Skip to main content
Log in

Warm-glow giving in networks with multiple public goods

  • Original Paper
  • Published:
International Journal of Game Theory Aims and scope Submit manuscript

Abstract

This paper explores a voluntary contribution game in the presence of warm-glow effects. There are many public goods and each public good benefits a different group of players. The structure of the game induces a bipartite network structure, where players are listed on one side and the public good groups they form are listed on the other side. The main result of the paper shows the existence and uniqueness of a Nash equilibrium. The unique Nash equilibrium is also shown to be asymptotically stable. Then the paper provides some comparative statics analysis regarding pure redistribution, taxation and subsidies. It appears that small redistributions of wealth may sometimes be neutral, but generally, the effects of redistributive policies depend on how public good groups are related in the contribution network structure.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. See, e.g., Brekke et al. (2007) for more stylized examples.

  2. See Becker (1974) for an early analysis of altruism and voluntary contributions.

  3. Bipartite graphs have previously been used, for example, to model economic exchange when buyers have relationships with sellers (Kranton and Minehart 2001), and water extraction when users draw on resource from multiple sources (Ilkiliç 2011).

  4. Furthermore, the comparative statics results involving corner solutions carry over exactly from the pure altruism case with many public goods (see Cornes and Itaya 2010).

  5. Previous results in this literature are restricted to purely altruistic agents. See Kemp (1984), Bergstrom et al. (1986) and Cornes and Itaya (2010) for neutrality and other comparative statics results. For the design of efficient mechanisms, see Cornes and Schweinberger (1996) and Mutuswami and Winter (2004). For the characterization of strategy-proof social choice functions, see Barberà et al. (1991) and Reffgen and Svensson (2012).

  6. Much of this literature is concerned with games in which agents decide how much to contribute to a single public good (i.e., strategies are unidimensional). See Bramoullé and Kranton (2007), Bloch and Zenginobuz (2007) and Bramoullé et al. (2014) for the case of linear best-responses. For the non-linear case, see Bramoullé et al. (2014), Rébillé and Richefort (2014) and Allouch (2015).

  7. An undirected bipartite graph is connected if any two nodes are connected by a path.

  8. This assumption is almost innocuous. See, e.g., Bergstrom et al. (1986, p. 31) for a discussion.

  9. There exist at least three alternative approaches to model impure altruism: one in which people care about the well-being of others (Margolis 1982; Bourlès et al. 2017), another one in which voluntary contributions are subject to a principle of reciprocity (Sudgen 1984), and a third one in which public goods are jointly produced with private goods (Cornes and Sandler 1984).

  10. When \(P=\{p_1\}\), the utility function of agent \(a_i\) reduces to

    $$\begin{aligned} U_i = b_1 \left( G_1 \right) + \delta _{i1} \left( x_{i1} \right) + c_i \left( q_i \right) . \end{aligned}$$

    This specification complies with the assumptions of the usual impure altruism model with a single public good (Andreoni 1990). It is also a special case of the joint production model by Cornes and Sandler (1984). This further indicates that the model developped in this paper is not a direct extension of Bramoullé and Kranton (2007)’s network public good game.

  11. Assumption 1, though, does not prevent the model from exhibiting free-riding effects.

  12. This claim is confirmed by a close inspection of the actual proof of Theorem 1.

  13. This would establish that a contribution increases (resp. decreases) with the number of even (resp. odd) length paths that start from it in the (corresponding undirected) contribution structure.

  14. In particular, it can be shown that all Nash equilibria admitted by the multiple public goods game are solutions to a non-linear complementarity problem (Rébillé and Richefort 2015). See, e.g., Karamardian (1969) for fundamental results in the field.

  15. See, e.g., Definition 4.1 in Khalil (2002).

  16. See, e.g., Cornes and Itaya (2010, p. 364) for a discussion.

  17. Another possible justification for Assumption 2’ may be that agents must be active, even very slightly, to secure their memberships in public good groups. The interiority of the equilibrium would then be the result of group formation processes, not studied in this paper and well worth exploring in future research. See, e.g., Brekke et al. (2007) for the analysis of a group formation game in which group membership is only available to active agents.

  18. An example of such a situation is given in Kemp (1984), in which agents are countries and public goods are international pure public consumption goods or global-level common-pool resources. In this case, warm-glow can be thought of as being a local, country-specific benefit derived from own contribution. For instance, national policy measures to protect the environment provide benefits which are both local (i.e., private) and global (i.e., collective). See, e.g., Kaul et al. (1999) for more details and examples.

  19. For example, quadratic value functions such that

    $$\begin{aligned} \delta _{ij} \left( x_{ij} \right) = \sigma _{ij} x_{ij} - \frac{\theta _j}{2} x_{ij}^2 \quad \text {and} \quad c_i \left( q_i \right) = \xi _i q_i - \frac{\psi }{2} q_i^2 \end{aligned}$$

    for all \(ij \in L\), where \(\sigma _{ij}, \xi _i >0\), \(\theta _j \in (0,\sigma _{ij}/w_i)\) and \(\psi \in (0,\xi _i/w_i)\), fulfil the neutrality condition over the altruism coefficients.

  20. Intuitively, considering incomplete contribution structures almost amounts to relaxing the assumption that \(D = \emptyset \), since clearly inactive links could be practically treated as missing links (see, e.g., Ilkiliç 2011).

  21. The effects of government intervention on the private provision of public goods has a long tradition in economics. The main question is to which extent public provision crowds out private contributions. See, e.g., Abrams and Schmitz (1984), Andreoni (1993), Eckel et al. (2005), Gronberg et al. (2012) and Ottoni-Wilhelm et al. (2014) for empirical studies on this issue.

