Skip to main content
Log in

Reduced-form budget allocation with multiple public alternatives

  • Original Paper
  • Published:
Social Choice and Welfare Aims and scope Submit manuscript

Abstract

We consider a reduced-form implementation problem where two players bargain over how to allocate some resources among a finite set of social alternatives. Each player’s payoff depends on how much of the resources is allocated towards each alternative. Distributional constraints on some targeted groups of alternatives are allowed. Using a network flow approach, we provide a characterization of the reduced-form implementability condition. We show that our results can be useful to study public good problems with budget constraints where a joint implementation of private and public alternatives is possible.

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

Similar content being viewed by others

Notes

  1. Jehiel et al. (1996, 1999) introduce mechanism design problems with allocative externalities, i.e., in auctions with externalities, when a buyer does not obtain the object, his payoff is affected by the identity of the winner.

  2. See for example Armstrong (2000), Brusco and Lopomo (2002), Manelli and Vincent (2010), Hörner and Samuelson (2011), Miralles (2012), Pai and Vohra (2014), Mylovanov and Zapechelnyuk (2017), and Gershkov et al. (2021).

  3. See for example Che et al. (2013) and Budish et al. (2013) for a discussion on hierarchical constraints in auctions and two-sided assignment markets.

  4. For classic analysis on bargaining and public good problems, see for example Myerson and Satterthwaite (1983), Mailath and Postlewaite (1990), Ledyard and Palfrey (1999).

  5. In auctions, the reduced-form approach has been used to characterize the optimal auction when agents have limited budgets, see for example Pai and Vohra (2014).

  6. Budish et al. (2013) obtain a generalization of Birkhoff-von Neumann theorem when a constraint structure forms two laminars. Che et al. (2013) show that group capacity functions preserve submodularity for a laminar family. It can be shown that independent families allow non-submodular quotas. For example, consider Example 3 with \(L_i=6\) and \(B_i=4\). Define \(\bar{b}(G)=\max _{q}\sum _{i\in G}q_i\). We have \(\bar{b}(\{1,2,3\})=9\) \(\bar{b}(i)=5\). Then \(\bar{b}(\{1,2\})+\bar{b}(\{1,3\})<\bar{b}(\{1,2,3\})+\bar{b}(\{1\})\). Hence \(\bar{b}\) is not submodular.

  7. Note that Border’s condition can be interpreted as the probability that a buyer from any set of types wins must not exceed the probability that there exists a buyer with type in that the set.

  8. Goeree and Kushnir (2011, 2016, 2020) obtain a geometric characterization for a general social choice problem with the probability simplex constraint. Lang and Yang (2021) consider bilateral trade and compromise problems and provide a characterization by the projection cone. Lang (2021) considers a two-person bargaining problem in Myerson (1979, 1984) and uses a network flow analysis to obtain a characterization of implementability.

  9. Given the original linear system of feasible allocations, i.e., \( Aq\le b\), we can introduce auxiliary variables \(\tilde{q}^k=(q_+^k, q_-^k)\) and replace \(q^k\) by \(q_+^k-q_-^k\), and add nonnegative constraints \(\tilde{q}^k\). This gives a new system \( \tilde{ A}\tilde{q}\le \tilde{b}\) and \( \tilde{q}\ge 0\). Here \(q_+^k, q_-^k\) denote the resource flows transferred into (out of) alternative k from (to) the designer. For each subgroup G, the allocation constraints require that the net flow of resources arriving alternatives in G is constrained.

  10. With value interdependencies, reduced-from allocation rules alone do not determine interim utilities (excluding transfers) and reduced-form values are desirable for the implementation.

  11. In Theorem 1, condition (4) with the role of the two players interchanged are redundant, as it is implied by condition (3) and (4) in Theorem 1, by adding (3) onto (4).

  12. The set of feasible reduced forms lies in a linear subspace given by condition (3) and hence the representation of the implementation inequalities is non-unique.

  13. Budish et al. (2013) show that if the constraint structure is bihierarchical, then the constraint matrix is totally unimodular.

  14. Here we refer to classic network flow problems with arc capacities. Che et al. (2013) construct a polymatroidal flow network, which allows a richer structure than a classic flow problem such that the constraint matrix needs not to be totally unimodular.

  15. For illustration, let \(M=\begin{bmatrix} 1&{} 1 &{}0 \\ 1&{} 0 &{} 1\\ 0&{} 1 &{} 1\\ \end{bmatrix}.\) Then \(\det (M)=2\) and M is not totally unimodular. Since B contains M as a submatrix, B is also not totally unimodular.

