Skip to main content

Marginal Pricing and Marginal Cost Pricing Equilibria in Economies with Externalities and Infinitely Many Commodities

  • Chapter
  • First Online:
Trends in Mathematical Economics

Abstract

This paper considers a general equilibrium model of an economy in which some firms may exhibit various types of non-convexities in production, there are external effects among agents and the commodity space is infinite dimensional. The consumption sets, the preferences of the consumers and the production possibilities are represented by correspondences in order to take into account the external effects. The firms are instructed to follow the marginal pricing rule from which we obtain an existence theorem. Then, the existence of a marginal cost pricing equilibrium is proved by adding additional assumptions. The simultaneous presence of externalities and infinitely many commodities are sources of technical difficulties when attempting to generalize previous existence results in the literature.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    \(\mathcal{L}_{\infty }(M,\mathcal{M},\mu )\) is the set of equivalence classes of all μ-essentially bounded, \(\mathcal{M}\)-measurable functions on M. Let x be an element of \(\mathcal{L}_{\infty }(M,\mathcal{M},\mu )\), then x ≥ 0 if x(m) ≥ 0 μ-a. e. (almost everywhere); x > 0 if x ≥ 0 and x ≠ 0, and x > > 0 if x(m) > 0 μ-a.e. Hence, if \(x,x\acute{}\) \(\in \mathcal{L}_{\infty }(M,\mathcal{M},\mu )\), then \(x \geq x\acute{}\) (respectively, \(x > x\acute{},\) \(x >> x\acute{}\)) if \(x - x\acute{} \geq 0\) (respectively, \(x - x\acute{} > 0,\) \(x - x\acute{} >> 0\)). \(L_{+} = \left \{x \in L: x \geq 0\right \}\) is the positive cone of L, and \(L_{++} = \left \{x \in L: x > 0\right \}\) is the strict positive cone or the quasi-interior of L. Let A and B be subsets of L. The difference of A and B is defined by A\(B = \left \{x: x \in A\text{ and }x\notin B\right \}\). The open ball of centre x and radius ɛ is \(B\left (x,\varepsilon \right ) = \left \{x^{{\prime}}\in L: \left \Vert x^{{\prime}}- x\right \Vert _{\infty } <\varepsilon \right \}\), while the closed ball of centre x and radius ɛ is \(\overline{B}\left (x,\varepsilon \right ) = \left \{x^{{\prime}}\in L: \left \Vert x^{{\prime}}- x\right \Vert _{\infty }\leq \varepsilon \right \}\).

  2. 2.

    \(\mathcal{L}_{1}(M,\mathcal{M},\mu )\) is classes of all \(\mathcal{M}\)-measurable functions f on M such that \(\int _{m\in M}\left \vert f\left (m\right )\right \vert d\mu \left (m\right ) < \infty \).

  3. 3.

    We say that a production vector y is weakly efficient if and only if \(y \in \partial _{\infty }Y \left (z\right )\). This is equivalent to say that \(\left (\left \{y\right \} + \mathrm{int}L_{+}\right ) \cap Y \left (z\right ) =\emptyset\). A stronger concept is that of efficiency. We say that a production vector y is efficient if and only if \(\left (\left \{y\right \} + L_{+}\right ) \cap Y \left (z\right ) =\emptyset\).

  4. 4.

    \(\nabla _{1}f_{j}\left (y,z\right )\) denotes the gradient vector of f j with respect to y in the sense of Fréchet, that is, \(\nabla _{1}f_{j}\left (y_{j},z\right )\left (x\right ) =\mathop{\lim }\limits_{ t \rightarrow 0}\frac{f_{j}\left (y_{j}+tx,z\right )-f_{j}\left (y_{j},z\right )} {t}\) for all x ∈ L, and the convergence is uniform with respect to x in bounded sets.

  5. 5.

    See Guesnerie (1975).

  6. 6.

    F and F , the topological dual of F, are isomorphic (See MacLane and Garret 1999, Theorem 9, p. 357).

  7. 7.

    Bonnisseau and Médecin 2001, p. 277

  8. 8.