  22. See, e.g., Theorem 4.7 in Khalil (2002).

References

  • Abrams BA, Schmitz MA (1984) The crowding-out effect of government transfers on private charitable contributions: cross-section evidence. Natl Tax J 37(4):563–568

    Google Scholar 

  • Allouch N (2015) On the private provision of public goods on networks. J Econ Theory 157:527–552

    Article  Google Scholar 

  • Andreoni J (1990) Impure altruism and donations to public goods: a theory of warm-glow giving. Econ J 100:464–477

    Article  Google Scholar 

  • Andreoni J (1993) An experimental test of the public-goods crowding-out hypothesis. Am Econ Rev 83(5):1317–1327

    Google Scholar 

  • Andreoni J (1995) Cooperation in public goods experiments: kindness or confusion? Am Econ Rev 85(4):891–904

    Google Scholar 

  • Andreoni J, Miller J (2002) Giving according to GARP: an experimental test of the consistency of preferences for altruism. Econometrica 70(2):737–753

    Article  Google Scholar 

  • Ballantine CS (1968) Products of positive definite matrices. III. J Algebra 10(2):174–182

    Article  Google Scholar 

  • Barberà S, Sonnenschein H, Zhou L (1991) Voting by committees. Econometrica 59(3):595–609

    Article  Google Scholar 

  • Becker GS (1974) A theory of social interactions. J Political Econ 82(6):1063–1093

    Article  Google Scholar 

  • Bergstrom T, Blume L, Varian H (1986) On the private provision of public goods. J Public Econ 29(1):25–49

    Article  Google Scholar 

  • Bloch F, Zenginobuz U (2007) The effects of spillovers on the provision of local public goods. Rev Econ Des 11(3):199–216

    Google Scholar 

  • Bourlès R, Bramoullé Y, Perez-Richet E (2017) Altruism in networks. Econometrica 85(2):675–689

    Article  Google Scholar 

  • Bramoullé Y, Kranton R (2007) Public goods in networks. J Econ Theory 135(1):478–494

    Article  Google Scholar 

  • Bramoullé Y, Kranton R, D’Amours M (2014) Strategic interaction and networks. Am Econ Rev 104(3):898–930

    Article  Google Scholar 

  • Brekke KA, Nyborg K, Rege M (2007) The fear of exclusion: individual effort when group formation is endogenous. Scand J Econ 109(3):531–550

    Article  Google Scholar 

  • Cornes R, Itaya J-I (2010) On the private provision of two or more public goods. J Public Econ Theory 12(2):363–385

    Article  Google Scholar 

  • Cornes R, Sandler T (1984) Easy riders, joint production, and public goods. Econ J 94:580–598

    Article  Google Scholar 

  • Cornes R, Sandler T (1986) The theory of externalities, public goods and club goods. Cambridge University Press, Cambridge

    Google Scholar 

  • Cornes R, Schweinberger AG (1996) Free riding and the inefficiency of the private production of pure public goods. Can J Econ 29(1):70–91

    Article  Google Scholar 

  • Eckel CC, Grossman PJ, Johnston RM (2005) An experimental test of the crowding out hypothesis. J Public Econ 89(8):1543–1560

    Article  Google Scholar 

  • Feldstein M, Taylor A (1976) The income tax and charitable contributions. Econometrica 44(6):1201–1222

    Article  Google Scholar 

  • Gronberg TJ, Luccasen RA, Turocy TL, Van Huyck JB (2012) Are tax-financed contributions to a public good completely crowded-out? Experimental evidence. J Public Econ 96(7–8):596–603

    Article  Google Scholar 

  • Hochman HM, Rodgers JD (1973) Utility interdependence and income transfers through charity. In: Boulding KE, Pfaff M, Pfaff A (eds) Transfers in an urbanized economy: theories and effects of the grants economy. Wadsworth Publishing Company, Belmont, pp 63–77

    Google Scholar 

  • Holmstrom B (1982) Moral hazard in teams. Bell J Econ 13(2):324–340

    Article  Google Scholar 

  • Ilkiliç R (2011) Networks of common property resources. Econ Theory 47(1):105–134

    Article  Google Scholar 

  • Karamardian S (1969) The nonlinear complementarity problem with applications, part 1. J Optim Theory Appl 4(2):87–98

    Article  Google Scholar 

  • Kaul I, Grunberg I, Stern MA (1999) Global public goods: international cooperation in the 21st century. Oxford University Press, Oxford

    Book  Google Scholar 

  • Kemp MC (1984) A note on the theory of international transfers. Econ Lett 14(2–3):259–262

    Article  Google Scholar 

  • Khalil HK (2002) Nonlinear systems, 3rd edn. Prentice Hall, Upper Saddle River

    Google Scholar 

  • Kranton RE, Minehart DF (2001) A theory of buyer–seller networks. A Econ Rev 91(3):485–508

    Article  Google Scholar 

  • Margolis H (1982) Selfishness, altruism, and rationality. Cambridge University Press, Cambridge

    Google Scholar 

  • Mutuswami S, Winter E (2004) Efficient mechanisms for multiple public goods. J Public Econ 88(3–4):629–644

    Article  Google Scholar 

  • Null C (2011) Warm glow, information, and inefficient charitable giving. J Public Econ 95(5–6):455–465

    Article  Google Scholar 

  • Olson M (1965) The logic of collective action. Harvard University Press, Cambridge

    Google Scholar 

  • Ottoni-Wilhelm M, Vesterlund L, Xie H (2014), Why do people give? Testing pure and impure altruism, NBER working paper no. 20497

  • Palfrey TR, Prisbrey JE (1996) Altruism, reputation and noise in linear public goods experiments. J Public Econ 61(3):409–427

