}}{{\widetilde{\theta }}_{h}}\). Then, for every \(h, {\widetilde{x}}_{h}\) solves the maximization problem: \( {\max \nolimits _{x_{h} \in {\mathbb {R}}^{C}_{++}}}{\overline{u}}_{h}(x_{h})\) subject to \(\frac{{\widetilde{\gamma }}}{{\widetilde{\theta }}_{h}} \cdot x_{h} \le \frac{{\widetilde{\gamma }}}{{\widetilde{\theta }}_{h}} \cdot {\widetilde{x}}_{h}\) because KKT are sufficient conditions to solve this problem. Thus, the uniqueness of \({\widetilde{x}}_{h}\) follows from the strict quasi-concavity of \({\overline{u}}_{h}\). Analogously, by Eqs. (4) and (5) of system (6), \({\widetilde{y}}_{j}\) solves the maximization problem: \( {\max \nolimits _{y_{j} \in {\mathbb {R}}^{C}}} \frac{{\widetilde{\gamma }}}{{\widetilde{\beta }}_{j}} \cdot y_{j}\) subject to \({\overline{t}}_{j}(y_{j})\le 0\) for every j. Thus, the uniqueness of \({\widetilde{y}}_{j}\) follows from the continuity and the strict quasi-convexity of \({\overline{t}}_{j}\). Therefore, \(({\widetilde{x}},{\widetilde{y}})\) is unique and consequently, the uniqueness of \(({\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})\) follows from equations (1), (2) and (4) of system (6). \(\square \)

Proof of Proposition 14

We use the functions \({\overline{u}}_{h}\) and \({\overline{t}}_{j}\) defined in Subsection 5.1. We have already pointed out that \(G({\widetilde{\xi }})=0\). Let \(\xi ^{\prime }=(x^{\prime },\lambda ^{\prime },y^{\prime },\alpha ^{\prime }, p^{\prime \backslash }) \in \Xi \) be such that \(G(\xi ^{\prime })=0\), we show that \({\widetilde{\xi }}=\xi ^{\prime }\).

First, notice that

$$\begin{aligned} \sum _{h\in {\mathcal {H}}}x_{h}^{\prime } - \sum _{j\in {\mathcal {J}}} y_{j}^{\prime } ={ \sum _{h\in {\mathcal {H}}}} e_{h} \end{aligned}$$
(16)

Indeed, summing \(G^{h.2}(\xi ^{\prime })=0\) over h, one gets \(\sum \nolimits _{h\in {\mathcal {H}}}x_{h}^{\prime } - \sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\prime } ={ \sum \nolimits _{h\in {\mathcal {H}}}} {\widetilde{e}}_h\) by \(G^{M}(\xi ^{\prime })=0\). Using the definition of \({\widetilde{e}}_h\) given in (7) and Proposition 13, one deduces (16).

Second, we show that

$$\begin{aligned} {\overline{u}}_{h}\left( x_{h}^{\prime }\right) = {\overline{u}}_{h}\left( {\widetilde{x}}_{h}\right) ,\quad \forall h \in {\mathcal {H}} \end{aligned}$$
(17)

Using the definition of \({\widetilde{e}}_h\) given in (7) and \(G^{h.1}(\xi ^{\prime })=G^{h.2}(\xi ^{\prime })=0, x^{\prime }_{h}\) solves the following maximization problem

$$\begin{aligned} \begin{array}{l} {\max \limits _{x_{h}\in {\mathbb {R}}_{++}^{C}}}\; {\overline{u}}_{h}(x_{h}) \\ \text {subject to}\;\;p^{\prime }\cdot x_{h}\le p^{\prime }\cdot {\widetilde{x}}_{h}+ \sum \limits _{j\in {\mathcal {J}}}s_{jh} p^{\prime }\cdot \left( y^{\prime }_{j} -{\widetilde{y}}_{j}\right) \\ \end{array} \end{aligned}$$
(18)