    Let us suppose that \(p_{I^{j}}\left (-y_{I^{j}}\right ) \geq p_{I^{j}}\left (a\right )\) for all \(a \in \cup _{t>0}\mathrm{int}Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\). For any t > 0 and a sufficiently large α > 0, we have \(-y_{I^{j}} +\alpha \chi _{I^{j}} \in \mathrm{int}Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\) by free disposal condition. Hence, \(p_{I^{j}}\left (-y_{I^{j}}\right ) \geq p_{I^{j}}\left (-y_{I^{j}} +\alpha \chi _{I^{j}}\right )\), a contradiction.

References

  • Aliprantis, C., Border, K.: Infinite Dimensional Analysis. Springer, Berlin (1994)

    Book  Google Scholar 

  • Aumann, R.: Existence of competitive equilibria in markets with a continuum of traders. Econometrica 34, 1–17 (1966)

    Article  Google Scholar 

  • Balder, E.: More on equilibria in competitive markets with externalities and a continuum of agents. J. Math. Econ. 44, 575–602 (2008)

    Article  Google Scholar 

  • Beato, P.: The existence of marginal cost pricing equilibria with increasing returns. Q. J. Econ. 97, 669–688 (1982)

    Article  Google Scholar 

  • Beato, P., Mas-Colell, A.: On marginal cost pricing with given tax-subsidy rules. J. Econ. Theory 37, 356–365 (1985)

    Article  Google Scholar 

  • Bewley, T.: Existence of equilibria in economies with infinitely many commodities. J. Econ. Theory 4, 209–226 (1972)

    Article  Google Scholar 

  • Bonnisseau, J.M.: Existence of equilibria in economies with externalities and nonconvexities. Set Valued Anal. 5, 209–226 (1997)

    Article  Google Scholar 

  • Bonnisseau, J.M.: The marginal pricing rule in economies with infinitely many commodities. Positivity 6, 275–296 (2002)

    Article  Google Scholar 

  • Bonnisseau, J.M., Cornet, B.: Existence of marginal cost pricing equilibria in an economy with several non convex firms. Econometrica 58, 661–682 (1990)

    Article  Google Scholar 

  • Bonnisseau, J.M., Cornet, B.: Existence of marginal cost pricing equilibria: the nonsmooth case. Int. Econ. Rev. 31, 685–708 (1990)

    Article  Google Scholar 

  • Bonnisseau, J.M., Créttez, B.: On the characterization of efficient production vectors. Econ Theory 31, 213–223 (2007)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Brown, D.: Equilibrium analysis with non-convex technologies. In: Hildenbrand, W., Sonnenschein, H. (eds.) Handbook of Mathematical Economics, vol. 4, pp. 1963–1995. North-Holland, Amsterdam (1991)

    Google Scholar 

  • Clarke, F.: Optimization and Nonsmooth Analysis. Wiley, New York (1983)

    Google Scholar 

  • Cornet, B.: Existence of equilibria in economies with increasing returns. In: Cornet, B., Tulkens, H. (eds.) Contributions to Operations Research and Economics: The XXth Anniversary of CORE, pp. 79–97. The MIT Press, Cambridge (1990)

    Google Scholar 

  • Cornet, B., Topuzu, M.: Existence of equilibria for economies with externalities and a measure space of consumers. Econ. Theory 26, 397–421 (2005)

    Article  Google Scholar 

  • Dunford, N., Schwarz, J.: Linear Operators, Part I. Wiley-Interscience, New York (1958)

    Google Scholar 

  • Fuentes, M.: Existence of equilibria in economies with externalities and non-convexities in an infinite dimensional commodity space. J. Math. Econ. 47, 768–776 (2011)

    Article  Google Scholar 

  • Guesnerie, R.: Pareto-optimality in nonconvex economies. Econometrica 43, 1–29 (1975)

    Article  Google Scholar 

  • Guesnerie, R.: First best allocation of resources with non convexities in production. In: Cornet, B., Tulkens, H. (eds.) Contributions to Operations Research and Economics: The XXth Anniversary of CORE, pp. 99–143. The MIT Press, Cambridge (1990)

    Google Scholar 

  • Hotelling, H.: The general welfare in relation to problems of taxation and of railway and utility rates. Econometrica 6, 242–269 (1938)

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Laffont, J.: Effets externes et théorie économique. Monographies du Séminaire d’Econométrie. Editions du CNRS, París (1977)

    Google Scholar 

  • Laffont, J.J., Laroque, G.: Effets externes et théorie de l’équilibre general. Cahiers du Séminaire d’Econométrie, vol. 14. Editions du CNRS, París (1972)

