This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Extreme Points in Multi-Dimensional Screening

Patrick Lahr  and  Axel Niemeyer
Abstract.

This paper characterizes extreme points of the set of incentive-compatible mechanisms for screening problems with linear utility. Extreme points are exhaustive mechanisms, meaning their menus cannot be scaled and translated to make additional feasibility constraints binding. In problems with one-dimensional types, extreme points admit a tractable description with a tight upper bound on their menu size. In problems with multi-dimensional types, every exhaustive mechanism can be transformed into an extreme point by applying an arbitrarily small perturbation. For mechanisms with a finite menu, this perturbation displaces the menu items into general position. Generic exhaustive mechanisms are extreme points with an uncountable menu. Similar results hold in applications to delegation, veto bargaining, and monopoly problems, where we consider mechanisms that are unique maximizers for specific classes of objective functionals. The proofs involve a novel connection between menus of extreme points and indecomposable convex bodies, first studied by [46].


JEL Codes: D82, D44, D86, C78, C65

Keywords: Multi-Dimensional Types, Extreme Points, Exposed Points, Indecomposable Convex Bodies, Multi-Good Monopoly Problem, Linear Delegation, Linear Veto Bargaining

Acknowledgments: We thank Felix Bierbrauer, Peter Caradonna, Simone Cerreia-Vioglio, Gregorio Curello, Laura Doval, Mira Frick, Alkis Georgiadis-Harris, Daniel Gottlieb, Michael Greinecker, Nima Haghpanah, Deniz Kattwinkel, Andreas Kleiner, Elliot Lipnowski, Fabio Maccheroni, Alejandro Manelli, Jeffrey Mensch, Benny Moldovanu, Roger Myerson, Efe Ok, Tom Palfrey, Martin Pollrich, Luciano Pomatto, Justus Preußer, Kota Saito, Larry Samuelson, Fedor Sandomirskiy, Mario Schulz, Ina Taneva, Omer Tamuz, Tristan Tomala, Aleh Tsyvinsky, Rakesh Vohra, and Mark Whitmeyer, as well as seminar audiences at ASU, Columbia, the Paris Game Theory seminar, Paris 1, UCR, and USC for helpful discussions and comments.
Lahr: ENS Paris-Saclay, Department of Economics. Email: patrick.lahr@ens-paris-saclay.fr
Niemeyer: Caltech, Division of the Humanities and Social Sciences. Email: niemeyer@caltech.edu

1. Introduction

Much of the mechanism design literature assumes that agents’ preferences can be described by a single dimension of private information. Under this assumption, the theory has delivered remarkably clean predictions for optimal mechanisms across various applications. However, in many environments, agents’ preferences are more realistically modeled assuming multiple dimensions of private information, for instance, in allocation problems with multiple heterogeneous goods or collective decision problems with several alternatives. Despite their importance, much less is known about multi-dimensional settings. Several results highlight an inherent complexity of optimal mechanisms in these settings, but explicit descriptions have not been obtained outside of a few special cases.111See, for example, [104], [80, 81], [56], [35], [36], or [55].

In this paper, we study the structure of optimal mechanisms for a class of mechanism design problems featuring one- and multi-dimensional types. Specifically, we consider linear screening problems. A principal makes an allocation that affects their own and an agent’s utility. Both parties’ utilities are linear in allocations and depend on the agent’s type, where the allocation space and type space are convex sets in Euclidean space. Linear screening covers a range of problems with and without transfers, for example, monopoly and bilateral trade problems or delegation and veto bargaining problems.

Our main results characterize the extreme points of the set of incentive-compatible (IC) mechanisms for linear screening problems. Since the principal maximizes a linear functional—their expected utility—over the set of IC mechanisms, an optimal mechanism can always be found among the extreme points. While every optimal mechanism is a mixture over optimal extreme points, generic objective functionals are uniquely maximized at an extreme point.222We show that the set of IC mechanisms is norm-compact and convex. The first claim then follows from Choquet’s theorem. The second claim follows from a theorem by [76] (where genericity is in a topological sense). Moreover, essentially every extreme point is the unique maximizer of some objective functional.333This claim follows from a theorem by Straszewicz and Klee ([67]). More precisely, the mechanisms that are uniquely optimal for some instance of the principal’s problem, i.e., exposed points, are dense in the set of extreme points. Thus, determining the structure of optimal mechanisms across instances of the principal’s problem is tantamount to determining the structure of the extreme points.

The extreme-point approach has seen successful applications in a number of other mechanism design settings, but with the sole exception of [81] (MV), it has not been applied to settings with multi-dimensional types.444See, for example, [23], [82], [70], [96, 97], or [125]. Although MV laid important groundwork for the monopoly problem, our characterizations reveal more explicit insights into the structure of extreme points and apply to a broader class of problems.555 We provide a detailed discussion of our relation to MV in Section 9.

Our main insight is that in every one-dimensional problem, the set of extreme points admits a tractable description, whereas in every multi-dimensional problem, the set of extreme points is virtually as rich as the set of all incentive-compatible mechanisms. An important observation is that every extreme point is exhaustive: the allocations made by the mechanism—its menu—cannot be scaled and translated to make additional feasibility constraints binding.666A feasibility constraint is an affine restriction on the set of feasible allocations, i.e., a halfspace. In one-dimensional problems, extreme points admit a tight upper bound on their menu size on top of exhaustiveness. In contrast, in multi-dimensional problems, every exhaustive mechanism can be transformed into an extreme point by applying an arbitrarily small perturbation. For exhaustive mechanisms with a finite menu, this perturbation simply displaces the menu items into general position. In particular, generic exhaustive mechanisms are extreme points.

1.1. Discussion

A common explanation for the difficulty with multi-dimensional screening is that binding incentive constraints depend on the choice of mechanism, making it a priori unclear which constraints will be binding in an optimal mechanism; our results corroborate this explanation. The perturbation described in the previous paragraph modifies the binding incentive constraints of an exhaustive mechanism for an arbitrarily small set of types. Thus, since exhaustive mechanisms are defined only in terms of binding feasibility constraints, the qualitative properties that distinguish extreme points from other mechanisms are essentially only properties of binding feasibility constraints. In contrast, for all one-dimensional problems, properties of binding incentive constraints impose significant restrictions on the structure of the extreme points, e.g., by limiting their menu size to no more than a few allocations in typical applications.

A potential concern is that our results characterize the structure of optimal mechanisms across all instances of the principal’s problem, i.e., for arbitrary utility functions and beliefs about the agent’s type, while in some applications, the principal’s utility function is known. For example, when a monopolist maximizes revenue, certain extreme points are suboptimal for every belief of the monopolist about the agent’s valuations. Our main insights remain the same in sample applications where the principal’s utility is fixed and state-independent, such as in the monopoly problem. In particular, with multi-dimensional types, we show that the extreme points that are (uniquely) optimal for some belief of the principal are again virtually as rich as the set of all IC mechanisms.

Our results offer some insights into the capabilities and limitations of the classical mechanism design paradigm. An important pillar for the success of the theory is that, in many applications, it makes predictions for optimal mechanisms that are independent of the specific details of the environment. We confirm that such predictions are obtainable for all one-dimensional linear screening problems, whereas they are largely unattainable for all multi-dimensional linear screening problems. When the structure of the optimal mechanism depends too finely on the model parameters, it is difficult to derive tangible practical guidance and testable implications from the theory since parameters such as type distributions may be unknown or unobservable in practice.

We emphasize that we do not provide a full solution to multi-dimensional linear screening in that we do not identify the optimal mechanism for each instance of the principal’s problem and show how this mechanism varies across instances. However, given the overwhelming complexity of the structure of extreme points, it seems implausible that such comparative statics exercises are feasible in full generality.