because KKT are sufficient conditions to solve this problem. Analogously, from \(G^{j.1}(\xi ^{\prime })=G^{j.2}(\xi ^{\prime })=0, y^{\prime }_{j}\) solves the maximization problem: \({ \max \nolimits _{y_{j} \in {\mathbb {R}}^{C}}} p^{\prime }\cdot y_{j}\) subject to \({\overline{t}}_{j}(y_{j})\le 0\). Notice that \({\widetilde{y}}_{j}\) satisfies the constraint of this problem because \(G^{j.2}({\widetilde{\xi }})=0\). Thus, \(p^{\prime } \cdot (y^{\prime }_{j}-{\widetilde{y}}_{j}) \ge 0\) for all j, and consequently \({\widetilde{x}}_{h}\) belongs to the budget constraint of problem (18). Then, \({\overline{u}}_{h}(x_{h}^{\prime }) \ge {\overline{u}}_{h}({\widetilde{x}}_{h})\) for all h. Now, suppose that \({\overline{u}}_{k}(x_{k}^{\prime }) > {\overline{u}}_{k}({\widetilde{x}}_{k})\) for some \(k \in {\mathcal {H}}\). From (16) and \(G^{j.2}(\xi ^{\prime })=0\) for all j, one deduces that \((x^{\prime },y^{\prime })\) is a feasible allocation of the production economy \({\overline{E}}\), and then one gets a contradiction since \(({\widetilde{x}},{\widetilde{y}})\) is a Pareto optimal allocation of \({\overline{E}}\) by Proposition 13. Thus, (17) is completely proved.

Now, define \(\beta ^{\prime }:=(\beta _{j}^{\prime })_{j\in {\mathcal {J}}}\) where \(\beta _{j}^{\prime }:= \lambda _{1}^{\prime } \alpha _{j}^{\prime }\) for all \(j, \theta ^{\prime }:=(\theta _{h}^{\prime }) _{h \ne 1}\) where \(\theta _{h}^{\prime }:= \frac{\lambda _{1}^{\prime } }{\lambda _{h}^{\prime } }\) for all \(h \ne 1\) and \(\gamma ^{\prime }:= \lambda _{1}^{\prime } p^{\prime }\). From \(G^{h.1}(\xi ^{\prime })=0\) for all \(h, G^{j.1}(\xi ^{\prime })=G^{j.2}(\xi ^{\prime })=0\) for all j, (16) and (17), it is an easy matter to check that \((x^{\prime },y^{\prime }, \beta ^{\prime },\theta ^{\prime },\gamma ^{\prime })\) solves system (6). Then, Proposition 13 implies that \(({\widetilde{x}},{\widetilde{y}},{\widetilde{\beta }},{\widetilde{\theta }},{\widetilde{\gamma }})=(x^{\prime },y^{\prime }, \beta ^{\prime },\theta ^{\prime },\gamma ^{\prime })\), and consequently, one deduces that \({\widetilde{\xi }}=\xi ^{\prime }\).

We remark that G is \(C^{1}\) by Point 1 of Assumptions 1 and 6. Finally, in order to show that 0 is a regular value for G, one proves that \(D_{\xi }G( {\widetilde{\xi }})\) has full row rank. In this regard, we show that if \(\Delta D_{\xi }G( {\widetilde{\xi }})=0\), then \(\Delta =0\) where \(\Delta :=((\Delta x_{h},\Delta \lambda _{h})_{h\in {\mathcal {H}}},(\Delta y_{j},\Delta \alpha _{j})_{j\in {\mathcal {J}}},\Delta p^{\backslash })\in {\mathbb {R}}^{\dim \Xi }\). The system \(\Delta D_{\xi }G( {\widetilde{\xi }})=0\) is given below.

$$\begin{aligned} \left\{ \begin{array}{l} (h.1)\;\;\Delta x_{h} D_{x_{h}}^{2}{\overline{u}}_{h}\left( {\widetilde{x}}_{h}\right) - \Delta \lambda _{h} {\widetilde{p}} + \Delta p^{\backslash } \left[ I_{C-1}|0\right] =0, \quad \forall \;h\in {\mathcal {H}} \\ (h.2) \;\; - \Delta x_{h} \cdot {\widetilde{p}}=0, \forall \;h\in {\mathcal {H}} \\ (j.1) \;\; \sum \limits _{h\in {\mathcal {H}}} \Delta \lambda _{h} s_{jh} {\widetilde{p}} - {\widetilde{\alpha }}_{j} \Delta y_{j} D_{y_{j}}^{2}{\overline{t}}_{j}\left( {\widetilde{y}}_{j}\right) -\Delta \alpha _{j} D_{y_{j}} {\overline{t}}_{j}\left( {\widetilde{y}}_{j}\right) - \Delta p^{\backslash } \left[ I_{C-1}|0\right] =0, \quad \forall \;j\in {\mathcal {J}} \\ (j.2) \;\; - \Delta y_{j} \cdot D_{y_{j}} {\overline{t}}_{j}\left( {\widetilde{y}}_{j}\right) =0, \quad \forall \;j\in {\mathcal {J}} \\ (M) \;\; -{\displaystyle \sum _{h \in {\mathcal {H}}}} {\widetilde{\lambda }}_{h} \Delta x_{h}^{\backslash } + \sum _{j\in {\mathcal {J}}}\Delta y_{j}^{ \backslash }=0 \end{array} \right. \end{aligned}$$

We first prove that \(\Delta x_{h}=0\) for all \(h\in {\mathcal {H}}\). Otherwise, suppose that there is \({\overline{h}}\in {\mathcal {H}}\) such that \(\Delta x_{{\overline{h}}} \ne 0\). The proof goes through the two following claims that contradict each others.