    Google Scholar 

  • MacLane, S., Garret, B.: Algebra, 3rd edn. AMS Chelsea Publishing, New York (1999)

    Google Scholar 

  • Mantel, R.: Equilibrio con rendimientos crecientes a escala. An. Asoc. Argentina de Economía Política 1, 271–283 (1979)

    Google Scholar 

  • Mas-Colell, A.: The price equilibrium existence problem in topological vector lattices. Econometrica 54, 1039–1054 (1986)

    Article  Google Scholar 

  • Mas-Colell, A., Zame, W.: Equilibrium theory in infinite dimensional spaces. In: Hildenbrand, W., Sonnenschein, H. (eds.) Handbook of Mathematical Economics, vol. 4, pp. 1836–1898. North-Holland, Amsterdam (1991)

    Google Scholar 

  • Mas-Colell, A., Whinston, M., Green, J.: Microeconomic Theory. Oxford University Press, New York (1995)

    Google Scholar 

  • Noguchi, M., Zame, W.: Competitive markets with externalities. Theor. Econ. 1, 143–166 (2006)

    Google Scholar 

  • Rustichini, A., Yannelis, N.: What is perfect competition. In: Khan, M.A., Yannelis, N. (eds.) Equilibrium Theory in Infinite Dimensional Spaces. Springer, New York (1991)

    Google Scholar 

  • Schaefer, H.H., Wolf, M.P.: Topological Vector Spaces, 2nd edn. Springer, New York (1999)

    Book  Google Scholar 

  • Shannon C.: Increasing returns in infinite-horizon economies. Rev. Econ. Stud. 64, 73–96 (1996)

    Article  Google Scholar 

  • Villar, A.: Curso de Microeconomía Avanzada. Un enfoque de equilibrio general. Antoni Bosch, Barcelona (1997)

    Google Scholar 

Download references

Acknowledgements

Earlier versions of this paper were presented at the Seminario de Funciones Generalizadas, Universidad de Buenos Aires (2013); the Primer Workshop en Economía Matemática, Universidad de San Andrés (2014); the XLIX Reunión Anual de la Asociación Argentina de Economía Política, Universidad Nacional de Posadas (2014); and the 15th SAET Conference on Current Trends in Economics, University of Cambridge (2015). I would like to thank their audiences. I also wish to thank Juan José Martínez and two anonymous reviewers whose comments and suggestions have improved the quality of this paper. Mistakes and other shortcomings are, of course, entirely my own.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matías Fuentes .

Editor information

Editors and Affiliations

Appendix

Appendix

7.1.1 Proof of Proposition 1

Before proving Proposition 1, we show that, given Assumption P, for every z ∈ L m+n and every j, \(\overline{\mathrm{int}Y _{j}\left (z\right )} = Y _{j}\left (z\right )\). Let y belong to \(Y _{j}\left (z\right )\). If y belongs to \(\mathrm{int}Y _{j}\left (z\right )\), then y belongs to \(\overline{\mathrm{int}Y _{j}\left (z\right )}\). If y belongs to \(\partial _{\infty }Y _{j}\left (z\right )\), then for all ɛ > 0, \(y - \frac{\varepsilon } {2}\chi _{M}\) belongs to \(\mathrm{int}Y _{j}\left (z\right )\) by free disposal. Consequently, \(B\left (y,\varepsilon \right ) \cap \mathrm{int}Y _{j}\left (z\right )\neq \varnothing \) for all ɛ > 0, and thus, y belongs to \(\overline{\mathrm{int}Y _{j}\left (z\right )}.\)

We now prove that the correspondence Y j : L m+nL is \(\left (\prod \nolimits _{L^{m+n}}\sigma ^{\infty },\mathcal{T}\right )\)-l.h.c. From the above result and Lemma 14.21 in Aliprantis and Border (1994), it is enough to prove that intY j : L m+nL is \(\left (\prod \nolimits _{L^{m+n}}\sigma ^{\infty },\mathcal{T}\right )\)-l.h.c. Let y ∈ \(\mathrm{int}Y _{j}\left (z\right )\) and let z α be a net which \(\prod \nolimits _{L^{m+n}}\sigma ^{\infty }-\) converges to z. Since f j is \(\sigma ^{\infty }\times \prod \nolimits _{L^{m+n}}\sigma ^{\infty }\)-continuous, there exists α 0 ∈ Γ such that α > α 0 implies \(f_{j}\left (y,z^{\alpha }\right ) < 0\). Hence, there exists a net y α \(\left (= y\right ) \in \mathrm{int}Y _{j}\left (z^{\alpha }\right )\) for all α and y α → y. 