1.2. Technical Contributions

We obtain our results by establishing a connection between extreme points of the set of IC mechanisms and extremal elements of certain spaces of convex sets. Instead of studying the set of IC mechanisms or the agent’s associated indirect utility functions,777For the indirect-utility approach, see e.g. [103], [104], [80, 81], and [36]. we study the space of all menus that the principal could offer the agent. By the well-known taxation principle, any IC mechanism is the agent’s choice function from some menu of allocations and vice versa. Since preferences are linear, offering the agent a menu is payoff-equivalent to offering the agent the menu’s convex hull. Thus, we can establish a bijection between payoff-equivalence classes of IC mechanisms and certain convex sets contained in the allocation space. We show that this bijection preserves convex combinations (in the sense of Minkowski) and therefore preserves extreme points. Analogous bijections hold onto the set of indirect utility functions.

The extremal elements of the space of compact convex sets in Euclidean space are relatively well understood in the mathematical literature and are referred to as indecomposable convex bodies, first studied by [46]. Most of our results are derived from translating these mathematical insights into economic insights via the connection between IC mechanisms and menus in the form of convex sets. Two kinds of complications arise in this translation. First, feasibility requires that menus are contained in the space of allocations; these constraints are not generally considered in the literature on indecomposability. Second, certain menus are equivalent from the agent’s perspective when the type space is restricted, i.e., when the agent’s preferences are constrained to a subset of all linear preferences.

Indecomposable convex bodies in the plane are points, line segments, and triangles, but they are so plentiful and complex in higher dimensions that a complete description has not been obtained and is not to be expected.888[115] writes (p. 166): “Most [(in the sense of topological genericity)] convex bodies in d\mathbb{R}^{d}, d3d\geq 3, are smooth, strictly convex and indecomposable. It appears that no concrete example of such a body is explicitly known. This is not too surprising, since it is hard to imagine how such a body should be described.” We note that algebraic characterizations of indecomposable polytopes are known; see [92, 88, 118]. We provide a characterization along these lines in Appendix B. However, what is known in the mathematical literature is enough to obtain the relevant economic insights we present in this paper. The complexity of indecomposable convex bodies in two- versus higher dimensions mirrors the dichotomy between one- and multi-dimensional screening problems since, with linear utility and up to redundancies, an allocation space of a given dimension always corresponds to a type space of one dimension less. (Transfers would here be counted as an allocation dimension of its own.)

1.3. Structure of the Paper

Section 2 introduces relevant notation and mathematical definitions. Section 3 introduces the model. Section 4 gives a characterization of extreme points in terms of mechanisms that make an inclusion-wise maximal set of incentive and feasibility constraints binding. Section 5 clarifies the role of feasibility constraints by defining and characterizing exhaustive mechanisms. Section 6 presents our core results for one- versus multi-dimensional problems, along with several supporting results. Section 7 introduces the relevant mathematical tools and sketches the proof of our core results in the context of a delegation problem among lotteries over finitely many alternatives, with an emphasis on the special role of the three-alternative case.999Problems with three alternatives have been considered as the simplest departure from the two-alternative case often studied in the literature on mechanism design without transfers; see [25]. Section 8 discusses applications to monopolistic selling and veto bargaining, including essentially complete characterizations of undominated mechanisms in the sense of [81] for these settings. Section 9 provides an extensive discussion of the related literature, including multi-dimensional screening, extreme points in mechanism design, delegation and veto bargaining, and the mathematical foundations underlying this paper. Section 10 concludes.

Appendix A collects several auxiliary results, including the translation between the set of IC mechanisms and a certain space of convex sets. Appendix B deals with the geometry of the set of finite-menu mechanisms and provides an algebraic characterization of finite-menu extreme points (which generalizes the main result in [81]). Appendix C provides a complete characterization of extreme points for one-dimensional problems omitted from the main text for brevity. Appendix D contains the proofs for all results in the main text.

2. Notation and Mathematical Definitions

Let XX be a subset of a topological vector space EE. Δ(X)\Delta(X) denotes the set of Borel probability measures on XX. intX\operatorname{int}X denotes the interior of XX, bndrX\operatorname{bndr}X denotes the boundary of XX, and clX\operatorname{cl}X denotes the closure of XX. convX\operatorname{conv}X denotes the convex hull, coneX\operatorname{cone}X denotes the conical hull, and affX\operatorname{aff}X denotes the affine hull.

Suppose XEX\subseteq E is convex. extX\operatorname{ext}X denotes the set of extreme points of XX, i.e., those xXx\in X for which x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} and λ(0,1)\lambda\in(0,1) implies x=x=x′′x=x^{\prime}=x^{\prime\prime}. expX\exp X denotes the set of exposed points of XX, i.e., those xXx\in X for which there exists a continuous linear functional f:Ef:E\to\mathbb{R} such that f(x)>f(x)f(x)>f(x^{\prime}) for all xXx^{\prime}\in X, xxx\neq x^{\prime}. Every exposed point is extreme, but the converse is not generally true. A face ff of XX is a convex subset of XX such that for all xfx\in f, x,x′′Xx^{\prime},x^{\prime\prime}\in X, and λ(0,1)\lambda\in(0,1), x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} implies x,x′′fx^{\prime},x^{\prime\prime}\in f. The set XEX\subseteq E is a polytope if it is the convex hull of finitely many (extreme) points.

We use the following standard terminology for convex sets in Euclidean space. A convex body KdK\subset\mathbb{R}^{d} is a non-empty compact convex set. A polyhedron PdP\subseteq\mathbb{R}^{d} is the finite intersection of closed halfspaces. A polyhedral cone is a cone that is also a polyhedron. A polytope in Euclidean space is a bounded polyhedron. Every face ff of a polyhedron PP can be represented as f=argmaxaPaθf=\operatorname*{arg\,max}_{a\in P}a\cdot\theta for some θd\theta\in\mathbb{R}^{d}. A face ff is proper if fPf\neq P. A vertex vv of PP is a face of dimension 0, i.e., an extreme point of PP.101010The dimension of a convex set XdX\subseteq\mathbb{R}^{d}, denoted dimX\dim X, is the dimension of its affine hull. A facet FF of PP is a face of PP such that dimF=dimP1\dim F=\dim P-1. If PdP\subseteq\mathbb{R}^{d} is dd-dimensional, then the facet-defining hyperplane of FF is the unique supporting hyperplane H={ydynHcH}H=\{y\in\mathbb{R}^{d}\mid y\cdot n_{H}\leq c_{H}\} of PP such that FHF\subseteq H, where nHn_{H} is the outer (unit) normal vector to PP on FF.

3. Model and Preliminaries

3.1. Allocations and Types

There is a principal and an agent. The principal chooses an allocation aAda\in A\subset\mathbb{R}^{d}, where AA is a dd-dimensional polytope. The principal’s preferences over allocations depend on the agent’s private information, their type θΘd{0}\theta\in\Theta\subset\mathbb{R}^{d}\setminus\{0\}, where the set {λθθΘ,λ+}\{\lambda\theta\mid\theta\in\Theta,\,\lambda\in\mathbb{R}_{+}\} of all rays through the type space Θ\Theta is a dd-dimensional polyhedral cone. We say that the type space is unrestricted if coneΘ=d\operatorname{cone}\Theta=\mathbb{R}^{d}. An agent of type θΘ\theta\in\Theta derives utility aθa\cdot\theta from allocation aAa\in A. Given the agent’s type θΘ\theta\in\Theta, the principal derives utility av(θ)a\cdot v(\theta) from allocation aAa\in A, where v:Θdv:\Theta\to\mathbb{R}^{d} is a bounded objective function that captures the conflict of interest between both parties. There may be a veto allocation a¯extA\underaccent{\bar}{a}\in\operatorname{ext}A that the agent can enforce unilaterally.

Remark.

The model subsumes several screening problems as special cases; see Sections 7 and 8 for examples. In particular, we subsume problems with transferable utility by interpreting one allocation dimension as a numeraire for which the principal and the agent have a known marginal utility. That is, Θ=Θ~×{1}\Theta=\tilde{\Theta}\times\{-1\}, v(θ)=(,1)v(\theta)=(\ldots,1) for all θΘ\theta\in\Theta, and A=A~×[0,κ]A=\tilde{A}\times[0,\kappa], where κ\kappa\in\mathbb{R} is the total endowment of the numeraire.