We first claim that \(\Delta p^{\backslash }\cdot (\sum \nolimits _{h\in {\mathcal {H}}} {\widetilde{\lambda }}_{h} \Delta x_{h}^{\backslash })>0\). Multiplying (h.1) by \( {\widetilde{\lambda }}_{h} \Delta x_{h} \) and summing over h, from (h.2) we get \(\sum \nolimits _{h\in {\mathcal {H}}} {\widetilde{\lambda }}_{h} \Delta x_{h}D_{x_{h}}^{2}{\overline{u}}_{h}({\widetilde{x}}_{h}) (\Delta x_{h})=-\Delta p^{\backslash }\cdot ( \sum \nolimits _{h\in {\mathcal {H}}} {\widetilde{\lambda }}_{h} \Delta x_{h}^{\backslash })\). Multiplying \(G^{h.1}({\widetilde{\xi }})=0\) by \(\Delta x_{h}\) and using (h.2), we get \( \Delta x_{h} \cdot D_{x_{h}} {\overline{u}}_{h}({\widetilde{x}}_{h})=0\) for all h. Therefore, Point 3 of Assumption 6 completes the proof of the claim since \({\widetilde{\lambda }}_{h}>0\) for all h and \(\Delta x_{{\overline{h}}}\ne 0\).

Second, we claim that \(\Delta p^{\backslash }\cdot (\sum \nolimits _{h\in {\mathcal {H}}} {\widetilde{\lambda }}_{h} \Delta x_{h}^{\backslash })\le 0\). Multiplying both sides of \(G^{j.1}({\widetilde{\xi }})=0\) by \(\Delta y_{j}\) and using (j.2), we get \( \Delta y_{j} \cdot {\widetilde{p}} =0\). Then, multiplying (j.1) by \(\Delta y_{j}\) and summing over j, from (j.2) we get \(- \sum \nolimits _{j\in {\mathcal {J}}} {\widetilde{\alpha }}_{j} \Delta y_{j} D_{y_{j}}^{2} {\overline{t}}_{j}({\widetilde{y}}_{j})(\Delta y_{j})=\Delta p^{\backslash }\cdot {\sum \nolimits _{j\in {\mathcal {J}}}}\Delta y_{j}^{\backslash } \). Since, \({\widetilde{\alpha }}_{j}>0\) for all j, Point 3 of Assumption 1 and (j.2) imply that \(\Delta p^{\backslash }\cdot {\sum \nolimits _{j\in {\mathcal {J}}}}\Delta y_{j}^{\backslash } \le 0.\) Using (M), the claim is completely proved.

Since \({\widetilde{p}}^{C}=1\) and \(\Delta x_{h} = 0\) for all \(h\in {\mathcal {H}}\), from (h.1) we get \(\Delta \lambda _{h}=0\) for all h, and then \(\Delta p^{\backslash }=0\). Thus, multiplying (j.1) by \(\Delta y_{j}\), Point 3 of Assumption 1 and (j.2) imply that \(\Delta y_{j}=0\). Therefore, using once again (j.1), we get \(\Delta \alpha _{j}=0\) by Point 4 of Assumption 1. Thus, we get \(\Delta = 0\). \(\square \)

Proof of Proposition 16

Observe that \(H^{-1}(0)=\Phi ^{-1}(0)\cup \Gamma ^{-1}(0)\). Since the union of a finite number of compact sets is compact, it is enough to show that \(\Phi ^{-1}(0)\) and \(\Gamma ^{-1}(0)\) are compact. \(\square \)

Claim 1

\(\Phi ^{-1}(0)\) is compact.