    Article  Google Scholar 

  • Palfrey TR, Prisbrey JE (1997) Anomalous behavior in public goods experiments: how much and why? Am Econ Rev 87(5):829–846

    Google Scholar 

  • Rébillé Y, Richefort L (2014) Equilibrium existence and uniqueness in network games with additive preferences. Eur J Oper Res 232(3):601–606

    Article  Google Scholar 

  • Rébillé Y, Richefort L (2015) Networks of many public goods with non-linear best replies. FEEM Nota di Lavoro 2015:057

    Google Scholar 

  • Reece WS (1979) Charitable contributions: new evidence on household behavior. Am Econ Rev 69(1):142–151

    Google Scholar 

  • Reffgen A, Svensson L-G (2012) Strategy-proof voting for multiple public goods. Theor Econ 7(3):663–688

    Article  Google Scholar 

  • Reinstein DA (2011) Does one charitable contribution come at the expense of another. BE J Econ Anal Policy 11(1):1–54

    Google Scholar 

  • Rosen JB (1965) Existence and uniqueness of equilibrium points for concave \(N\)-person games. Econometrica 33(3):520–534

    Google Scholar 

  • Strang G (1988) Linear algebra and its applications, 3rd edn. Thomson Learning, Boston

    Google Scholar 

  • Sudgen R (1984) The supply of public goods through voluntary contributions. Econ Journal 94:772–787

    Article  Google Scholar 

  • Voorneveld M (2000) Best-response potential games. Econ Lett 66(3):289–295

    Article  Google Scholar 

  • Zhang A, Zhang Y (1996) Stability of a Cournot-Nash equilibrium: the multiproduct case. J Math Econ 26(4):441–462

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lionel Richefort.

Additional information

I would like to thank participants to the 2016 UECE Lisbon Meeting in Game Theory and Applications and the 2017 Coalition Theory Network Workshop in Glasgow for helpful comments. I would also like to thank two anonymous reviewers for their thoughtful review of this paper. All remaining errors are mine.

Appendix

Appendix

Given a contribution structure g, let \({\mathbf {x}}_g\) stand for the column vector of contributions: \({\mathbf {x}}_g\) is the link by link profile of contributions and has size r(g). The links in \({\mathbf {x}}_g\) are sorted in lexicographic order: the contribution \(x_{ij}\) is listed above the contribution \(x_{kl}\) when \(i<k\) or when \(i = k\) and \(j < l\). For the contribution structures \(g_1\) and \(g_2\) given in Fig. 2,

$$\begin{aligned} {\mathbf {x}}_{g_1}= \left( \begin{array}{l} x_{11} \\ x_{12} \\ x_{21} \\ x_{22} \end{array} \right) \quad \text {and} \quad {\mathbf {x}}_{g_2}= \left( \begin{array}{l} x_{11} \\ x_{21} \\ x_{22} \\ x_{32} \end{array} \right) . \end{aligned}$$

The Nash equilibrium of the multiple public goods game is noted \({\mathbf {x}}_g^*\).

Proof of Theorem 1

Because of the budget constraints, the allowed contributions are limited by the requirement that \({\mathbf {x}}_g\) be selected from a convex and compact set S such that

$$\begin{aligned} S = \prod \limits _{ij \in L}{[0,w_i]} \subset {\mathbb {R}}_+^{r(g)}. \end{aligned}$$

Then, the existence of a Nash equilibrium follows from fixed point arguments (such as Kakutani fixed point theorem) as in Theorem 1 of Rosen (1965).

To prove the uniqueness of the Nash equilibrium, Theorems 2 and 6 of Rosen (1965) are applied, which entail that the Nash equilibrium of the multiple public goods game is unique whenever the \(r(g) \times r(g)\) Jacobian matrix of marginal utilities \({\mathbf {J}}({\mathbf {x}}_g)\) is a symmetric negative definite matrix for all \({\mathbf {x}}_g \in S\). Observe that, for all \(ij \in L\),

$$\begin{aligned} \frac{\partial ^2 U_i}{\partial x_{kl} \partial x_{\textit{ij}}} \left( {\mathbf {x}}_g \right) = \left\{ \begin{array}{ll} b''_j \left( G_j \right) + \delta ''_{ij} \left( x_{ij} \right) + c''_i \left( w_i - X_i \right) ,&{}\text { for } kl \in L \text { s.t. } kl = \textit{ij};\\ c''_i \left( w_i - X_i \right) , &{}\text { for } kl \in L \text { s.t. } k = i \hbox { and } l \ne j;\\ b''_j \left( G_j \right) , &{} \text { for } kl \in L \text { s.t. } k \ne i\hbox { and }l = j;\\ 0, &{}\text { for } kl \in L \text { s.t. } k \ne i\hbox { and }l \ne j, \end{array} \right. \end{aligned}$$

so \({\mathbf {J}}( {\mathbf {x}}_g)\) is a symmetric matrix which can be decomposed as

$$\begin{aligned} {\mathbf {J}} \left( {\mathbf {x}}_g \right) = {\mathbf {B}}\left( {\mathbf {x}}_g \right) + \mathbf {\Delta } \left( {\mathbf {x}}_g \right) + {\mathbf {C}} \left( {\mathbf {x}}_g \right) , \end{aligned}$$