Since utility is linear, we can identify types on the same ray from the origin because they have the same preferences over the allocations in AA. We select normalized types in the unit sphere 𝕊d1={yd:y=1}\mathbb{S}^{d-1}=\{y\in\mathbb{R}^{d}:\>||y||=1\} as canonical representatives, i.e., Θ𝕊d1\Theta\subseteq\mathbb{S}^{d-1}. In applications, we occasionally make other selections, e.g., when considering transferable utility. Thus, in our model, a dd-dimensional allocation space AA always corresponds to a (d1)(d-1)-dimensional type space Θ\Theta.111111Contrary to other notions of one-dimensionality in the mechanism design literature (see e.g. [24, Chapter 5.6]), a one-dimensional type space need here not imply a linear order on the underlying preferences. For example, Θ=𝕊1\Theta=\mathbb{S}^{1} may be a circle.

3.2. Mechanisms

The principal designs a (direct and measurable) mechanism x:ΘAx:\Theta\to A to screen the agent.121212It is without loss of generality to consider deterministic mechanisms: every randomized allocation in Δ(A)\Delta(A) can be replaced with its barycenter since both principal and agent have linear utility. In applications, we may think of the allocation space AA as a set of lotteries over an underlying finite set of alternatives. In this case, a mechanism can be interpreted as a stochastic mechanism. A mechanism asks the agent to report their type θ\theta and then implements an allocation x(θ)x(\theta). By the revelation principle, it is without loss of generality for the principal to focus on mechanisms that are incentive-compatible (IC) and individually rational (IR):

x(θ)θ\displaystyle x(\theta)\cdot\theta x(θ)θ\displaystyle\geq x(\theta^{\prime})\cdot\theta\quad θ,θΘ;\displaystyle\forall\theta,\theta^{\prime}\in\Theta; (IC)
x(θ)θ\displaystyle x(\theta)\cdot\theta a¯θ\displaystyle\geq\underaccent{\bar}{a}\cdot\theta\quad θΘ.\displaystyle\forall\theta\in\Theta. (IR)

IC means that the agent has no incentive to misreport their type. IR means the agent has no incentive to veto the principal’s choice. To simplify the analysis, we assume that there exists a type θ¯Θ\underaccent{\bar}{\theta}\in\Theta for whom the veto allocation is one of their favorite allocations, i.e., a¯argmaxaAaθ¯\underaccent{\bar}{a}\in\operatorname*{arg\,max}_{a\in A}a\cdot\underaccent{\bar}{\theta}. If no veto allocation exists, IR is satisfied by convention.

An optimal mechanism is any solution to the principal’s problem

supx:ΘA\displaystyle\sup_{x:\Theta\to A} Θ(x(θ)v(θ))𝑑μ\displaystyle\int_{\Theta}(x(\theta)\cdot v(\theta))\,d\mu (OPT)
s.t. (IC) and (IR),\displaystyle\quad\eqref{eq:IC}\text{ and }\eqref{eq:IR},

where μΔ(Θ)\mu\in\Delta(\Theta) is the principal’s belief about the agent’s type. We assume that μ\mu admits a bounded probability density, i.e., is absolutely continuous.

We say that a set of (IC) and (IR) mechanisms is a candidate set for optimality if it contains an optimal mechanism for every objective function vv and belief μ\mu of the principal.

3.3. Menus and Payoff-Equivalence

Instead of designing a mechanism, the principal can equivalently offer the agent a menu (or delegation set) MAM\subseteq A, with a¯M\underaccent{\bar}{a}\in M, from which the agent may choose their favorite allocation. That is,

x(θ)argmaxaMaθx(\theta)\in\operatorname*{arg\,max}_{a\in M}a\cdot\theta

defines an IC and IR mechanism x:ΘAx:\Theta\to A (if maximizers exist). The value function U(θ)=θx(θ)U(\theta)=\theta\cdot x(\theta) is the agent’s indirect utility function associated with the mechanism xx.

Mechanisms defined by the same menu are payoff-equivalent, i.e., the associated indirect utility functions are the same. For IC mechanisms, it can be shown that payoff-equivalence is equivalent to equality almost everywhere (Corollary A.5).131313 Almost everywhere equality is with respect to the spherical measure (since Θ𝕊d1\Theta\subseteq\mathbb{S}^{d-1}). For a Borel subset B𝕊d1B\subseteq\mathbb{S}^{d-1}, the spherical measure is proportional to the Lebesgue measure of the set {λθθB,λ[0,1]}\{\lambda\theta\mid\theta\in B,\,\lambda\in[0,1]\}. Thus, payoff-equivalent mechanisms yield the principal the same expected utility since the belief μ\mu is absolutely continuous.

We define the (essential) menu

menu(x)=cl{x(Θ)x satisfies (IC) and (IR) and is payoff-equivalent to x}\operatorname{menu}(x)=\operatorname{cl}\bigcap\{x^{\prime}(\Theta)\mid x^{\prime}\text{ satisfies \eqref{eq:IC} and \eqref{eq:IR} and is payoff-equivalent to }x\}

associated with an IC and IR mechanism as (the closure of) the set of allocations that are commonly made by all mechanisms in its payoff-equivalence class. For example, if the menu size |menu(x)||\operatorname{menu}(x)| is finite, then the menu simply consists of the allocations that are made by the mechanism with strictly positive probability (cf. [36, Definition 7]).

We henceforth identify payoff-equivalent mechanisms, i.e., x=xx=x^{\prime} if x(θ)=x(θ)x(\theta)=x^{\prime}(\theta) for almost every θΘ\theta\in\Theta, and write 𝒳\mathcal{X} for the set of payoff-equivalence classes of IC and IR mechanisms.141414See Section A.1 for a brief discussion of tie-breaking. In Section A.2, we show that 𝒳\mathcal{X} is L1L^{1}-compact and convex. Therefore, a solution to (OPT) exists and can be found among the extreme points of 𝒳\mathcal{X} (Bauer’s maximum principle).

4. Binding Incentive and Feasibility Constraints

In this section, we provide a characterization of the extreme points of the set of IC and IR mechanisms in terms of binding incentive and feasibility constraints. Optimal mechanisms solve a linear optimization problem, and therefore, identifying the binding constraints is crucial for finding a solution. This perspective will prove useful in the subsequent sections.

An (IC) constraint is represented by a pair of types (θ,θ)Θ×Θ(\theta,\theta^{\prime})\in\Theta\times\Theta, and we define

𝒞(x)={(θ,θ)Θ×Θ|argmaxamenu(x)θaargmaxamenu(x)θa}\mathcal{IC}(x)=\left\{(\theta,\theta^{\prime})\in\Theta\times\Theta\ \middle|\ \operatorname*{arg\,max}_{a\in\operatorname{menu}(x)}\theta^{\prime}\cdot a\subseteq\operatorname*{arg\,max}_{a\in\operatorname{menu}(x)}\theta\cdot a\right\} (1)

as the set of binding IC constraints of mechanism x𝒳x\in\mathcal{X}. This definition considers a constraint as binding if type θ\theta is indifferent to mimicing type θ\theta^{\prime} regardless of how θ\theta^{\prime} breaks ties.151515Since ties are null events, 𝒞(x)\mathcal{IC}(x) coincides for every type θ\theta and almost every deviation θ\theta^{\prime} with defining an IC constraint as binding if x(θ)θ=x(θ)θx(\theta)\cdot\theta=x(\theta^{\prime})\cdot\theta. The latter definition of binding constraints is not robust to tie-breaking.