We prove that, up to a subsequence, every sequence \(( \xi ^{\nu }, \tau ^{\nu }) _{\nu \in {\mathbb {N}}} \subseteq \Phi ^{-1}(0)\) converges to an element of \(\Phi ^{-1}( 0) \), where \(\xi ^{\nu }:=(x^{\nu },\lambda ^{\nu }, y^{\nu },\alpha ^{\nu },p^{\nu \;\backslash })_{\nu \in {\mathbb {N}}}\). Since \(\{\tau ^{\nu }: \nu \in {\mathbb {N}}\}\subseteq [0,1]\), up to a subsequence, \(( \tau ^{\nu }) _{\nu \in {\mathbb {N}}}\) converges to some \(\tau ^{*} \in [0,1]\). From Steps 1.1–1.4 below, up to a subsequence, \((\xi ^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \(\xi ^{*}:=(x^{*},\lambda ^{*}, y^{*},\alpha ^{*},p^{*\;\backslash })\in \Xi \). Since \(\Phi \) is continuous, taking the limit, one gets \((\xi ^{*},\tau ^{*})\in \Phi ^{-1}( 0)\).

We remind that for every \(\tau \in [0,1], e_{h}({\tau })\) defined in (11) is given by

$$\begin{aligned} e_{h}({\tau })= \tau e_{h}+(1-\tau ) {\widetilde{e}}_{h} \end{aligned}$$

Step 1.1. Up to a subsequence, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \((x^{*},y^{*})\in {\mathbb {R}}_{+}^{\textit{CH}}\times {\mathbb {R}}^{\textit{CJ}}\). We show that for \(r=\sum \nolimits _{h\in {\mathcal {H}}}e_{h}\), the sequence \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) is included in the bounded set K(r) given by Lemma 5. By \(\Phi ^{j.2}(\xi ^{\nu },\tau ^{\nu })=0\), for every j we get

$$\begin{aligned} t_{j}\left( y_{j}^{\nu },{\overline{y}}_{-j},{\overline{x}}\right) =0, \quad \forall \; \nu \in {\mathbb {N}} \end{aligned}$$

Thus, the sequence \((y^{\nu })_{\nu \in {\mathbb {N}}}\) is included in the set \(Y({\overline{x}},{\overline{y}})\) given by (1). Summing \(\Phi ^{h.2}(\xi ^{\nu },\tau ^{\nu })=0\) over h, by \(\Phi ^{M}(\xi ^{\nu },\tau ^{\nu })=0\) we have \(\sum \nolimits _{h\in {\mathcal {H}}} x_{h}^{\nu } - \sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\nu }= \sum \nolimits _{h\in {\mathcal {H}}} e_{h}( \tau ^{\nu })\) for every \(\nu \in {\mathbb {N}}\). Using the definition of \({\widetilde{e}}_h\) given by (7) and Proposition 13, one gets \(\sum \nolimits _{h\in {\mathcal {H}}} e_{h}( \tau ) = r\) for every \(\tau \in [0,1]\), and then \(\sum \nolimits _{h\in {\mathcal {H}}} x_{h}^{\nu } - \sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\nu } = r\) for every \(\nu \in {\mathbb {N}}\). Thus, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}} \subseteq A({\overline{x}},{\overline{y}};r) \subseteq K(r)\). Consequently, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) is included in \(\mathrm{cl}\,K(r)\) which is a compact set. Therefore, up to a subsequence, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \((x^{*},y^{*})\in \mathrm{cl}\,K(r) \subseteq {\mathbb {R}}_{+}^{\textit{CH}}\times {\mathbb {R}}^{\textit{CJ}}\), and then \((x^{*},y^{*})\in {\mathbb {R}}_{+}^{CH}\times {\mathbb {R}}^{\textit{CJ}}\).

Step 1.2. The consumption allocation \(x^{*}\) is strictly positive, i.e. \(x_{h}^{*} \gg 0\) for every \(h\in {\mathcal {H}}\). By \(\Phi ^{h.1}(\xi ^{\nu },\tau ^{\nu })=\Phi ^{h.2}(\xi ^{\nu },\tau ^{\nu })=0\) and KKT sufficient conditions, \(x_{h}^{\nu }\) solves the following problem for every \(\nu \in {\mathbb {N}}\).

$$\begin{aligned} \begin{array}{ll} \max \limits _{x_{h}\in {\mathbb {R}}_{++}^{C}} \; u_{h}\left( x_{h}, {\overline{x}}_{- h}, {\overline{y}}\right) \\ \begin{aligned} \text {subject to } &{} p^{\nu }\cdot x_{h} \le p^{\nu }\cdot \left[ \tau ^{\nu } e_{ h}+\left( 1-\tau ^{\nu }\right) {\widetilde{x}}_{h}\right] + p^{\nu }\cdot \sum _{j\in {\mathcal {J}}}s_{jh}\left( y_{j}^{\nu }- \left( 1-\tau ^{\nu }\right) {\widetilde{y}}_{j}\right) \end{aligned} \end{array}\nonumber \\ \end{aligned}$$
(19)