  16. The submoduar flow polyhedron is a generalization of network flows and polymatroid intersections.

  17. That is, find a vector c such that \(c^\top Q>\max \{c^\top x: x\in \mathcal {Q}\}\).

  18. A set function g is modular if for all \(U, U'\subseteq V\), \(g (U)+g(U')=g(U\cap U')+g (U\cup U')\).

  19. Zheng (2021) considers an implementation problem with several types of private goods. He then applies his condition to study a problem with objects and monetary payments with money also being one type of private goods. The implementation problem of private-public goods in this section differs from that in Zheng (2021).

  20. Here the projecting away a variable \(x^1\) from a system \(Ax\le b\) of linear inequalities refers to the creation of another system without the variable \(x^1\) such that both systems have the same solutions over the remaining variables.

  21. For example, outcome A can be implemented by the following weak referendum with appropriate taxes: \(q=2\) or 0 if two agents vote yes or no, \(q=3/2\) if only one agent votes yes; and each agent is taxed to the limit as long as the other agent votes yes, taxed half of the budget if she votes yes but the other votes no, and not taxed if otherwise.

References

  • Alaei S, Fu H, Haghpanah N, Hartline J, Malekian A (2019) Efficient computation of optimal auctions via reduced forms. Math Oper Res 44(3):767–1144

    Article  Google Scholar 

  • Armstrong M (2000) Optimal multi-object auctions. Rev Econ Stud 67(3):455–481

    Article  Google Scholar 

  • Border K (1991) Implementation of reduced form auctions: a geometric approach. Econometrica 59(4):1175–1187

    Article  Google Scholar 

  • Border K (2007) Reduced form auctions revisited. Econ Theor 31:167–181

    Article  Google Scholar 

  • Brusco S, Lopomo G (2002) Collusion via signalling in simultaneous ascending bid auctions with heterogeneous objects, with and without complementarities. Rev Econ Stud 69(2):407–436

    Article  Google Scholar 

  • Budish E, Che Y, Kojima F, Milgrom P (2013) Designing random allocation mechanisms: theory and applications. Am Econ Rev 103(2):585–623

    Article  Google Scholar 

  • Cai Y, Daskalakis C, Weinberg M (2012) Optimal multi-dimensional mechanism design: reducing revenue to welfare maximization. In: Proceedings of 53rd annual IEEE symposium on foundations of computer science, pp 130–139

  • Che Y, Kim J, Mierendorff K (2013) Generalized reduced-form auctions: a network-flow approach. Econometrica 81(6):2487–2520

    Article  Google Scholar 

  • Gershkov A, Moldovanu B, Strack P, Zhang M (2021) A theory of auctions with endogenous valuations. J Polit Econ 129(4):1011–1051

    Article  Google Scholar 

  • Goeree J, Kushnir A (2011) A geometric approach to mechanism design. Working paper

  • Goeree J, Kushnir A (2016) Reduced form implementation with value interdependencies. Games Econ Behav 99:250–256

    Article  Google Scholar 

  • Goeree J, Kushnir A (2020) A geometric approach to mechanism design. Working paper

  • Gopalan P, Nisan N, Roughgarden T (2015) Public projects, boolean functions and the borders of border’s theorem. In: Proceedings of the sixteenth ACM conference on economics and computation, pp 395–415

  • Hart S, Reny P (2015) Implementation of reduced form mechanisms: a simple approach and a new characterization. Econ Theory Bull 3:1–8

    Article  Google Scholar 

  • Hörner J, Samuelson L (2011) Managing strategic buyers. J Polit Econ 119(3):379–425

    Article  Google Scholar 

  • Jehiel P, Moldovanu B, Stacchetti E (1996) How (not) to sell nuclear weapons. Am Econ Rev 86(4):814–829

    Google Scholar 

  • Jehiel P, Moldovanu B, Stacchetti E (1999) Multidimensional mechanism design for auctions with externalities. J Econ Theory 85(2):258–293

    Article  Google Scholar 

  • Kleiner A, Moldovanu B, Strack P (2021) Extreme points and majorization: economic applications. Econometrica 89(4):1557–1593

    Article  Google Scholar 

  • Lang X (2021) Reduced-form allocations with complementarity: a 2-person case. Working paper

  • Lang X, Yang Z (2021) Reduced-form allocations for multiple indivisible objects under constraints. Working paper

  • Ledyard J, Palfrey T (1999) A characterization of interim efficiency with public goods. Econometrica 67(2):435–448

    Article  Google Scholar 

  • Ledyard J, Palfrey T (2002) The approximation of efficient public good mechanisms by simple voting schemes. J Public Econ 83(2):153–171

    Article  Google Scholar 

  • Mailath G, Postlewaite A (1990) Asymmetric information bargaining problems with many agents. Rev Econ Stud 57(1):351–367