To define feasibility constraints, recall that the allocation space AA is a polytope. Thus, there exists a finite set \mathcal{F} of facet-defining hyperplanes H={ynynH=cH}H=\{y\in\mathbb{R}^{n}\mid y\cdot n_{H}=c_{H}\} of AA. That is, A={H:H}A=\bigcap\{H_{-}:\>H\in\mathcal{F}\}, where H={ynynHcH}H_{-}=\{y\in\mathbb{R}^{n}\mid y\cdot n_{H}\leq c_{H}\} are the associated halfspaces containing AA. Each halfspace corresponds to an affine restriction on the space of available allocations, and no restriction is redundant given the others; see Figure 1 for an illustration.

We define

(x)={Hmenu(x)H}\mathcal{F}(x)=\{H\in\mathcal{F}\mid\operatorname{menu}(x)\cap H\neq\emptyset\} (2)

as the set of binding feasibility constraints of mechanism xx.

Individual rationality constraints are irrelevant for the formulation of the following result; see Section A.3 for an explanation.

Theorem 4.1.

A mechanism x𝒳x\in\mathcal{X} with finite menu size is an extreme point of 𝒳\mathcal{X} if and only if there is no other mechanism x𝒳x^{\prime}\in\mathcal{X} such that (x)(x)\mathcal{F}(x)\subseteq\mathcal{F}(x^{\prime}) and 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}).

Proof.

See Section D.1. ∎

H2H_{2}H1H_{1}H5H_{5}H4H_{4}H3H_{3}AAmenu(x)\operatorname{menu}(x)
Figure 1. The menu of a mechanism x𝒳x\in\mathcal{X} which is not an extreme point. The allocation space AA is a polytope defined by five facet-defining hyperplanes ={H1,,H5}\mathcal{F}=\{H_{1},\ldots,H_{5}\}. The four allocations marked with dots are the menu of the mechanism. The two rightmost allocations in the menu can be translated horizontally while maintaining the orientation of all dotted lines, keeping the set of binding constraints unchanged. This is clear for the feasibility constraints, and can be seen for the incentive constraints because each type—represented by a direction in 2\mathbb{R}^{2}—still chooses the same menu item(s).
Remark.

The inclusions (x)(x)\mathcal{F}(x)\subseteq\mathcal{F}(x^{\prime}) and 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}) in Theorem 4.1 can equivalently be replaced by the equalities (x)=(x)\mathcal{F}(x)=\mathcal{F}(x^{\prime}) and 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}).

A mechanism with finite menu size is an extreme point if and only if it is the only mechanism that makes a given inclusion-wise maximal set of constraints binding; Figure 1 illustrates. Of the two types of constraints, binding feasibility constraints are easier to analyze and will be treated separately in the next section.

Let us briefly discuss the proof of Theorem 4.1. If x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} is a finite menu mechanism in 𝒳\mathcal{X}, where λ(0,1)\lambda\in(0,1) and x,x′′𝒳x^{\prime},x^{\prime\prime}\in\mathcal{X}, then 𝒞(x)=𝒞(x)𝒞(x′′)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime})\cap\mathcal{IC}(x^{\prime\prime}).161616For almost all type pairs, this is immediate from the definition of the (IC) constraints. See Lemma D.1 for a complete argument. Thus, an important object for understanding extreme points is the set {x𝒳𝒞(x)𝒞(x)}\{x^{\prime}\in\mathcal{X}\mid\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime})\} of mechanisms that make an inclusion-wise larger set of IC constraints binding than a given finite-menu mechanism x𝒳x\in\mathcal{X}. We show that this set is a polytope and a face of 𝒳\mathcal{X}; in particular, xext𝒳x\in\operatorname{ext}\mathcal{X} if and only if xext{x𝒳𝒞(x)𝒞(x)}x\in\operatorname{ext}\{x^{\prime}\in\mathcal{X}\mid\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime})\}. Extreme points of a polytope are uniquely determined by their incident facets, i.e., binding constraints. Thus, xext𝒳x\in\operatorname{ext}\mathcal{X} if and only if xx is uniquely determined by its binding feasibility constraints within the face, which completes the proof.

The result does not extend to mechanisms with infinite menu size because the relevant face is no longer a polytope.171717For example, one can show the existence of strictly incentive-compatible extreme points x,xextXx,x^{\prime}\in\operatorname{ext}X, i.e., 𝒞(x)=𝒞(x)=\mathcal{IC}(x)=\mathcal{IC}(x^{\prime})=\emptyset, that make the same feasibility constraints binding, i.e., (x)=(x)\mathcal{F}(x)=\mathcal{F}(x^{\prime}). In the linear delegation problem discussed in Section 7, this amounts to showing that there exist smooth and indecomposable convex bodies, i.e., extended menus, that touch the same facets of the unit simplex, which follows by arguments similar to those in the proof of Theorem 6.6; see [115, Theorems 2.7.1 and 3.2.18 ]. All our subsequent results will nevertheless accommodate mechanisms of infinite menu size.

Remark.

The required steps for the proof outlined in the previous paragraph generalize the main results of [81, Theorems 17,19, 20, and 24 ] about extreme points of the multi-good monopoly problem to arbitrary linear screening problems; see Appendix B.

5. Exhaustive Mechanisms

In this section, we introduce and characterize exhaustive mechanisms and show that every extreme point is exhaustive. Exhaustiveness allows us to isolate the role of binding feasibility constraints in determining which mechanisms are extreme points. Our main results in the next section will clarify the role of binding incentive constraints.

Definition 5.1.

Mechanisms x,x𝒳x,x^{\prime}\in\mathcal{X} are positively homothetic if there exists λ++\lambda\in\mathbb{R}_{++} and tdt\in\mathbb{R}^{d} such that x=λx+tx=\lambda x^{\prime}+t. Mechanisms x,x𝒳x,x^{\prime}\in\mathcal{X} are homothetic if they are positively homothetic or one of them is constant. A mechanism x𝒳x\in\mathcal{X} is exhaustive if there does not exist a mechanism x𝒳x^{\prime}\in\mathcal{X} positively homothetic to xx such that (x)(x)\mathcal{F}(x)\subseteq\mathcal{F}(x^{\prime}).

Two mechanisms are (positively) homothetic if one can be obtained from the other by scaling (with a strictly positive scalar) and translation. In geometric terms, a positive homothety leaves invariant the “shape” and “orientation” of menus. In economic terms, a positive homothety leaves invariant the agent’s ordinal preferences over menu items and, in particular, the binding incentive constraints. Positive homothethy defines an equivalence relation on 𝒳\mathcal{X} and every equivalence class of positively homothetic mechanisms contains an exhaustive mechanism, but this mechanism need not be unique; see Figure 2.

menu(x)\operatorname{menu}(x)menu(x)\operatorname{menu}(x^{\prime})
Figure 2. Two menus of exhaustive mechanisms homothetic to each other.
Theorem 5.2.

Every extreme point xext𝒳x\in\operatorname{ext}\mathcal{X} is exhaustive.

Proof.

See Section D.2. ∎

For mechanisms with finite menu size, Theorem 5.2 is a corollary of Theorem 4.1. If xx is not exhaustive, then there exists a mechanism x𝒳x^{\prime}\in\mathcal{X} positively homothetic to xx such that (x)(x)\mathcal{F}(x)\subseteq\mathcal{F}(x^{\prime}). 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}) follows immediately from the definition of positive homothety. Therefore, xx is not uniquely pinned down by its binding constraints. If xx has a finite menu, then Theorem 4.1 completes the proof by contraposition. In general, the argument in Section D.2 shows that a mechanism that leaves slack in the feasibility constraints can be decomposed into mechanisms homothetic to itself.

We proceed by characterizing the set of exhaustive mechanisms more explicitly. This characterization is important since every property of exhaustive mechanisms is also a property of extreme points and hence of optimal mechanisms.181818While every optimal mechanism is a mixture over optimal extreme points, exhaustiveness is not necessarily preserved under convex combinations. Thus, technically not every optimal mechanism for a given instance (v,μ)(v,\mu) of the principal’s problem need be exhaustive. However, topologically generic linear objective functionals are uniquely maximized at an extreme point ([76]). Thus, optimal mechanisms are generically exhaustive. Recall that nHn_{H} is the normal vector of the facet-defining hyperplane HH\in\mathcal{F} of the allocation space AA.