We first claim that for every \(\nu \in {\mathbb {N}}\), the following vector

$$\begin{aligned} {\widehat{e}}_{h}\left( \tau ^{\nu }\right) :=\tau ^{\nu }e_{ h}+\left( 1-\tau ^{\nu }\right) {\widetilde{x}}_{h} \end{aligned}$$
(20)

belongs to the budget constraint of the problem above. By \(\Phi ^{j.1}(\xi ^{\nu },\tau ^{\nu })=\Phi ^{j.2}(\xi ^{\nu },\tau ^{\nu })=0\) and KKT sufficient conditions, \(y_{j}^{\nu }\) solves the following problem for every \(\nu \in {\mathbb {N}}\).

$$\begin{aligned} \begin{array}{l} \max \limits _{y_{j}\in {\mathbb {R}}^{C}}p^{\nu }\cdot y_{j} \\ \text {subject to } t_{j}\left( y_{j},{\overline{y}}_{- j},{\overline{x}}\right) \le 0 \end{array} \end{aligned}$$
(21)

\(t_{j}({\widetilde{y}}_{j},{\overline{y}}_{- j},{\overline{x}})= 0\) since \(G^{j.2}({\widetilde{\xi }})=0\), see (8). By Point 2 of Assumption 1, \( t_{j}(0,{\overline{y}}_{- j},{\overline{x}})= 0\). Then, we get \(t_{j}((1-\tau ^{\nu }){\widetilde{y}}_{j},{\overline{y}}_{-j},{\overline{x}})< 0\) since \(t_{j}(\cdot ,{\overline{y}}_{-j},{\overline{x}}) \) is strictly quasi-convex, that is, the production plan \((1-\tau ^{\nu }){\widetilde{y}}_{j}\) belongs to the constraint set of problem (21). Thus, \(p^{\nu }\cdot (y_{j}^{\nu } -(1-\tau ^{\nu }){\widetilde{y}}_{j})\ge 0\) for every j, and then \(p^{\nu }\cdot \sum \nolimits _{j\in {\mathcal {J}}}s_{jh}( y_{j}^{\nu }- (1-\tau ^{\nu }){\widetilde{y}}_{j}) \ge 0\) which completes the proof of the claim.

Therefore, for every \(\nu \in {\mathbb {N}}, u_{h}(x_{h}^{\nu }, {\overline{x}}_{- h}, {\overline{y}}) \ge u_{h}( {\widehat{e}}_{h}(\tau ^{\nu }), {\overline{x}}_{- h}, {\overline{y}})\). By Point 2 of Assumption 6, for every \(\varepsilon >0\) we get \(u_{h}(x_{h}^{\nu }+\varepsilon \mathbf{1}, {\overline{x}}_{- h}, {\overline{y}}) > u_{h}( {\widehat{e}}_{h}(\tau ^{\nu }), {\overline{x}}_{- h}, {\overline{y}})\) where \(\mathbf{1}:=(1,\dots ,1)\in {\mathbb {R}}_{++}^{C}\). Taking the limit over \(\nu \) and using Point 1 of Assumption 6, we get \(u_{h}(x_{h}^{*} +\varepsilon \mathbf{1}, {\overline{x}}_{- h}, {\overline{y}}) \ge u_{h}( {\widehat{e}}_{h}(\tau ^{*}), {\overline{x}}_{- h}, {\overline{y}})\) for every \(\varepsilon >0\). Then, \(x_{h}^{*}\) belongs to the closure of the upper counter set of \(({\widehat{e}}_{h}(\tau ^{*}), {\overline{x}}_{- h}, {\overline{y}})\), which is included in \({\mathbb {R}}_{++}^{C}\) by Point 4 of Assumption 6. Thus, \(x_{h}^{*} \gg 0\).

Step 1.3. Up to a subsequence, \((\lambda ^{\nu },p^{\nu \;\backslash }) _{\nu \in {\mathbb {N}}}\) converges to some \(\lambda ^{*} \in {\mathbb {R}}_{++}^{H}\times {\mathbb {R}}_{++}^{C-1}\). The proof is similar to the proof of Step 2.3.