The weak* closeness of Y j is immediate from Assumption P(ii).

Remark 3.

Given Assumption P, if the correspondence Y j is convex valued, then \(\overline{\mathrm{int}Y _{j}\left (z\right )} = Y _{j}\left (z\right )\) without free disposal requirement (Schaefer and Wolf 1999, p. 38, 1.3). On the other hand, we can repeat the argument made above to show that the correspondence Y j F: F m+nF is l.h.c.

7.1.2 Proof of Proposition 2

We omit the index j in order to simplify the notation. We first state the following Lemma:

Lemma 4.

For a given \(\overline{z} \in L^{m+n}\) , let \(T_{Y \left (\overline{z}\right )}\left (\overline{y}\right )\) be the Clarke tangent cone of \(Y \left (\overline{z}\right )\) at \(\overline{y}\) . Let \(v \in T_{Y \left (\overline{z}\right )}\left (\overline{y}\right )\) and δ > 0. There exist weak* open neighbourhoods of \(\overline{z}\) and \(\overline{y}\), \(W^{\overline{z}}\) and \(W\overline{^{y}}\) , respectively, such that for all ɛ > 0, for all \(z \in W^{\overline{z}}\) and for all \(y \in W\overline{^{y}} \cap B\left (\overline{y},\varepsilon \right ),\) \(v + \overline{y} - y -\delta \chi _{M} \in T_{Y \left (z\right )}\left (y\right )\).

Proof.

Given \(\overline{z} \in Z,\) we have to prove that \(\nabla _{1}f\left (\overline{y},\overline{z}\right )\left (v + \overline{y} - y -\delta \chi _{M}\right ) \leq 0\). Let \(0 <\alpha < \frac{\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right )} {2\left (\left \Vert v\right \Vert +\varepsilon +\delta \right )}.\) From Assumption P(v), there exists a \(\prod \nolimits _{L^{m+n+1}}\sigma ^{\infty }-\) open neighbourhood of \(\left (\overline{z},\overline{y}\right ),\) \(U^{\overline{z}} \times U\overline{^{y}},\) such that for all \(\left (z,y\right ) \in U^{\overline{z}} \times U\overline{^{y}}\), \(\left \vert \nabla _{1}f\left (y,z\right ) -\nabla _{1}f\left (\overline{y},\overline{z}\right )\right \vert <\alpha\). Let us consider the following σ −open neighbourhood of \(\overline{y,}\)

$$\displaystyle{V \overline{^{y}} = \left \{y \in L: \left \vert \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\overline{y} - y\right )\right \vert < \frac{\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right )} {2} \right \}}$$

Let \(W\overline{^{y}} = U\overline{^{y}} \cap V \overline{^{y}}\), \(W^{\overline{z}} = U^{\overline{z}}\) and ɛ > 0. For all \(\left (y,z\right ) \in W\overline{^{y}} \cap B\left (\overline{y},\varepsilon \right ) \times W^{\overline{z}}\),

$$\displaystyle\begin{array}{rcl} \nabla _{1}f\left (y,z\right )\left (v + \overline{y} - y -\delta \chi _{M}\right )& =& \left (\nabla _{1}f\left (y,z\right ) -\nabla _{1}f\left (\overline{y},\overline{z}\right )\right )\left (v + \overline{y} - y -\delta \chi _{M}\right ) {}\\ & & +\nabla _{1}f\left (\overline{y},\overline{z}\right )\left (v + \overline{y} - y -\delta \chi _{M}\right ) {}\\ & & <\alpha \left (\left \Vert v\right \Vert +\varepsilon +\delta \right ) {}\\ & +& \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (v\right ) + \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\overline{y} - y\right ) -\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right ) {}\\ & & < \frac{\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right )} {2} + \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (v\right ) {}\\ & & +\frac{\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right )} {2} -\delta \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (\chi _{M}\right ) {}\\ & =& \nabla _{1}f\left (\overline{y},\overline{z}\right )\left (v\right ) \leq 0 {}\\ \end{array}$$