    Article  Google Scholar 

  • Manelli A, Vincent D (2010) Bayesian and dominant-strategy implementation in the independent private values model. Econometrica 78(6):1905–1938

    Article  Google Scholar 

  • Maskin E, Riley J (1984) Optimal auctions with risk averse buyers. Econometrica 52(6):1473–1518

    Article  Google Scholar 

  • Matthews S (1984) On the implementability of reduced form auctions. Econometrica 52(6):1519–1522

    Article  Google Scholar 

  • Mierendorff K (2011) Asymmetric reduced form auctions. Econ Lett 110:41–44

    Article  Google Scholar 

  • Miralles A (2012) Cardinal Bayesian allocation mechanisms without transfers. J Econ Theory 147(1):179–206

    Article  Google Scholar 

  • Myerson R (1979) Incentive compatibility and the bargaining problem. Econometrica 47(1):58–73

    Article  Google Scholar 

  • Myerson R (1981) Optimal auction design. Math Oper Res 6(1):58–73

    Article  Google Scholar 

  • Myerson R (1984) Two-person bargaining problems with incomplete information. Econometrica 52(2):461–488

    Article  Google Scholar 

  • Myerson R, Satterthwaite M (1983) Efficient mechanisms and bilateral trading. J Econ Theory 29(2):265–281

    Article  Google Scholar 

  • Mylovanov T, Zapechelnyuk A (2017) Optimal allocation with ex post verification and limited penalties. Am Econ Rev 107(9):2666–2694

    Article  Google Scholar 

  • Pai M, Vohra R (2014) Optimal auctions with financially constrained bidders. J Econ Theory 150(1):383–425

    Article  Google Scholar 

  • Schrijver A (2004) Combinatorial optimization: polyhedra and efficiency. Springer, Berlin

    Google Scholar 

  • Vohra R (2011) Mechanism design: a linear programming approach (econometric society monographs). Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Zheng C (2021) Reduced-form auctions of multiple objects. Working paper

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xu Lang.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

I am grateful to Eric van Damme, Dolf Talman, and Zaifu Yang for their helpful feedbacks, and in particular Debasis Mishra for his numerous comments which greatly improved this paper. I would also like to thank the Associate Editor and one anonymous referee for their insightful comments. Financial support from National Natural Science Foundation of China (NSFC-72033004) and Humanities and Social Sciences Foundation of Ministry of Education of China (18YJC190008) are acknowledged.

Appendix

Appendix

Proof of Corollary 2

We show that if Q satisfies condition (16), Q is implementable by a feasible s-symmetric mechanism. By Theorem 1, condition (16) is sufficient for the existence of a feasible (possibly asymmetric) mechanism q. We can define another feasible mechanism \(q'\) by interchanging the labels of agents and define \(\hat{q}=1/2 q+1/2q'\). By convexity, \(\hat{q}\) is feasible. Moreover, \(\hat{q}\) is s-symmetric. Let \(\hat{Q}\) be the reduced form of \(\hat{q}\). Since Q is s-symmetric, we have \(Q=Q'=\hat{Q}\). Hence Q is implemented by \(\hat{q}\).

Proof of Proposition 4

(1) The monotonicity condition in (a) is familiar from the literature on incentive compatibility. For the implementability, note that the constraint structure is a two-tier laminar, by Theorem 1 and r-symmetry, we have

  1. (i)

    for all \(j\in J\),

    $$\begin{aligned} M_j+N_j-Q_j=0, \end{aligned}$$
    (27)
  2. (ii)
    $$\begin{aligned} \sum _j M_j \lambda _j= \sum _j N_j \lambda _j, \end{aligned}$$
    (28)
  3. (iii)

    for all \(A,B\subseteq J\),

    $$\begin{aligned} \sum _{j\in A}M_j \lambda _j - \sum _{j\in B} N_{j} \lambda _{j}\le \lambda (A)\lambda (B^c). \end{aligned}$$
    (29)

We can use condition (27) to project away variables \(N_j=Q_j-M_j\) from the system. That is, substitute \(N_j\) into conditions (28) and (29), we obtain conditions (25) and (26).