Step 1.4. Up to a subsequence, \((\alpha ^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \(\alpha ^{*} \in {\mathbb {R}}_{++}^{J}\). The proof is similar to the proof of Step 2.4.

Claim 2

\(\Gamma ^{-1}(0) \) is compact.

Let \(( \xi ^{\nu }, \tau ^{\nu }) _{\nu \in {\mathbb {N}}}\) be a sequences in \(\Gamma ^{-1}( 0)\). As in Claim \(1, ( \tau ^{\nu }) _{\nu \in {\mathbb {N}}}\) converges to \(\tau ^{*}\in [0,1]\). From Steps 2.1–2.4 below, up to a subsequence, \((\xi ^{\nu })_{\nu \in {\mathbb {N}}}\) converges to an element \(\xi ^{*}:=( x^{*},\lambda ^{*}, y^{*},\alpha ^{*},p^{*\backslash }) \in \Xi \). Since \(\Gamma \) is a continuous function, taking limit, we get \((\xi ^{*},\tau ^{*})\in \Gamma ^{-1}( 0)\).

We remind that for every \(\tau \in [0,1], x({\tau })\) and \(y({\tau })\) defined in (11) are given by

$$\begin{aligned} x({\tau })= \tau x+(1-\tau ) {\overline{x}} \quad \text { and }\quad y({\tau })=\tau y+(1-\tau ) {\overline{y}} \end{aligned}$$

Step 2.1. Up to a subsequence, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \((x^{*},y^{*})\in {\mathbb {R}}_{+}^{CH}\times {\mathbb {R}}^{CJ}\). We show that for \(r=\sum \nolimits _{h\in {\mathcal {H}}}e_{h}\), the sequence \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}}\) is included in the bounded set K(r) given by Lemma 5. Then, one completes the proof as in Step 1.1. By \(\Gamma ^{j.2}(\xi ^{\nu },\tau ^{\nu })=0\), for every j we have that

$$\begin{aligned} t_{j}\left( y_{j}^{\nu },y_{-j}^{\nu }\left( \tau ^{\nu }\right) ,x^{\nu }\left( \tau ^{\nu }\right) \right) =0, \quad \forall \nu \in {\mathbb {N}} \end{aligned}$$

Thus, for every \(\nu \in {\mathbb {N}}\), the production allocation \(y^{\nu }\) belongs to the set \(Y(x^{\nu }(\tau ^{\nu }),y^{\nu }(\tau ^{\nu }))\) given by (1). Now, summing \(\Gamma ^{h.2}(\xi ^{\nu },\tau ^{\nu })=0\) over h, by \(\Gamma ^{M}(\xi ^{\nu },\tau ^{\nu })=0\), we get \(\sum \nolimits _{h\in {\mathcal {H}}}x_{h}^{\nu }- \sum \nolimits _{j\in {\mathcal {J}}} y_{j}^{\nu }=r\). Then, for every \(\nu \in {\mathbb {N}}\), the allocation \((x^{\nu },y^{\nu })\) belongs to the set \(A(x^{\nu }(\tau ^{\nu }),y^{\nu }(\tau ^{\nu });r) \subseteq K(r)\), and consequently, \((x^{\nu },y^{\nu })_{\nu \in {\mathbb {N}}} \subseteq K(r)\).

Step 2.2. The consumption allocation \(x^{*}\) is strictly positive, i.e. \(x_{h}^{*} \gg 0\) for every \(h\in {\mathcal {H}}\). The argument is similar to the one used in Step 1.2 except for the last part which is quite different due to the presence of consumption externalities in the utility functions. For this reason, at the end of this step we need Assumption 7 or, alternatively, Assumption 8.

First, according to \(\Gamma ^{h.1}(\xi ^{\nu },\tau ^{\nu })=\Gamma ^{h.2}(\xi ^{\nu },\tau ^{\nu })=0\), replace problem (19) with the following problem

$$\begin{aligned} \begin{array}{ll} \max \limits _{x_{h}\in {\mathbb {R}}_{++}^{C}} u_{h}\left( x_{h}, x^{\nu }_{- h}\left( \tau ^{\nu }\right) , y^{\nu }\left( \tau ^{\nu }\right) \right) \\ \begin{aligned} \text {subject to } &{}p^{\nu }\cdot x_{h} \le p^{\nu }\cdot e_{h}+p^{\nu }\cdot \sum _{j\in {\mathcal {J}}}s_{jh} y_{j}^{\nu } \end{aligned} \end{array} \end{aligned}$$
(22)