where \({\mathbf {B}}({\mathbf {x}}_g)\) is the Jacobian matrix of marginal collective benefits, \(\mathbf {\Delta }({\mathbf {x}}_g)\) is the Jacobian matrix of marginal warm-glow, and \({\mathbf {C}}({\mathbf {x}}_g)\) is the Jacobian matrix of marginal private consumption. Both \({\mathbf {B}}({\mathbf {x}}_g)\), \(\mathbf {\Delta }({\mathbf {x}}_g)\) and \({\mathbf {C}}({\mathbf {x}}_g)\) are symmetric matrices. Moreover, \(\mathbf {\Delta }({\mathbf {x}}_g)\) is a diagonal matrix with all diagonal elements negative since under Assumption 1, \(\delta ''_{ij}(.) < 0\) for all \(ij \in L\). Then, \(\mathbf {\Delta }({\mathbf {x}}_g)\) is negative definite for all \({\mathbf {x}}_g \in S\). In the following lemmas, it is shown that both \({\mathbf {B}}({\mathbf {x}}_g)\) and \({\mathbf {C}}({\mathbf {x}}_g)\) are negative semidefinite for all \({\mathbf {x}}_g \in S\), so \({\mathbf {J}}({\mathbf {x}}_g)\) is a sum of a symmetric negative definite matrix and two symmetric negative semidefinite matrices. Hence, \({\mathbf {J}}({\mathbf {x}}_g)\) is symmetric negative definite for all \({\mathbf {x}}_g \in S\), and uniqueness is established. \(\square \)

Lemma 1

\({\mathbf {B}}({\mathbf {x}}_g)\) is negative semidefinite for all \({\mathbf {x}}_g \in S\).

Proof

To show that \({\mathbf {B}}({\mathbf {x}}_g)\) is negative semidefinite for all \({\mathbf {x}}_g \in S\), it is proved that there exists a matrix \({\mathbf {R}}_g\), with possibly dependent columns, such that \(-{\mathbf {B}}({\mathbf {x}}_g) = {{\mathbf {R}}_g}^\mathsf{T} {\mathbf {R}}_g\) (see Strang 1988, p. 333). Observe that, for all \(ij \in L\),

$$\begin{aligned} - \frac{\partial ^2 b_j}{\partial x_{kl} \partial x_{ij}} \left( {\mathbf {x}}_g \right) = \left\{ \begin{array}{ll} - b''_j \left( G_j \right) , &{} \text { for } kl \in L \text { s.t. } l = j;\\ 0, &{}\text { for } kl \in L \text { s.t. } l \ne j, \end{array} \right. \end{aligned}$$

so \(- {\mathbf {B}}({\mathbf {x}}_g)\) is a symmetric matrix. For \(s \in \{1, \ldots ,m\}\), let \({\mathbf {v}}^s \in {\mathbb {R}}_+^{r(g)}\) be such that

$$\begin{aligned} v_{ij}^s = \left\{ \begin{array}{ll} \sqrt{-b''_j ( G_j )}, &{} \text { for } ij \in L \text { s.t. } j = s;\\ 0, &{} \text { for } ij \in L \text { s.t. } j \ne s. \end{array} \right. \end{aligned}$$

Define \({\mathbf {R}}_g\) as a partitioned matrix such that

$$\begin{aligned} {{\mathbf {R}}_g}^\mathsf{T} = \left( \begin{array}{ccc} {\mathbf {v}}^1&\ldots&{\mathbf {v}}^m \end{array} \right) _{r(g) \times m}. \end{aligned}$$

It is straightforward to check that \(-{\mathbf {B}}({\mathbf {x}}_g) = {{\mathbf {R}}_g}^\mathsf{T} {\mathbf {R}}_g\), so \({\mathbf {B}}({\mathbf {x}}_g)\) is negative semidefinite for all \({\mathbf {x}}_g \in S\). \(\square \)

Lemma 2

\({\mathbf {C}}({\mathbf {x}}_g)\) is negative semidefinite for all \({\mathbf {x}}_g \in S\).

Proof

Let’s prove that there exists a matrix \({\mathbf {R}}_g\) such that \(- {\mathbf {C}}({\mathbf {x}}_g) = {{\mathbf {R}}_g}^\mathsf{T} {\mathbf {R}}_g\). Observe that, for all \(ij \in L\),

$$\begin{aligned} - \frac{\partial ^2 c_i}{\partial x_{kl} \partial x_{ij}} \left( {\mathbf {x}}_g \right) = \left\{ \begin{array}{ll} - c''_i \left( w_i - X_i \right) , &{} \text { for } kl \in L \text { s.t. } k = i;\\ 0, &{}\text { for } kl \in L \text { s.t. } k \ne i, \end{array} \right. \end{aligned}$$

so \(- {\mathbf {C}}({\mathbf {x}}_g)\) is a symmetric matrix. For \(t \in \{1 , \ldots , n\}\), let \({\mathbf {w}}^t \in {\mathbb {R}}_+^{r(g)}\) be such that

$$\begin{aligned} w_{ij}^t = \left\{ \begin{array}{ll} \sqrt{- c''_i ( w_i - X_i )}, &{} \text { for } ij \in L \text { s.t. } i = t; \\ 0, &{} \text { for } ij \in L \text { s.t. } i \ne t. \end{array} \right. \end{aligned}$$

Define \({\mathbf {R}}_g\) as a partitioned matrix such that

$$\begin{aligned} {{\mathbf {R}}_g}^\mathsf{T} = \left( \begin{array}{ccc} {\mathbf {w}}^1&\ldots&{\mathbf {w}}^n \end{array} \right) _{r(g) \times n}. \end{aligned}$$

It is straightforward to check that \(- {\mathbf {C}}({\mathbf {x}}_g) = {{\mathbf {R}}_g}^\mathsf{T} {\mathbf {R}}_g\), so \({\mathbf {C}}({\mathbf {x}}_g)\) is negative semidefinite for all \({\mathbf {x}}_g \in S\). \(\square \)