Theorem 5.3.

A mechanism x𝒳x\in\mathcal{X} is exhaustive if and only if one of the following holds:

  1. (1)

    There exists aextAa\in\operatorname{ext}A such that menu(x)={a}\operatorname{menu}(x)=\{a\}.

  2. (2)

    (a) span{nH}H(x)=d\operatorname{span}\{n_{H}\}_{H\in\mathcal{F}(x)}=\mathbb{R}^{d} and (b) H(x)H=\bigcap\limits_{H\in\mathcal{F}(x)}H=\emptyset.

Proof.

See Section D.2. ∎

That is, a non-constant mechanism is exhaustive if and only if the facet-defining hyperplanes corresponding to the binding feasibility constraints satisfy two conditions: (a) their normal vectors span the ambient space and (b) they have an empty intersection. These conditions ensure that the mechanism can neither be translated or scaled relative to a point in a way that would make additional feasibility constraints binding. Figure 3 illustrates.

menu(x)\operatorname{menu}(x)
H(x)H\bigcap_{H\in\mathcal{F}(x^{\prime})}Hmenu(x)\operatorname{menu}(x^{\prime})
Figure 3. Illustrations of conditions (a) and (b) from Theorem 5.3. Left: condition (a) is violated by a menu touching two parallel facets of a rectangle. The menu can be translated horizontally until it touches an inclusion-wise larger set of facets. Right: condition (b) is violated by a menu touching only two facets of a pentagon. The menu can be scaled relative to the intersection point of the two facet-defining lines until it touches an inclusion-wise larger set of facets.

An equivalent formulation of condition (2) in Theorem 5.3 is that (x)\mathcal{F}(x) contains d+1d+1 hyperplanes of which (a) dd intersect in a single point and (b) the last does not. In particular, if the facet-defining hyperplanes of the allocation space AA are in general position, then a non-constant mechanism x𝒳x\in\mathcal{X} is exhaustive if and only if |(x)|d+1|\mathcal{F}(x)|\geq d+1.191919The hyperplanes in \mathcal{F} are in general position if every subset of more than dd hyperplanes in \mathcal{F} has an empty intersection. If d=2d=2, then the facet-defining hyperplanes are always in general position; thus, a non-constant mechanism x𝒳x\in\mathcal{X} is exhaustive if and only if |(x)|3|\mathcal{F}(x)|\geq 3.

We illustrate the characterization of exhaustiveness and its economic implications with two examples.

Example 5.4.

Let A={a+di=1dai1}A=\{a\in\mathbb{R}^{d}_{+}\mid\sum_{i=1}^{d}a_{i}\leq 1\} be the dd-dimensional unit simplex embedded in d\mathbb{R}^{d}. The unit simplex is the allocation space when considering lotteries over finitely many alternatives or when dividing time or a budget across a finite set of options (see Sections 7 and 8 for applications). By Theorem 5.3, a non-constant mechanism x:ΘAx:\Theta\to A is exhaustive if and only if it makes all d+1d+1 feasibility constraints binding.

A facet of the unit simplex, i.e., feasibility constraint, is characterized by those lotteries in which some alternative is chosen with probability 0. Therefore, in economic terms, a non-constant exhaustive mechanism must allow the agent to avoid any particular alternative with probability 1.

Example 5.5.

Let A=[0,1]dA=[0,1]^{d} be the unit cube in d\mathbb{R}^{d}. The unit cube is the allocation space in a problem with dd goods, one of which could be money. For example, consider a bilateral trade problem where kk goods are owned by the principal, dkd-k goods are owned by the agent, and the principal proposes a menu of possible trades to the agent. By Theorem 5.3, a non-constant mechanism x:ΘAx:\Theta\to A is exhaustive if and only if it makes dd non-parallel feasibility constraints and at least one additional feasibility constraint binding.

A facet of the unit cube, i.e., feasibility constraint, is characterized by those allocations in which some good is either allocated to the principal with probability 1 or to the agent with probability 1. Therefore, in economic terms, an exhaustive mechanism must offer the agent a menu designating at least one good for which the menu contains an option where the agent receives the good with probability 1 and an option where the principal receives the good with probability 1. In addition, for every other good, there must be an option where at least one of the two parties receives the good with probability 1. (The latter condition is automatically satisfied if the menu must include the status quo in which every agent keeps their endowment.)

6. Extreme Points in One- versus Multi-Dimensional Type Spaces

In this section, we show that the extreme points of the set of IC and IR mechanisms have a simple structure in every problem with one-dimensional types but are virtually as rich as the set of exhaustive mechanisms in every problem with multi-dimensional types. Recall that, in our model, a dd-dimensional allocation space AA always corresponds to a (d1)(d-1)-dimensional type space Θ\Theta. Also recall that \mathcal{F} is the set of feasibility constraints defining the allocation space AA.

Theorem 6.1.

Suppose d=2d=2. Then, every extreme point xext𝒳x\in\operatorname{ext}\mathcal{X} is exhaustive and satisfies |menu(x)||||\operatorname{menu}(x)|\leq|\mathcal{F}|.

Proof.

See Section D.3. ∎

Remark.

The bound is tight for the unrestricted type space and attained by allocating to each type one of their most preferred extreme points of the allocation space AA.

Theorem 6.1 is the essential insight of a complete characterization of the extreme points for problems with one-dimensional types (Theorem C.1 in Appendix C): extreme points can be succinctly described as choice functions from a limited number of menu items, akin to the well-known posted-price result for the monopoly problem ([94]; [102]). The complete characterization shows that a mechanism x𝒳x\in\mathcal{X} is an extreme point if and only if menu(x)\operatorname{menu}(x) lacks a certain geometric structure, which we call a flexible chain.

In the multi-dimensional case, the structure of extreme points is fundamentally different and markedly more complex. To make this point, we equip the set of IC and IR mechanisms 𝒳\mathcal{X} with the L1L^{1}-norm

x=Θx(θ)𝑑θ.||x||=\int_{\Theta}||x(\theta)||\,d\theta. (3)

We say that a property holds for most elements of a subset of a topological space if it holds on a dense set that is also a countable intersection of relatively open sets (i.e., a dense GδG_{\delta}); this is a standard notion of topological genericity.

Theorem 6.2.

Suppose d3d\geq 3. Then, every extreme point is exhaustive and most exhaustive mechanisms are extreme points.

Theorems 6.1 and 6.2 together show that properties of binding incentive constraints further discipline the set of exhaustive mechanisms if and only if the type space is one-dimensional. Exhaustiveness is a property of binding feasibility constraints alone. Thus, our results corroborate the heuristic understanding in the mechanism design literature that the difficulty with multi-dimensional screening lies in identifying the incentive constraints that are binding in an optimal mechanism.

6.1. Additional Results

In the remainder of this section, we present additional results for the multi-dimensional case that further strengthen Theorem 6.2. We separately discuss extreme points of finite and infinite menu size as well as uniquely optimal mechanisms. All proofs are in Section D.3.

We first provide a genericity condition under which an exhaustive mechanism of finite menu size is an extreme point. For this, we say that a set of points MAM\subseteq A is in general position if every hyperplane in d\mathbb{R}^{d} intersects MM in at most dd points.

Theorem 6.3.

Suppose d3d\geq 3. If x𝒳x\in\mathcal{X} is exhaustive and menu(x)\operatorname{menu}(x) is finite and in general position, then xext𝒳x\in\operatorname{ext}\mathcal{X}.

That is, every exhaustive mechanism with a finite menu can be transformed into an extreme point by perturbing its menu into general position. By carrying out such perturbations, we obtain the following genericity result:

Theorem 6.4.