(2) We show that with incentive compatibility we can obtain a reduction of inequalities for condition (26). Note that by incentive compatibility, \(M_j\) is non-decreasing in j. Also \(M_j-Q_j\) is non-increasing in j, since condition (a) holds and \(j/c\le 1\) for all j. By a similar argument as Che et al. (2013), to find the tightest inequalities, we can choose A and B to maximize the set function f(AB) defined by the left hand side of (26), and the optimal solutions are of the form \(A=\{j\in J: M_j\ge \lambda (B^c)\}\) and \(B=\{j\in J: M_j-Q_j\ge -\lambda (A)\}\).\(\square\)

Proof of Proposition 5

For each \(Q\in \mathcal {Q}\), define \(\mathcal {F}(Q)=\{M\in \mathbb {R}^J: (Q,M)\in \mathcal {F}\}\). By construction, \( \mathcal {F}(Q)\) is non-empty, i.e., there exists M such that (QM) is feasible. Then \((Q,M)\in \mathcal {F}\) if and only if \(Q \in \mathcal {Q}\) and \(M\in \mathcal {F}(Q)\). Consider a linear welfare objective \( f(U)=w\cdot U\) with the interim utility weights \(w=(w_j)\in \mathbb {R}_{++}^J\). By quasilinearity, we have \(f(U)= f^1(Q)+f^2(M)\) and the problem becomes \(\max _{(Q,M)\in \mathcal {F}}f^1(Q)+f^2(M)\). We can first find for each \(Q\in \mathcal {Q}\), the solutions to the subproblem \(\max _{M\in \mathcal {F}(Q)}f^2(M)\). It is well-known that for linear programming with the right hand side parameters (vector b in \(Ax\le b\)), the value function depends on b continuously and is piecewise linear. It implies that value function \(F^2(Q)=\max _{M\in \mathcal {F}(Q)}f^2(M)\) depends on Q continuously. We then choose \(Q\in \mathcal {Q}\) to maximize \(f^1(Q)+F^2(Q)\). Because \(F^2(Q)\) is in general nonlinear, the solution can arise at a (possibly non-extreme) boundary point of \(\mathcal {Q}\).\(\square\)

Lemma 2

Let \(J=\{a,b\}\) with \(a<b\le c\). A reduced-form mechanism \((Q,M)\in \mathcal {F}\) if and only if (a) \(Q_a \le Q_b\) and \((a/c) (Q_b- Q_a)\le M_b-M_a\le (b/c) (Q_b- Q_a)\), and (b)

$$\begin{aligned}&2 (M_a \lambda _a+M_b \lambda _b) = Q_a \lambda _a+ Q_b \lambda _b, \end{aligned}$$
(30)
$$\begin{aligned}&2M_a \lambda _a+M_b\lambda _b- Q_a \lambda _a\le \lambda _b, \end{aligned}$$
(31)
$$\begin{aligned}&M_a \lambda _a+M_b\lambda _b- Q_a \lambda _a\le \lambda _b^2, \end{aligned}$$
(32)
$$\begin{aligned}&M_a-Q_a \le 0, \end{aligned}$$
(33)
$$\begin{aligned}&0\le M_a, M_b \le 1. \end{aligned}$$
(34)

Proof of Lemma 2

First notice that in (26), take \(B=\{\emptyset \}\) we obtain \(M_a,M_b\le 1\). Take \(B=\{a,b\}\) and substitute (25) into (26) we have \(M_a,M_b\ge 0\). By Proposition 5 (2), we only need to consider \(A= \{a,b\}, \{b\},\{\emptyset \}\) and \(B=\{\emptyset \}, \{a\},\{a,b\}\). Take all pairs of such (AB) we obtain:

$$\begin{aligned}&M_a \lambda _a+M_b \lambda _b \le 1, \end{aligned}$$
(35)
$$\begin{aligned}&2M_a \lambda _a+M_b\lambda _b- Q_a \lambda _a\le \lambda _b, \end{aligned}$$
(36)
$$\begin{aligned}&2M_a \lambda _a+2M_b\lambda _b- Q_a \lambda _a- Q_b \lambda _b\le 0, \end{aligned}$$
(37)
$$\begin{aligned}&M_b\lambda _b \le \lambda _b, \end{aligned}$$
(38)
$$\begin{aligned}&M_a \lambda _a+M_b\lambda _b- Q_a \lambda _a\le \lambda _b^2, \end{aligned}$$
(39)
$$\begin{aligned}&M_a \lambda _a+2M_b\lambda _b- Q_a \lambda _a- Q_b \lambda _b\le 0, \end{aligned}$$
(40)
$$\begin{aligned}&M_a \lambda _a- Q_a \lambda _a \le 0,\end{aligned}$$
(41)
$$\begin{aligned}&M_a \lambda _a+M_b\lambda _b- Q_a \lambda _a-Q_b \lambda _b\le 0. \end{aligned}$$
(42)

Note that the inequalities (35), (37), (40), (42) are implied by \(0\le M_a,M_b\le 1\) and (30), and thus redundant. We then obtain the conditions in the Lemma.\(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lang, X. Reduced-form budget allocation with multiple public alternatives. Soc Choice Welf 59, 335–359 (2022). https://doi.org/10.1007/s00355-022-01398-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-022-01398-3

Navigation