and replace \({\widehat{e}}_h(\tau ^{\nu })\) given by (20) with \({\widehat{e}}_h(\tau ^{\nu }):=e_h\). Second, according to \(\Gamma ^{j.1}(\xi ^{\nu },\tau ^{\nu })=\Gamma ^{j.2}(\xi ^{\nu },\tau ^{\nu })=0\), replace problem (21) with the following problem

$$\begin{aligned} \begin{array}{l} \max \limits _{y_{j}\in {\mathbb {R}}^{C}}p^{\nu }\cdot y_{j} \\ \text {subject to } t_{j}\left( y_{j},y_{-j}^{\nu }\left( {\tau ^{\nu }}\right) ,x^{\nu }\left( {\tau ^{\nu }}\right) \right) \le 0 \end{array} \end{aligned}$$

Third, one follows the same strategy as for Step 1.2. For every \(\nu \in {\mathbb {N}}, u_{h}(x_{h}^{\nu }, x^{\nu }_{- h}(\tau ^{\nu }), y^{\nu }(\tau ^{\nu })) \ge u_{h}( e_{h}, x^{\nu }_{- h}(\tau ^{\nu }), y^{\nu }(\tau ^{\nu }))\). Notice that \(x^{\nu }_{- h}(\tau ^{\nu })\) belongs to \({\mathbb {R}}_{++}^{C(H-1)}\), because \(x^{\nu }_{- h}(\tau ^{\nu })= \tau ^{\nu } x^{\nu }_{- h}+(1-\tau ^{\nu }) {\overline{x}}_{- h}\), and \(x^{\nu }_{- h}\) and \({\overline{x}}_{- h}\) are both strictly positive. Then, by Point 2 of Assumption 6, for every \(\nu \in {\mathbb {N}}\) and for every \(\varepsilon >0\) we get

$$\begin{aligned} u_{h}\left( x_{h}^{\nu }+\varepsilon \mathbf{1}, x^{\nu }_{- h}\left( \tau ^{\nu }\right) , y^{\nu }\left( \tau ^{\nu }\right) \right) > u_{h}\left( e_{h}, x^{\nu }_{- h}\left( \tau ^{\nu }\right) , y^{\nu }\left( \tau ^{\nu }\right) \right) \end{aligned}$$
(23)

Now, take the limit over \(\nu \). Differently from Step 1.2, we have the two following cases: \(x^{*}_{-h}(\tau ^{*}) \gg 0\) or \(x^{*}_{-h}(\tau ^{*}) \in \mathrm{Bd}\,( {\mathbb {R}}_{++}^{C(H-1)})\). The second case may happen here, but not in Step 1.2, because now there are consumption externalities in the utility functions. Indeed, if \(( \tau ^{\nu }) _{\nu \in {\mathbb {N}}}\) converges to \(\tau ^{*}=1\), then \(x^{*}_{- h}(\tau ^{*})=x^{*}_{-h}\) which a priori is not necessarily strictly positive.

Suppose that \(x^{*}_{-h}(\tau ^{*}) \gg 0\). Using (23) and Point 1 of Assumption 6, we get \(u_{h}(x_{h}^{*} +\varepsilon \mathbf{1}, x^{*}_{-h}(\tau ^{*}), y^{*}(\tau ^{*})) \ge u_{h}( e_{h}, x^{*}_{-h}(\tau ^{*}), y^{*}(\tau ^{*}))\) for every \(\varepsilon >0\). Then, \(x_{h}^{*}\) is strictly positive, because it belongs to the closure of the upper counter set of \((e_{h}, x^{*}_{-h}(\tau ^{*}), y^{*}(\tau ^{*}))\), which is included in \({\mathbb {R}}_{++}^{C}\) by Point 4 of Assumption 6.

Suppose that \(x^{*}_{-h}(\tau ^{*}) \in \mathrm{Bd}\,( {\mathbb {R}}_{++}^{C(H-1)})\). Under Assumption 7, the previous argument still holds true.