Proof of Theorem 2

Since asymptotic stability is a local property, it can be assumed that clearly inactive links remain inactive following a small change of contributions at the other links (see, e.g.,Bramoullé et al. 2014; Allouch 2015). Hence, under Assumption 2, the dynamic system reduces to

$$\begin{aligned} {\dot{x}_{\textit{ij}}} = \phi _{ij} \left( G_{-i,j} , w_i - X_{i,-j} \right) - x_{\textit{ij}}, \quad \text {for all }ij \in B. \end{aligned}$$

Let \({\mathbf {Z}} ( {\mathbf {x}}_B )\) be the \(r(B) \times r(B)\) Jacobian matrix of the function \(z_{ij}({\mathbf {x}}_B ) = \phi _{ij} ( G_{-i,j} , w_i - X_{i,-j} ) - x_{ij}\) for all \(ij \in B\). To prove the asymptotic stability of the Nash equilibrium, Lyapunov’s indirect method is applied. This entails that the Nash equilibrium of the multiple public goods game is asymptotically stable whenever the real part of each eigenvalue of \({\mathbf {Z}} ( {\mathbf {x}}^*_B )\) is negative.Footnote 22

For all \(ij \in B\), observe that

$$\begin{aligned} \frac{\partial z_{ij}}{\partial x_{kl}} \left( {\mathbf {x}}_B \right) = \left\{ \begin{array}{ll} - 1 , &{} \text { for } kl \in B \text { s.t. } kl = \textit{ij};\\ \frac{- c''_i \left( w_i - X_i \right) }{b''_j \left( G_j \right) + \delta ''_{\textit{ij}} \left( x_{ij} \right) + c''_i \left( w_i - X_i \right) }, &{} \text { for } kl \in B \text { s.t. } k = i\hbox { and }l \ne j;\\ \frac{- b''_j \left( G_j \right) }{b''_j \left( G_j \right) + \delta ''_{ij} \left( x_{ij} \right) + c''_i \left( w_i - X_i \right) }, &{} \text { for } kl \in B \text { s.t. } k \ne i\hbox { and } l = j;\\ 0, &{} \text { for } kl \in B \text { s.t. } k \ne i\hbox { and }l \ne j, \end{array} \right. \end{aligned}$$

so \({\mathbf {Z}} ( {\mathbf {x}}_B )\) is an asymmetric matrix which can be decomposed as

$$\begin{aligned} {\mathbf {Z}} \left( {\mathbf {x}}_B \right) = {\mathbf {Y}} \left( {\mathbf {x}}_B \right) {\mathbf {J}} ( {\mathbf {x}}_B ) , \end{aligned}$$

where \({\mathbf {J}} ( {\mathbf {x}}_B )\) is the Jacobian matrix of marginal utilities and \({\mathbf {Y}} ( {\mathbf {x}}_B )\) is a diagonal matrix with all diagonal elements positive, i.e.,

$$\begin{aligned} \left[ {\mathbf {Y}} \left( {\mathbf {x}}_B \right) \right] _{ij,ij} = - \frac{1}{b''_j \left( G_j \right) + \delta ''_{ij} \left( x_{ij} \right) + c''_i \left( w_i - X_i \right) } > 0, \quad \text {for all } ij \in B. \end{aligned}$$

Then, \({\mathbf {Y}} ( {\mathbf {x}}_B )\) is a symmetric positive definite matrix for all \({\mathbf {x}}_B \in S_B = \prod \nolimits _{ij \in B}{[0,w_i]}\). It has been shown in the proof of Theorem 1 that under Assumption 1, \({\mathbf {J}} ( {\mathbf {x}}_g )\) is a symmetric negative definite matrix for all \({\mathbf {x}}_g \in S\). Given that any principal submatrix of a symmetric negative definite matrix is symmetric negative definite, \({\mathbf {J}} ( {\mathbf {x}}_B )\) is a symmetric negative definite matrix for all \({\mathbf {x}}_B \in S_B\). It follows that \(- {\mathbf {Z}}({\mathbf {x}}_B)\) is the product of two symmetric positive definite matrices, \({\mathbf {Y}}({\mathbf {x}}_B)\) and \(-{\mathbf {J}}({\mathbf {x}}_B)\). By Theorem 2 in Ballantine (1968), all the eigenvalues of \(- {\mathbf {Z}}({\mathbf {x}}_B)\) are real and positive for all \({\mathbf {x}}_B \in S_B\). Thus, all the eigenvalues of \({\mathbf {Z}}({\mathbf {x}}^*_B)\) are real and negative, and asymptotic stability of the Nash equilibrium is established. \(\square \)

Proof of Proposition 1

Totally differentiating the best-response functions at each link \(ij \in B\) yields

$$\begin{aligned} dx_{ij} = \frac{\partial \phi _{ij}}{\partial G_{-i,j}}dG_{-i,j} + \frac{\partial \phi _{ij}}{\partial (w_i - X_{i,-j})} \left( dw_i - dX_{i,-j} \right) . \end{aligned}$$

It follows that

$$\begin{aligned} dx_{ij} = - \frac{b''_j}{b''_j + \delta ''_{ij} + c''_i} dG_{-i,j} + \frac{c''_i}{b''_j + \delta ''_{ij} + c''_i} \left( dw_i - dX_{i,-j} \right) , \end{aligned}$$

or equivalently, since \(dG_{-i,j} = dG_j - dx_{ij}\),

$$\begin{aligned} dx_{ij} = - \frac{b''_j}{\delta ''_{ij} + c''_i} dG_j + \alpha _{ij} \left( dw_i - dX_{i,-j} \right) . \end{aligned}$$