Suppose d3d\geq 3. For every kk\in\mathbb{N}, the set of extreme points of menu size kk is relatively open and dense in the set of exhaustive mechanisms of menu size kk.202020An alternative statement is that the set of extreme points of menu size kk is relatively open and dense in the set of exhaustive mechanisms of menu size k\leq k; see the proof.

Thus, extreme points remain prevalent among exhaustive mechanisms even when restricting attention to mechanisms that make only a limited number of allocations.

It is easy to show that mechanisms with a finite menu size are dense in the set of all mechanisms. Consequently, we have:

Corollary 6.5.

Suppose d3d\geq 3. The set of extreme points of finite menu size is dense in the set of exhaustive mechanisms.

We next turn to mechanisms of infinite menu size.

Theorem 6.6.

Suppose d3d\geq 3. Most exhaustive mechanisms are extreme points of uncountable menu size.

Remark.

The proof of Theorem 6.6 establishes the stronger claim that most exhaustive mechanisms are continuous functions (for which the menu is a connected subset of the allocation space). While examples of extreme points with uncountable menu size have been documented in the literature ([81]; [36]), the existence and prevalence of continuous extreme points is novel.

Exhaustive mechanisms can also be approximated by mechanisms that are uniquely optimal for some objective and prior of the principal. That is, even the most parsimonious candidate sets are dense in the set of exhaustive mechanisms. The formal result is a consequence of a theorem due to Straszewicz and Klee ([67]), which asserts that the exposed points of a norm-compact convex set are dense in its extreme points.

Corollary 6.7.

Suppose d3d\geq 3. For every exhaustive mechanism x𝒳x\in\mathcal{X} and every ε>0\varepsilon>0, there exists a mechanism xext𝒳x^{\prime}\in\operatorname{ext}\mathcal{X} such that xx<ε||x-x^{\prime}||<\varepsilon and such that xx^{\prime} is uniquely optimal for some objective function v:Θdv:\Theta\to\mathbb{R}^{d} and belief μΔ(Θ)\mu\in\Delta(\Theta).

In Section 8, we show that the gist of our results continues to hold if we only consider those extreme points that are unique maximizers for specific objectives of the principal such as revenue-maximization. That is, candidate sets remain complex even if the principal’s objective is a priori known and fixed and only their belief is considered a free parameter.

Remark.

We have given an essentially, though not entirely, complete characterization of the extreme points of the set of IC and IR mechanisms. For example, menus that are not in general position and allow some affine dependencies among menu items can still be extreme points. In Appendix B (Theorem B.6), we provide a complete algebraic characterization of finite-menu extreme points. Using the connection to the relevant mathematical concepts to be established in the next section, the reader can consult the references provided in Section 9 for additional conditions. A complete characterization of all extreme points is not to be expected (see Footnote 8 in the introduction).

7. Proof Ideas: The Case of Linear Delegation

In this section, we explain the methodology behind our results. Our approach is to translate between extreme points of the set of (IC) and (IR) mechanisms and extreme points of the set of all menus. Menus can be identified with convex bodies in allocation space, allowing us to draw upon a mathematical literature that has characterized extremal—there called indecomposable—elements of spaces of convex bodies. We illustrate this transfer of results from mathematics to economics through what we consider to be the simplest multi-dimensional screening problem; detailed proofs and generalizations are relegated to Appendices A and D.

7.1. Linear Delegation

We proceed in the context of the linear delegation problem and discuss the necessary adjustments for other problems at the end of this section:

  • A={a+di=1dai1}A=\{a\in\mathbb{R}^{d}_{+}\mid\sum_{i=1}^{d}a_{i}\leq 1\} is the unit simplex, i.e., the allocation space when considering lotteries over m=d+1m=d+1 alternatives or when dividing time or a budget across the alternatives (aa lists the probabilities or shares of the first dd alternatives);

  • Θ=𝕊d1\Theta=\mathbb{S}^{d-1} is the unrestricted type space, i.e., the agent can have all possible von Neumann-Morgenstern preferences over AA;

  • the principal’s objective function v:Θdv:\Theta\to\mathbb{R}^{d} is an arbitrary bounded function, i.e., the principal relies on the agent’s information in order to make an informed decision;

  • there is no veto alternative a¯\underaccent{\bar}{a} for the agent.

The linear delegation problem features multi-dimensional types whenever there are m4m\geq 4 alternatives and thus differs from classical formulations of delegation problems à la [59, 60], which assume one-dimensional allocation and type spaces and single-peaked preferences; see Section 9 for further discussion.

Next to being a natural application of our model, there are two systematic reasons for considering the linear delegation problem:

  1. (1)

    In the linear delegation problem, incentive constraints are completely independent from feasibility constraints in the sense that every mechanism that makes an inclusion-wise maximal set of incentive constraints binding is an extreme point up to positive homothety (Lemma 7.1). This independence simplifies our arguments and renders the connection between extremal menus and indecomposable convex bodies most transparent.

  2. (2)

    Every linear screening problem is linear delegation with a restricted type space (modulo IR constraints). This is because every linear screening problem can be represented with the unit simplex as its allocation space through an appropriate type space restriction.212121Consider a problem with allocation space AA and type space Θ\Theta. Any allocation polytope AdA\subset\mathbb{R}^{d} is the image of a higher-dimensional simplex SnS\subset\mathbb{R}^{n} under a linear map f:ndf:\mathbb{R}^{n}\to\mathbb{R}^{d} [49, Chapter 5.1]. An appropriate type space in n\mathbb{R}^{n} corresponding to the simplex is given by fT(Θ)f^{T}(\Theta), where fTf^{T} is the transpose of ff. With such reformulations, however, coneΘ\operatorname{cone}\Theta is no longer full-dimensional, and because of this additional complexity, we do not use reformulations to linear delegation in our general proofs.

7.2. From Mechanisms to Menus

So far, we have followed the literature in that we have stated our results in terms of direct mechanisms. However, IC mechanisms can equivalently be understood as the agent’s choice functions from different (closed) menus MAM\subseteq A.

We call a closed set MAM\subseteq A an extended menu if every allocation in AMA\setminus M is strictly preferred by at least one type θΘ\theta\in\Theta to every allocation in MM. In other words, if MM is an extended menu, then there is no allocation that can be added to MM without necessarily changing the agent’s choice function. Since the agent has linear utility and we are considering the unrestricted type space in this section, every menu MAM\subseteq A is extended by passing to its convex hull conv(M)\operatorname{conv}(M), which is a convex body in allocation space.

It is straightforward to show that the map which assigns to every mechanism x𝒳x\in\mathcal{X} the extended menu conv(menu(x))A\operatorname{conv}(\operatorname{menu}(x))\subseteq A is a bijection between the space 𝒳\mathcal{X} of (payoff-equivalence classes of) IC mechanisms and the space \mathcal{M} of convex bodies in allocation space.

In Appendix A (Theorem A.2), we show that the bijection between 𝒳\mathcal{X} and \mathcal{M} commutes with convex combinations and, therefore, preserves the linear structure of the underlying spaces. For convex bodies M,MdM,M^{\prime}\subset\mathbb{R}^{d}, this linear structure is given by Minkowski addition and positive scalar multiplication, defined as

λM+ρM={λa+ρaaM,aM},\lambda M+\rho M^{\prime}=\{\lambda a+\rho a^{\prime}\mid a\in M,a^{\prime}\in M^{\prime}\}, (4)

where λ,ρ+\lambda,\rho\in\mathbb{R}_{+}. In particular, extreme points of one space map to extreme points of the other space.

We also show in Appendix A that convergence of extended menus with respect to the Hausdorff distance implies convergence of the corresponding mechanisms in L1L^{1} (Lemma A.7). Thus, any statement about compactness or denseness in the former space carries over to the latter.