We now proceed to the proof of Proposition 2. Let \(\left (z^{\alpha },\pi ^{\alpha }\right )_{\left (\varGamma,\leq \right )}\) be a net of Z × S, \(\prod \nolimits _{L^{m+n}}\sigma ^{\infty }\times \sigma ^{ba}-\) converging to \(\left (\overline{z},\overline{\pi }\right )\). Let \(v \in T_{Y \left (\overline{z}\right )}\left (\overline{y}\right )\) and δ > 0. There exist ɛ > 0 and α 0 ∈ Γ such that for all α > α 0, \(y^{\alpha } \in B\left (0,\varepsilon \right )\). We note that \(\left \Vert y^{\alpha } -\overline{y}\right \Vert <\varepsilon +\left \Vert \overline{y}\right \Vert =\varepsilon ^{{\prime}}\). Hence, for all α > α 0, \(y^{\alpha } \in B\left (\overline{y},\varepsilon ^{{\prime}}\right )\). From the above lemma, there exist weak*-open neighbourhoods of \(\overline{z}\) and \(\overline{y}\), \(W^{\overline{z}}\) and \(W\overline{^{y}}\), respectively, such that for ɛ  > 0 and all α > α 0, \(\left (y^{\alpha },z^{\alpha }\right ) \in W\overline{^{y}} \cap B\left (\overline{y},\varepsilon ^{{\prime}}\right ) \times W^{\overline{z}}\) and \(v + \overline{y} - y^{\alpha } -\delta \chi _{M} \in T_{Y \left (z^{\alpha }\right )}\left (y^{\alpha }\right )\).

Since \(\pi ^{\alpha } \in N_{Y \left (z^{\alpha }\right )}\left (y^{\alpha }\right )\), \(\pi ^{\alpha }\left (v + \overline{y} - y^{\alpha } -\delta \chi _{M}\right ) \leq 0\) for all α > α 0. Passing to the limit, we obtain \(\overline{\pi }\left (v\right ) + \overline{\pi }\left (\overline{y}\right ) -\mathop{\lim }\limits_{\alpha }\pi ^{\alpha }\left (y^{\alpha }\right )-\delta \leq 0\). Since 0 ∈ \(T_{Y \left (\overline{z}\right )}\left (\overline{y}\right )\), \(\overline{\pi }\left (\overline{y}\right ) \leq \mathop{\lim }\limits_{\alpha }\pi ^{\alpha }\left (y^{\alpha }\right )+\delta\), and since this inequality holds true for all δ > 0, we have \(\overline{\pi }\left (\overline{y}\right ) \leq \mathop{\lim }\limits_{\alpha }\pi ^{\alpha }\left (y^{\alpha }\right )\).

Let \(v \in T_{Y \left (\overline{z}\right )}\left (\overline{y}\right ).\) If \(\mathop{\lim }\limits_{\alpha }\pi ^{\alpha }\left (y^{\alpha }\right ) = \overline{\pi }\left (\overline{y}\right ),\) then \(\overline{\pi }\left (v\right ) \leq 0\). Consequently, \(\overline{\pi } \in N_{Y \left (\overline{z}\right )}\left (\overline{y}\right ) \cap S\) since π α ∈ S for all α.

7.1.3 Proof of Proposition 4

We first state and prove the following lemma, which is used in the proof of Proposition 4. To simplify, we suppress index j.

Lemma 5.

Let \(p_{I} \in \left (L^{I}\right )_{+}^{{\ast}},\) then there exists \(\hat{p}_{I} \in L_{+}^{{\ast}}\) such that \(\hat{p}_{I}\left (x\right ) = p_{I}\left (x^{I}\right )\) if x ∉ L O and \(\hat{p}_{I}\left (x\right ) = 0\) if x ∈ L O.

Proof.