Since \(C=D=\emptyset \), all links are clearly active, i.e., \(L=B\). Hence, summing across all \(a_i \in N_g(p_j)\) and solving for \(d G_j\) yields

$$\begin{aligned} d G_j = k_j \sum \limits _{a_i \in N_g(p_j)}{\left\{ \alpha _{ij} \left( dw_i - d X_{i, -j} \right) \right\} , \quad \text { for all }\ p_j \in P, } \end{aligned}$$
(1)

where

$$\begin{aligned} k_j = \left( 1 + \sum \limits _{a_i \in N_g (p_j)}{\frac{b''_j}{\delta ''_{ij} + c''_i}} \right) ^{-1} \in \left( 0 , 1 \right] . \end{aligned}$$

Since \(\alpha _{ij} = \alpha _j\) for all \(ij \in L\), Eq. (1) becomes

$$\begin{aligned} dG_j = k_j \alpha _j \sum \limits _{a_i \in N_g(p_j)}{\{ dw_i - dX_{i,-j} \} }, \quad \text {for all }p_j \in P. \end{aligned}$$

Moreover, since g is a complete bipartite graph, it holds that \(N_g(a_i) = P\) for all \(a_i \in A\), and equivalently \(N_g(p_j) = A\) for all \(p_j \in P\). Hence,

$$\begin{aligned} \sum \limits _{a_i \in N_g(p_j)} dw_i = \sum \limits _{a_i \in A} dw_i = 0 \end{aligned}$$

and

$$\begin{aligned} \sum \limits _{a_i \in N_g(p_j)}{dX_{i,-j}} = \sum \limits _{a_i \in A}{dX_{i,-j}} = \sum \limits _{p_l \in P \backslash \{p_j \}}{dG_l}. \end{aligned}$$

It follows that, for all \(p_j \in P\),

$$\begin{aligned} dG_j = - k_j \alpha _j \sum \limits _{p_l \in P \backslash \{p_j \}}{dG_l}. \end{aligned}$$

From this last equation, it appears that

$$\begin{aligned} \sum \limits _{p_l \in P}{dG_l} = \left( 1 - \frac{1}{k_1 \alpha _1} \right) dG_1 = \ldots = \left( 1 - \frac{1}{k_m \alpha _m} \right) dG_m, \end{aligned}$$

so it holds that

$$\begin{aligned} \text {sign}\left( dG_1\right) = \ldots = \text {sign}\left( dG_m\right) . \end{aligned}$$

Then, for all \(p_j \in P\),

$$\begin{aligned} \text {sign}\left( dG_j\right)= & {} \text {sign}\left( \sum \limits _{p_l \in P \backslash \{p_j \}}{dG_l}\right) \\= & {} \text {sign}\left( k_j \alpha _j \sum \limits _{p_l \in P \backslash \{p_j \}}{dG_l}\right) \\= & {} \text {sign}\left( -dG_j\right) \end{aligned}$$

if and only if \(dG_j = 0\). \(\square \)

Proof of Proposition 2

When the contribution structure is complete, a best-response function of the form given is sufficient since identical values of the altruism coefficient among all agents with respect to each public good is sufficient. The remainder of the proof is therefore devoted to the necessary condition.

Since \(C=D=\emptyset \), \(x_{ij} = \phi _{ij}(G_{-i,j} , w_i - X_{i,-j})\) holds for all \(ij \in L\). Moreover, since \(dG_j = 0\) for all \(p_j \in P\), the total differential of the best-response functions given in the proof of Proposition 1 yields

$$\begin{aligned} dx_{ij} = \alpha _j \left( dw_i - dX_{i,-j} \right) , \quad \text { for all }\ ij \in L, \end{aligned}$$

where \(\alpha _j = \alpha _j({\mathbf {x}}_g^*)\). This implies that \(\phi _{ij}(G_{-i,j} , w_i - X_{i,-j})\) is linear in \(w_i - X_{i,-j}\). Then, it holds that

$$\begin{aligned} x_{ij} = \phi _{ij} \left( G_{-i,j} , w_i - X_{i,-j} \right) = \phi ^*_{ij} \left( G_{-i,j} \right) + \alpha _j \left( w_i - X_{i,-j} \right) , \quad \text { for all }\ ij \in L, \end{aligned}$$

where \(\phi ^*_{ij}\) is decreasing since \(\partial \phi _{ij} / \partial G_{-i,j} = -b''_j / (b''_j + \delta ''_{ij} + c''_i) \le 0\). \(\square \)

Proof of Proposition 3

Totally differentiating the best-response functions at each link \(ij \in B\) while keeping \(ds_{ij} = dw_i = 0\) yields

$$\begin{aligned} d {{\tilde{x}}_{ij}}= & {} \frac{\partial \phi _{ij}}{\partial G_{-i,j}}dG_{-i,j} + \frac{\partial \phi _{ij}}{\partial (\frac{\tau _{ij}}{1-s_{ij}})} \times \frac{1}{1-s_{ij}} d\tau _{ij} - \frac{\partial \phi _{ij}}{\partial (w_i - X_{i,-j})} dX_{i,-j}, \end{aligned}$$

or equivalently,

$$\begin{aligned} d {{\tilde{x}}_{ij}}= & {} -\frac{b''_j}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i}dG_{-i,j} + \frac{\frac{\delta ''_{ij}}{(1-s_{ij})^2}}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i}d\tau _{ij} - \frac{c''_i}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i} dX_{i,-j}. \end{aligned}$$

Since \(C=D=\emptyset \), all links are clearly active, i.e., \(L=B\). Hence, rearranging as in the proof of Proposition 1 yields

$$\begin{aligned} d {{\tilde{G}}_j} = {{\tilde{k}}_j} \sum \limits _{a_i \in N_g(p_j)}{\left\{ \left( 1 - {{{\tilde{\alpha }}}_{ij}} \right) d\tau _{ij} - {{{\tilde{\alpha }}}_{ij}} d{{\tilde{X}}_{i,-j}} \right\} }, \quad \text { for all }\ p_j \in P, \end{aligned}$$
(2)

where

$$\begin{aligned} {{\tilde{k}}_j} = \left( 1 + \sum \limits _{a_i \in N_g (p_j)}{\frac{b''_j}{\frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i}} \right) ^{-1} \in \left( 0 , 1 \right] . \end{aligned}$$