Alternatively, if Assumption 7 is not satisfied, one uses Assumption 8 as follows. By Assumption 8 and (23), there exists \(\delta >0\) such that for every \(\nu \in {\mathbb {N}}\) and for every \(\varepsilon >0\) one gets

$$\begin{aligned} u_{h}\left( x_{h}^{\nu }+\varepsilon \mathbf{1}, x^{\nu }_{- h}\left( \tau ^{\nu }\right) +\delta \mathbf{1}, y^{\nu }\left( \tau ^{\nu }\right) \right) \ge u_{h}\left( e_{h}, x^{\nu }_{- h}\left( \tau ^{\nu }\right) +\delta \mathbf{1}, y^{\nu }\left( \tau ^{\nu }\right) \right) \end{aligned}$$

Take the limit over \(\nu \) and remark that \(x^{*}_{-h}(\tau ^{*}) +\delta \mathbf{1} \gg 0\). Consequently, from Point 1 of Assumption 6 one gets \(u_{h}(x_{h}^{*} +\varepsilon \mathbf{1}, x^{*}_{-h}(\tau ^{*}) +\delta \mathbf{1}, y^{*}(\tau ^{*})) \ge u_{h}( e_{h}, x^{*}_{-h}(\tau ^{*}) +\delta \mathbf{1}, y^{*}(\tau ^{*}))\) for every \(\varepsilon >0\). Then, \(x_{h}^{*}\) is strictly positive, because it belongs to the closure of the upper counter set of \((e_{h}, x^{*}_{-h}(\tau ^{*}) +\delta \mathbf{1}, y^{*}(\tau ^{*}))\), which is included in \({\mathbb {R}}_{++}^{C}\) by Point 4 of Assumption 6.

Step 2.3. Up to a subsequence, \((\lambda ^{\nu },p^{\nu \;\backslash }) _{\nu \in {\mathbb {N}}}\) converges to some \((\lambda ^{*}, p^{*\;\backslash }) \in {\mathbb {R}}_{++}^{H} \times {\mathbb {R}}_{++}^{C-1}\). By \(\Gamma ^{h.1}(\xi ^{\nu },\tau ^{\nu })=0\), fixing commodity C, for every \(\nu \in {\mathbb {N}}\) we have \(\lambda _{h}^{\nu }=D_{x_{h}^{C}}u_{h}(x_{h}^{\nu }, x_{-h}^{\nu }(\tau ^{\nu }),y^{\nu }(\tau ^{\nu }))\). Taking the limit, by Points 1 and 2 of Assumption 6, we get \(\lambda _{h}^{*}:=D_{x_{h}^{C}}u_{h}(x_{h}^{*}, x_{-h}^{*}(\tau ^{*}),y^{*}(\tau ^{*})) >0\).

By \(\Gamma ^{h.1}(\xi ^{\nu },\tau ^{\nu })=0\), for all commodity \(c\ne C\) and for all \(\nu \in {\mathbb {N}}\) we have \(p^{\nu \;c}=\frac{D_{x_{h}^{c}}u_{h}(x_{h}^{\nu }, x_{-h}^{\nu }(\tau ^{\nu }),y^{\nu }(\tau ^{\nu }))}{\lambda _{h}^{\nu }}\). Taking the limit, by Points 1 and 2 of Assumption 6, we get \(p^{*\;c}:=\frac{D_{x_{h}^{c}}u_{h}(x_{h}^{*}, x_{-h}^{*}(\tau ^{*}),y^{*}(\tau ^{*}))}{\lambda _{h}^{*}} >0\). Therefore, \(p^{*\;\backslash } \gg 0\).

Step 2.4. Up to a subsequence, \((\alpha ^{\nu })_{\nu \in {\mathbb {N}}}\) converges to some \(\alpha ^{*} \in {\mathbb {R}}_{++}^{J}\). For every firm j, fix a commodity \(c(j) \in {\mathcal {C}}\). By \(\Gamma ^{j.1}(\xi ^{\nu },\tau ^{\nu })=0\), for every \(\nu \in {\mathbb {N}}\) we have that

$$\begin{aligned} \alpha _{j}^{\nu }= \frac{p^{\nu c(j)}}{ D_{y_{j}^{c(j)}}t_{j}\left( y_{j}^{\nu },y_{-j}^{\nu }\left( \tau ^{\nu }\right) , x^{\nu }\left( \tau ^{\nu }\right) \right) } \end{aligned}$$

which is strictly positive by Point 4 of Assumption 1. Taking the limit, by Points 1 and 4 of Assumption 1, we get \(\alpha _{j}^{*}:= \frac{p^{*\; c(j)}}{D_{y_{j}^{c(j)}}t_{j}(y_{j}^{*},y_{-j}^{*}(\tau ^{*}), x^{*}(\tau ^{*}))}>0\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

del Mercato, E.L., Platino, V. Private ownership economies with externalities and existence of competitive equilibria: a differentiable approach. J Econ 121, 75–98 (2017). https://doi.org/10.1007/s00712-017-0520-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00712-017-0520-1

Keywords

JEL Classification

Navigation