Let \(p_{I} \in \left (L^{I}\right )_{+}^{{\ast}}\). By a classical extension theorem, there exists a functional \(\tilde{p}_{I} \in L_{+}^{{\ast}}\), and hence, a measure \(\tilde{v}_{I}\) ∈ ba + M, , μ such that \(\tilde{p}_{I}\left (x\right ) =\int _{m\in M}x\left (m\right )d\tilde{v}_{I}\left (m\right )\) and \(p_{I}\left (x\right ) =\tilde{ p}_{I}\left (x\right )\) for all x ∈ L I, since L and \(ba\left (M,\mathcal{M},\mu \right )\) are isometrically isomorphic (Dunford and Schwarz 1958). We now define the measure \(\hat{v}_{I}\) as:

\(\hat{v}_{I}\left (A\right ) = \left \{\begin{array}{c} \tilde{v}_{I}\left (A^{I}\right )\quad \text{ if }A \subsetneq O \\ 0\quad \text{ otherwise} \end{array} \right.\).

One easily checks that \(\hat{v}_{I} \in ba^{+}\left (M,\mathcal{M},\mu \right )\) which is identified with a functional \(\hat{p}_{I} \in L_{+}^{{\ast}}\). Take xL O. There exists M  ⊂ I such that \(\mu \left (M^{{\prime}}\right )\neq 0\) and \(x^{I}\left (m\right )\neq 0\) for all m ∈ M . Consequently, \(\hat{p}_{I}\left (x\right ) =\hat{ p}_{I}\left (x^{I}\right )+\hat{p}_{I}\left (x^{O}\right ) =\int _{m\in M}x^{I}\left (m\right )d\hat{v}_{I}\left (m\right )+\int _{m\in M}x^{O}\left (m\right )d\hat{v}_{I}\left (m\right ) =\int _{m\in I}x^{I}\left (m\right )d\hat{v}_{I}\left (m\right ) =\int _{m\in I}x^{I}\left (m\right )d\tilde{v}_{I}\left (m\right ) =\tilde{ p}_{I}\left (x^{I}\right ) = p_{I}\left (x^{I}\right ).\) If x ∈ L O, \(\hat{p}_{I}\left (x\right ) =\int _{m\in M}x^{O}\left (m\right )d\hat{v}_{I}\left (m\right ) = 0\).

Remark 4.

The above lemma can be rewritten in terms of the subspace \(\left (L^{O}\right )^{{\ast}}\) as follows: for every \(p_{O} \in \left (L^{O}\right )_{+}^{{\ast}},\) there exists a functional \(\hat{p}_{O} \in L_{+}^{{\ast}}\) such that \(\hat{p}_{O}\left (x\right ) = p_{O}\left (x^{O}\right )\) if xL I and \(\hat{p}_{O}\left (x\right ) = 0\) if x ∈ L I.

First, we claim that for all t > 0, \(-y_{I^{j}}\) does not belong to the relative interior of \(Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\). Otherwise, \(y_{j} \in \mathrm{int}Y _{j}\left (z\right )\). We also note that for all t > 0, the relative interior of \(Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\) is non-empty. Finally, since for all t > 0, \(Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\) is convex, \(\cup _{t>0}\mathrm{int}Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\) is open, non-empty and convex (Schaefer and Wolf 1999, p. 38, 1.2).