Let \(\tau _j = \sum \nolimits _{a_i \in N_g(p_j)}{\tau _{ij}}\) denote the total lump sum taxes with respect to public good \(p_j\). Since the contribution structure is complete and \({{{\tilde{\alpha }}}_{ij}} = {{{\tilde{\alpha }}}_j}\) for all \(ij \in L\), Equation (2) can be rearranged as

$$\begin{aligned} d {{\tilde{G}}_j} = {{\tilde{k}}_j} \left( 1 - {{{\tilde{\alpha }}}_j} \right) d\tau _j - {{\tilde{k}}_j} {{{\tilde{\alpha }}}_j} \sum \limits _{p_l \in P \backslash \{p_j\}}{d{{\tilde{G}}_l}}, \quad \text { for all }\ p_j \in P. \end{aligned}$$

Hence, assuming that \(d\tau _1 \ne 0\) and \(d\tau _l = 0\) for all \(p_l \in P \backslash \{p_1\}\) yields

$$\begin{aligned} d {{\tilde{G}}_1} = {{\tilde{k}}_1} \left( 1 - {{{\tilde{\alpha }}}_1} \right) d\tau _1 - {{\tilde{k}}_1} {{{\tilde{\alpha }}}_1} \sum \limits _{p_l \in P \backslash \{p_1\}}{d{{\tilde{G}}_l}} \end{aligned}$$

and

$$\begin{aligned} d {{\tilde{G}}_l} = - {{\tilde{k}}_l} {{{\tilde{\alpha }}}_l} \sum \limits _{p_j \in P \backslash \{p_l\}}{d{{\tilde{G}}_j}}, \quad \text { for all }\ p_l \in P \backslash \{p_1\}. \end{aligned}$$

From this last equation, it appears that

$$\begin{aligned} \sum \limits _{p_j \in P}{d{{\tilde{G}}_j}} = \left( 1 - \frac{1}{{{\tilde{k}}_2}{{{\tilde{\alpha }}}_2}} \right) d{{\tilde{G}}_2} = ... = \left( 1 - \frac{1}{{{\tilde{k}}_m}{{{\tilde{\alpha }}}_m}} \right) d{{\tilde{G}}_m}. \end{aligned}$$
(3)

Hence, it holds that

$$\begin{aligned} d{{\tilde{G}}_l} = \beta _l d{{\tilde{G}}_1}, \quad \text {for all} p_l \in P \backslash \{p_1\}, \end{aligned}$$
(4)

where

$$\begin{aligned} \beta _l = \left( -\frac{1}{{{\tilde{k}}_l}{{{\tilde{\alpha }}}_l}} - \sum \limits _{p_j \in P \backslash \{p_1,p_l\}}{\left\{ \frac{1 - \frac{1}{{{\tilde{k}}_l}{{{\tilde{\alpha }}}_l}}}{1 - \frac{1}{{{\tilde{k}}_j}{{{\tilde{\alpha }}}_j}}}\right\} } \right) ^{-1} \in (-1,0). \end{aligned}$$

Now, let \(d\tau _1 > 0\) and suppose that \(d{{\tilde{G}}_1} \le 0\). Then, from Equation (4), \(d{{\tilde{G}}_l} \ge 0\) for all \(p_l \in P \backslash \{p_1\}\), and therefore, from Equation (3), \(\sum \nolimits _{p_j \in P}{d{{\tilde{G}}_j}} \le 0\). Hence,

$$\begin{aligned} -d{{\tilde{G}}_1} \ge \sum \limits _{p_l \in P \backslash \{p_1\}}{d{{\tilde{G}}_l}} \ge 0. \end{aligned}$$

It follows that

$$\begin{aligned} \begin{array}{rcl} d{{\tilde{G}}_1} &{} = &{} {{\tilde{k}}_1}\left( 1 - {{{\tilde{\alpha }}}_1} \right) d\tau _1 - {{\tilde{k}}_1}{{{\tilde{\alpha }}}_1}\sum \limits _{p_l \in P \backslash \{p_1\}}{d{{\tilde{G}}_l}}\\ &{}\ge &{} {{\tilde{k}}_1}\left( 1 - {{{\tilde{\alpha }}}_1} \right) d\tau _1 - {{\tilde{k}}_1}{{{\tilde{\alpha }}}_1} \left( -d{{\tilde{G}}_1} \right) \\ &{} = &{} {{\tilde{k}}_1}\left( 1 - {{{\tilde{\alpha }}}_1} \right) d\tau _1 + {{\tilde{k}}_1}{{{\tilde{\alpha }}}_1} d{{\tilde{G}}_1}. \end{array} \end{aligned}$$

Then, it appears that

$$\begin{aligned} d{{\tilde{G}}_1} \left( 1 - {{\tilde{k}}_1}{{{\tilde{\alpha }}}_1} \right) \ge {{\tilde{k}}_1} \left( 1 - {{{\tilde{\alpha }}}_1} \right) d\tau _1 \iff d{{\tilde{G}}_1} \ge \frac{{{\tilde{k}}_1} ( 1 - {{{\tilde{\alpha }}}_1} )}{ 1 - {{\tilde{k}}_1}{{{\tilde{\alpha }}}_1} }d\tau _1 > 0, \end{aligned}$$

a contradiction. The same contradiction can easily be obtained under the assumption that \(d{{\tilde{G}}_1} \ge 0\) when \(d\tau _1 < 0\). Hence, sign\((d\tau _1)\) = sign\((d{{\tilde{G}}_1})\) = sign\((-d{{\tilde{G}}_l})\) for all \(p_l \in P \backslash \{p_1\}\) = sign\((\sum \nolimits _{p_j \in P}{d{{\tilde{G}}_j}})\). \(\square \)