7.3. Indecomposability and Exhaustiveness

We next explain how extreme points of the set of extended menus \mathcal{M} can be understood in terms of the notion of indecomposability from the mathematical literature and the notion of exhaustiveness defined in Section 5.

A menu MM\in\mathcal{M} is an extreme point of \mathcal{M} if and only if it does not admit either of the following decompositions:

  1. (1)

    M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} for λ(0,1)\lambda\in(0,1) and M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} homothetic to MM;

  2. (2)

    M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} for λ(0,1)\lambda\in(0,1) and M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} not homothetic to MM.222222The two cases are mutually exclusive: if one of MM^{\prime} or M′′M^{\prime\prime} is homothetic to MM, then so is the other.

We call (1) a homothetic decomposition and (2) a non-homothetic decomposition.

Lemma D.2 in Appendix D shows that a mechanism x𝒳x\in\mathcal{X} is exhaustive if and only if the associated extended menu MM\in\mathcal{M} admits no homothetic decomposition. We can straightforwardly extend the definition of exhaustiveness to extended menus because exhaustiveness is solely a property of the feasibility constraints of the allocation space that are intersected by the menu of a mechanism.

Non-homothetic decompositions are closely related to the notion of decomposability from the mathematical literature. A convex body KdK\subset\mathbb{R}^{d} is decomposable if there exist convex bodies K,K′′dK^{\prime},K^{\prime\prime}\subset\mathbb{R}^{d} not homothetic to MM such that M=K+K′′M=K^{\prime}+K^{\prime\prime}. By scaling the summands, decomposability is equivalent to the existence of convex bodies K,K′′dK^{\prime},K^{\prime\prime}\subset\mathbb{R}^{d} not homothetic to KK such that K=λK+(1λ)K′′K=\lambda K^{\prime}+(1-\lambda)K^{\prime\prime} with λ(0,1)\lambda\in(0,1). A convex body that is not decomposable is indecomposable.

If an extended menu MM\in\mathcal{M} is indecomposable, then MM has no non-homothetic decomposition. The converse does not generally hold because the summands λM\lambda M^{\prime} and (1λ)M′′(1-\lambda)M^{\prime\prime} of a non-homothetic decomposition in our model are required to be subsets of AA, i.e., feasible extended menus.232323In the absence of feasibility constraints, every convex body trivially has homothetic decompositions, e.g. through translations into opposite directions. This is why homothetic decompositions are ruled out in the definition of indecomposability. However, when the allocation space is a simplex, indecomposability is necessary and sufficient for the absence of non-homothetic decompositions.

Lemma 7.1.

In the linear delegation problem, an extended menu MM\in\mathcal{M} is in ext\operatorname{ext}\mathcal{M} if and only if MM is indecomposable and exhaustive.

Proof.

See Section D.4. ∎

Before proceeding with a characterization of the indecomposable convex bodies, we briefly discuss the economic meaning of indecomposability. Recall that an extreme point of finite menu size is determined by its binding incentive and feasibility constraints (Theorem 4.1). Indecomposability of the associated extended menu ensures that (up to payoff-equivalence) there is no other, non-constant mechanism that makes an inclusion-wise larger set of incentive constraints binding; exhaustiveness ensures the same for the feasibility constraints. Thus, by Lemma 7.1 and in the linear delegation problem, the role of incentive and feasibility constraints in whether or not a mechanism is an extreme point can be completely separated. Indeed, in other linear screening problems, extreme points need not make inclusion-wise maximal sets of incentive constraints binding. (Nevertheless, it is helpful to analyze feasibility constraints separately from the incentive constraints, as we have done in Section 5.)

7.4. Characterizing Extreme Points

Given Lemma 7.1, it remains to characterize indecomposable and exhaustive extended menus. Indecomposability has been characterized in the mathematical literature.

Theorem ([117, 91]).

A convex body M2M\subset\mathbb{R}^{2} is indecomposable if and only if it is a point, line segment, or triangle.

Figure 1 depicts the proof idea for convex polygons. The figure shows a quadrilateral and two deformations of the quadrilateral that translate the right-most, vertical facet-defining line either to the left or to the right. The resulting deformed quadrilaterals yield a non-homothetic decomposition of the original quadrilateral. Similar deformations can be found for any polygon, but triangles are the only polygons for which these deformations yield homotheties of the triangle. Thus, (degenerate) triangles are the only indecomposable convex polygons. The extension to all plane convex bodies requires a more involved argument.

Theorem ([116]).

Let d3d\geq 3. The set of indecomposable convex bodies in d\mathbb{R}^{d} is Hausdorff-dense in the set of all convex bodies in d\mathbb{R}^{d}.

Shephard identifies a large class of indecomposable polytopes, with the simplest being the simplicial polytopes, i.e., polytopes of which every proper face is a simplex. Roughly speaking, a simplicial polytope SS is indecomposable because each two-dimensional face of SS is a triangle and any decomposition of SS into non-homothetic polytopes would also have to decompose every face of SS individually, which is impossible because triangles are indecomposable. Simplicial polytopes are Hausdorff-dense in the space of all convex bodies. First, every convex body is arbitrarily close to a polytope. (Take the convex hull of a finite set of points on the body’s boundary that is ε\varepsilon-dense in the boundary.) Second, every polytope can be transformed into a simplicial polytope by perturbing its vertices into general position; Figure 4 illustrates.

v1v_{1}v2v_{2}v3v_{3}v4v_{4}wwv1v_{1}v2v_{2}v3v_{3}v4v_{4}^{\prime}ww
Figure 4. Illustration of how to perturb a polytope into a nearby simplicial polytope. Left: a pyramid with apex ww and base v1v_{1}-v2v_{2}-v3v_{3}-v4v_{4}. Right: a simplicial polytope obtained from the pyramid by pulling the vertex v4v_{4} to a new vertex v4v_{4}^{\prime} such that the five vertices are in general position. This procedure can be iteratively applied to the vertices of any polytope to obtain a nearby simplicial polytope. (Incidentally, a pyramid is already indecomposable.)

Exhaustiveness admits a simple economic characterization in the linear delegation problem, which follows immediately from Theorem 5.3 (recall Example 5.4). We state the characterization in terms of mechanisms, but it can equivalently be stated in terms of the associated extended menus:

  • A constant mechanism xx is exhaustive if and only if it dictates an alternative: there exists aextAa\in\operatorname{ext}A such that menu(x)={a}\operatorname{menu}(x)=\{a\}.

  • A non-constant mechanism xXx\in X is exhaustive if and only if it grants a strike: for every alternative k=1,,mk=1,\ldots,m there exists a lottery amenu(x)a\in\operatorname{menu}(x) in which alternative kk is chosen with probability 0. That is, the agent is given the option to strike out any one of the alternatives. Geometrically speaking, this means that menu(x)\operatorname{menu}(x) touches all facets of the allocation simplex.

The following characterization result follows at once from the previous arguments and the bijection between the set of mechanisms 𝒳\mathcal{X} and the set of extended menus \mathcal{M}.

Theorem 7.2.

Consider the linear delegation problem:

  1. (1)

    With m=3m=3 alternatives, x𝒳x\in\mathcal{X} is in ext𝒳\operatorname{ext}\mathcal{X} if and only if one of the following holds:

    1. (a)

      xx dictates an alternative;

    2. (b)

      xx grants a strike and has menu size at most three.

  2. (2)

    With m4m\geq 4 alternatives, ext𝒳\operatorname{ext}\mathcal{X} is dense in the set of mechanisms that grant a strike.

Thus, the theory predicts simple solutions for linear delegation problems with three alternatives, but with four or more alternatives and up to approximation, the only distinguishing property of extreme points is that they dictate or grant a strike.

We remark that optimality in the linear delegation problem, even when there are only three alternatives, may require the use of stochastic mechanisms that offer the agent lotteries over the alternatives.242424See [75] and [70] for a discussion about the optimality of stochastic mechanisms in the classical one-dimensional delegation model. Lotteries can be interpreted as risky courses of action or as budget or time shares. Optimality may even require lotteries with full support, i.e., interior points of the simplex. Intuitively, lotteries give the principal more leeway in screening the agent and make it more difficult for the agent to align the allocation with their own preferences.