Since \(-y_{I^{j}}\notin \cup _{t>0}\mathrm{int}Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\), there exists a continuous linear functional \(p_{I^{j}} \in \left (L^{I^{_{j}} }\right )_{+}^{{\ast}}\) such that \(p_{I^{j}}\left (-y_{I^{j}}\right ) \leq p_{I^{j}}\left (a\right )\forall a \in \cup _{t>0}\mathrm{int}Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\),Footnote 8 whence \(p_{I^{j}}\left (-y_{I^{j}}\right ) \leq p_{I^{j}}\left (a^{{\prime}}\right )\) for all a  ∈ ∪ t > 0 \(Y _{j}\left (y_{O^{j}}+\right.\) \(\left.t\chi _{O^{j}},z\right )\). Consequently, \(p_{I^{j}}\left (-y_{I^{j}}\right ) = c_{j}\left (p_{I^{j}},y_{O^{j}} + t\chi _{O^{j}},z\right )\) since \(-y_{I^{j}} \in Y _{j}\left (y_{O^{j}} + t\chi _{O^{j}},z\right )\) for all t > 0. By the above lemma, we can extend the functional \(p_{I^{j}}\) to an element of L + —denoted by \(p_{I^{j}}\) as well—such that \(p_{I^{j}}\left (\xi \right ) = 0\) for all \(\xi \in L^{O^{j} }\). Let \(p_{O^{j}} = \bigtriangledown _{O^{j}}c_{j}\left (p_{I^{j}},y_{O^{j}},z\right ) \in \left (L_{+}^{O^{j} }\right )^{{\ast}}\). We also extend \(p_{O^{j}}\) to L + —denoted by \(p_{O^{j}}\) as well—such that \(p_{O^{j}}\left (\xi \right ) = 0\) for all \(\xi \in L^{I^{j} }\). Consequently, by Lemma 2, \(p_{j} = p_{I^{j}} + p_{O^{j}} \in N_{Y }{}_{j}\left (z\right )\left (y_{j}\right )\). Since, \(\left (z,\pi \right )\) is a marginal pricing equilibrium, π = λ p j for some λ > 0. Hence, \(\pi _{I^{j}} =\lambda p_{I^{j}}\) and \(\pi _{O^{j}} =\lambda p_{O^{j}} =\lambda \bigtriangledown _{O^{j}}c_{j}\left (p_{I^{j}},y_{O^{j}},z\right ) = \bigtriangledown _{O^{j}}c_{j}\left (\lambda p_{I^{j}},y_{O^{j}},z\right ) = \bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}},z\right )\). Consequently, \(\pi _{I^{j}}\left (-y_{I^{j}}\right ) = c_{j}\left (\pi _{I^{j}},y_{O^{j}},z\right )\) and \(\pi =\pi _{I^{j}} + \bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}},z\right ) \in N_{Y _{j}\left (z\right )}\left (y_{j}\right )\). Hence, conditions a., b’. and c’. of Definition 2 are satisfied.

Remark.

We point out that Bonnisseau and Cornet show that if \(\left (z,\pi \right )\) is a marginal pricing equilibrium, then there exists a vector \(\left (w_{j}\right )_{j=1}^{n}\) ∈ L n (our notation) defined as \(w_{j} = y_{I^{j}} + y_{O^{j}}^{+}\), such that \(\left (\left (x_{i}\right )_{i=1}^{m},\left (w_{j}\right )_{j=1}^{n},\pi \right )\) is a marginal cost pricing equilibrium. A significant difference between our approach and theirs is that in their case, \(\left (\left (x_{i}\right )_{i=1}^{m},\left (w_{j}\right )_{j=1}^{n}\right ) \in \varPi _{i=1}^{m}X_{i} \times \varPi _{j=1}^{n}Y _{j}\), while in ours, if z ∈ Z, x i i = 1 m, w j j = 1 n may not be in \(\varPi _{i=1}^{m}X_{i}\left (\left (\left (x_{i}\right )_{i=1}^{m},\left (w_{j}\right )_{j=1}^{n}\right )\right ) \times \varPi _{j=1}^{n}Y _{j}\left (\left (\left (x_{i}\right )_{i=1}^{m},\left (w_{j}\right )_{j=1}^{n}\right )\right )\) since the sets are not comparable. This justifies Assumption SPP.

Another important difference with the above paper is that, even if ​x i i = 1 m, w j j = 1 n belongs to \(\varPi _{i=1}^{m}X_{i}\left (\left (x_{i}\right )_{i=1}^{m},\left (z_{j}\right )_{j=1}^{n}\right ) \times \varPi _{j=1}^{n}Y _{j}\left (\left (\left (x_{i}\right )_{i=1}^{m},\left (z_{j}\right )_{j=1}^{n}\right )\right )\), we cannot prove that \(\bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}},z\right ) = \bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}}^{+},\right.\)   \(\left.z\right )\) as they did, since the argument they constructed does not work in Fréchet derivatives in infinite-dimensional spaces. Consequently, in the present context, \(\bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}},z\right ) = \bigtriangledown _{O^{j}}c_{j}\left (\pi _{I^{j}},y_{O^{j}}^{+},z\right )\) whenever \(y_{O^{j}} = y_{O^{j}}^{+}\) which also justifies the Assumption SPP.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Fuentes, M. (2016). Marginal Pricing and Marginal Cost Pricing Equilibria in Economies with Externalities and Infinitely Many Commodities. In: Pinto, A., Accinelli Gamba, E., Yannacopoulos, A., Hervés-Beloso, C. (eds) Trends in Mathematical Economics. Springer, Cham. https://doi.org/10.1007/978-3-319-32543-9_7

Download citation

Publish with us

Policies and ethics