Proof of Proposition 4

Totally differentiating the best-response functions at each link \(ij \in B\) while keeping \(d\tau _{ij} = dw_i = 0\) yields

$$\begin{aligned} d {{\tilde{x}}_{ij}}= & {} \frac{\partial \phi _{ij}}{\partial G_{-i,j}}dG_{-i,j} + \frac{\partial \phi _{ij}}{\partial s_{ij}}ds_{ij} + \frac{\partial \phi _{ij}}{\partial (\frac{\tau _{ij}}{1-s_{ij}})} \times \frac{\tau _{ij}}{(1-s_{ij})^2} ds_{ij} - \frac{\partial \phi _{ij}}{\partial (w_i - X_{i,-j})} dX_{i,-j}, \end{aligned}$$

or equivalently,

$$\begin{aligned} d {{\tilde{x}}_{ij}}= & {} -\frac{b''_j}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i}dG_{-i,j} - \frac{\frac{\delta '_{ij}}{(1-s_{ij})^2} - \frac{\delta ''_{ij} \tau _{ij} }{(1-s_{ij})^3}}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i} ds_{ij} - \frac{c''_i}{b''_j + \frac{\delta ''_{ij}}{(1-s_{ij})^2} + c''_i} dX_{i,-j}. \end{aligned}$$

Since \(C=D=\emptyset \), all links are clearly active, i.e., \(L=B\). Hence, rearranging as in the proof of Proposition 1 yields

$$\begin{aligned} d {{\tilde{G}}_j}= & {} {{\tilde{k}}_j} \sum \limits _{a_i \in N_g(p_j)}{\left\{ \left( {{{\tilde{\alpha }}}_{ij}} \kappa _{ij} + \left( 1 - {{{\tilde{\alpha }}}_{ij}} \right) \frac{\tau _{ij}}{1-s_{ij}} \right) ds_{ij} - {{{\tilde{\alpha }}}_{ij}} d{{\tilde{X}}_{i,-j}} \right\} }, \nonumber \\&\text { for all }\ p_j \in P, \end{aligned}$$
(5)

where

$$\begin{aligned} \kappa _{ij} = \frac{\frac{\partial \phi _{ij}}{\partial s_{ij}}}{\frac{\partial \phi _{ij}}{\partial (w_i - {{\tilde{X}}_{i,-j}})}} = \frac{- \frac{\delta '_{ij}}{(1-s_{ij})^2}}{c''_i} > 0, \end{aligned}$$

and \({{\tilde{k}}_j} \in ( 0 , 1 ]\) as in the proof of Proposition 3. Since the contribution structure is complete and \({{{\tilde{\alpha }}}_{ij}} = {{{\tilde{\alpha }}}_j}\) for all \(ij \in L\), Equation (5) can be rearranged as

$$\begin{aligned} d {{\tilde{G}}_j}= & {} {{\tilde{k}}_j} \sum \limits _{a_i \in A}{\left\{ \left( {{{\tilde{\alpha }}}_j} \kappa _{ij} + \left( 1 - {{{\tilde{\alpha }}}_j} \right) \frac{\tau _{ij}}{1-s_{ij}} \right) ds_{ij}\right\} } - {{\tilde{k}}_j} {{{\tilde{\alpha }}}_j} \sum \limits _{p_l \in P \backslash \{p_j\}}{d{{\tilde{G}}_l}}, \\&\text {for all }p_j \in P. \end{aligned}$$

Hence, assuming that \(ds_{i1} \ne 0\) for at least one agent \(a_i \in N_g(p_1)\) and \(ds_{il} = 0\) for all \(a_i \in N_g(p_l)\) for all \(p_l \in P \backslash \{p_1\}\) yields

$$\begin{aligned} d {{\tilde{G}}_1} = {{\tilde{k}}_1} \sum \limits _{a_i \in A}{\left\{ \left( {{{\tilde{\alpha }}}_1} \kappa _{i1} + \left( 1 - {{{\tilde{\alpha }}}_1} \right) \frac{\tau _{i1}}{1-s_{i1}} \right) ds_{i1}\right\} } - {{\tilde{k}}_1} {{{\tilde{\alpha }}}_1} \sum \limits _{p_l \in P \backslash \{p_1\}}{d{{\tilde{G}}_l}} \end{aligned}$$

and

$$\begin{aligned} d {{\tilde{G}}_l} = - {{\tilde{k}}_l} {{{\tilde{\alpha }}}_l} \sum \limits _{p_j \in P \backslash \{p_l\}}{d{{\tilde{G}}_j}}, \quad \text { for all }\ p_l \in P \backslash \{p_1\}. \end{aligned}$$

From this last equation, observe that Equations (3) and (4) hold, and since

$$\begin{aligned} {{{\tilde{\alpha }}}_1} \kappa _{i1} + \left( 1 - {{{\tilde{\alpha }}}_1} \right) \frac{\tau _{i1}}{1-s_{i1}} > 0, \quad \text { for all }\ a_i \in A, \end{aligned}$$

the same contradiction as in the proof of Proposition 3 can easily be obtained. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Richefort, L. Warm-glow giving in networks with multiple public goods. Int J Game Theory 47, 1211–1238 (2018). https://doi.org/10.1007/s00182-018-0616-z

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00182-018-0616-z

Keywords

JEL Classification

Navigation