7.5. General Linear Screening Problems

We finally discuss the necessary adjustments to our approach when considering (IR) constraints, allocation spaces different from the simplex, and restricted type spaces.

In the context of the linear delegation problem, IR would mean that the menu of a mechanism must contain the veto alternative a¯extA\underaccent{\bar}{a}\in\operatorname{ext}A. Any decomposition of a given convex body that contains a¯\underaccent{\bar}{a} must also contain a¯\underaccent{\bar}{a}. Thus, introducing IR constraints simply amounts to considering extreme points of the set of IC mechanisms that also satisfy IR. The same conclusion obtains in other linear screening problems; see Section A.3.

Suppose the allocation space AA differs from the simplex. If an extended menu MM\in\mathcal{M} is indecomposable, then it does not admit a non-homothetic decomposition. However, the converse is no longer true. This is inconsequential for the denseness results for multi-dimensional problems since we only get additional, extremal but decomposable extended menus. For one-dimensional type spaces, these additional extreme points drive the bound on the menu size from three up to the number of feasibility constraints of the allocation space (Theorem 6.1). We provide a complete characterization of extremal extended menus for one-dimensional type spaces and arbitrary allocation spaces AA (Theorem C.1 in Appendix C). This characterization builds on a mathematical result due to [93].

Suppose the type space is restricted, i.e. coneΘd\operatorname{cone}\Theta\neq\mathbb{R}^{d}. Extending a menu now entails more than taking the convex hull because there are certain directions in the allocation space along which all types are worse off. Geometrically speaking, these directions form the polar cone of the type space. To prove our result, it is a technical convenience to extend menus beyond the boundaries of the allocation space and work with closed convex sets that share the polar cone as a common recession cone. Indecomposability for closed convex sets with a common recession cone is analogous to indecomposability for convex bodies and has been discussed in [118].

8. Specific Objectives: Multi-Good Monopoly and Linear Veto Bargaining

Our previous analysis considered candidates for optimality that the principal must a priori consider when uncertain about both their objective function and the distribution of the agent’s types; we now fix the principal’s objective, e.g., revenue maximization, and characterize the mechanisms that remain relevant for optimality as the type distribution varies.

In applications to the multi-good monopoly problem and the linear veto bargaining problem, to be defined below, we show that the set of mechanisms that are uniquely optimal for some type distribution is dense in the set of undominated mechanisms. A mechanism is undominated if there is no other mechanism that yields the principal an unambiguously higher utility. We provide characterizations of undominated mechanisms, showing that they are almost as rich as the set of all (IC) and (IR) mechanisms. Thus, the gist of our main results holds when restricting attention to extreme points that are unique maximizers for specific objectives of the principal. We discuss the two applications after introducing undominated mechanisms.

8.1. Undominated Mechanisms

For multi-dimensional problems, we have identified exhaustive mechanisms as a reference set in which the extreme points lie dense. However, with a fixed objective vv, not every extreme point remains relevant for optimality. For example, an extreme point might minimize expected revenue for some type distribution μ\mu. The appropriate reference set now becomes the set of undominated mechanisms, originally defined for the multi-good monopoly problem by [81].

Definition 8.1.

A mechanism x𝒳x\in\mathcal{X} is dominated by another mechanism x𝒳x^{\prime}\in\mathcal{X} if x(θ)v(θ)x(θ)v(θ)x^{\prime}(\theta)\cdot v(\theta)\geq x(\theta)\cdot v(\theta) for almost all θΘ\theta\in\Theta, with strict inequality on a set of types of positive measure. A mechanism x𝒳x\in\mathcal{X} is undominated if it is not dominated by any other mechanism x𝒳x^{\prime}\in\mathcal{X}.

[81] show for the monopoly problem that every undominated mechanism is optimal for some belief about the agent’s type. Their benchmark result can be extended from revenue maximization to arbitrary objectives:

Theorem 8.2.

For every undominated mechanism x𝒳x\in\mathcal{X}, there exists a type distribution μΔ(Θ)\mu\in\Delta(\Theta) such that xx is an optimal mechanism for a principal with belief μ\mu.

Proof.

See Section D.5. ∎

Conversely, every mechanism that is optimal for some fully supported type distribution μΔ(Θ)\mu\in\Delta(\Theta) must clearly be undominated.

A priori, not every undominated mechanism is a necessary candidate for optimality. (Undominated mechanisms need not be extreme or exposed points). However, in the following applications and as long as types are multi-dimensional, we show that every undominated mechanism is arbitrarily close to a mechanism that is uniquely optimal for some type distribution, i.e., arbitrarily close to a mechanism that is a necessary candidate for optimality.

8.2. Multi-Good Monopoly

The multi-good monopoly problem is the following linear screening problem:

  • A=[0,1]m×[0,κ]A=[0,1]^{m}\times[0,\kappa], where the first mm allocation dimensions are the probabilities with which good i=1,,mi=1,\ldots,m is sold to the agent, and the last allocation dimension is the payment by the agent (and κ\kappa is some sufficiently large constant, which is without loss of generality whenever valuations are bounded);

  • Θ=[0,1]m×{1}\Theta=[0,1]^{m}\times\{-1\}, i.e., the consumer has valuations in [0,1][0,1] for each good i=1,,mi=1,\ldots,m and money is the numeraire;252525Due to linear utility, we implicitly assume that the goods are neither substitutes nor complements for the agent. This assumption is made in most papers on the multi-good monopoly problem. We could incorporate substitutes and complements by allowing the agent to have one valuation for each possible bundle B{1,,m}B\subseteq{\{1,\ldots,m\}}. The allocation space is then the unit simplex over 2m2^{m} deterministic allocations, i.e., all possible bundles, plus an extra dimension representing money as before. Free disposal, i.e., the agent being willing to pay weakly more for inclusion-wise larger bundles, and a fixed marginal utility for money can be modeled as a family of affine restrictions on the type space.

  • a¯=(0,,0,0)\underaccent{\bar}{a}=(0,\ldots,0,0), i.e., the consumer can leave without paying anything;

  • v(θ)=v¯={0,,0,1}v(\theta)=\bar{v}=\{0,\ldots,0,1\} for all θΘ\theta\in\Theta, i.e., the principal maximizes expected revenue (and goods can be produced at zero cost).262626The literature makes the zero-cost assumption for simplicity. It can easily be relaxed to a constant marginal cost for each good. With decreasing marginal costs, extreme points also remain the relevant candidates for optimality (see the discussion in [81]). With increasing marginal costs, one has to follow the approach taken by [104].

In line with standard terminology in mechanism design with transfers, we abuse our language by referring to a[0,1]ma\in[0,1]^{m} as an allocation and t[0,κ]t\in[0,\kappa] as the transfer. Instead of probabilities, allocations can also be interpreted as quantities or as quality-differentiated goods with multiple attributes (for which the consumer has unit demand).

We next show that a large class of mechanisms in the monopoly problem is undominated. A pricing function is a continuous convex function p:[0,1]m+p:[0,1]^{m}\to\mathbb{R}_{+} such that p(0)=0p(0)=0 that assigns a price to each possible allocation.272727Convexity and continuity are without loss of generality because the agent has linear utility. p(0)=0p(0)=0 reflects the (IR) constraint. See also [56, Appendix A.2 ]. The marginal price for good i=1,,di=1,\ldots,d at allocation a[0,1]ma\in[0,1]^{m} with ai<1a_{i}<1 is the directional derivative eip(a)\nabla_{e_{i}}p(a) of pp at aa in the coordinate direction eie_{i} (which exists by the convexity and continuity of pp). The mechanism x𝒳x\in\mathcal{X} obtained from a pricing function is the agent’s choice function from the menu M={(a,p(a))a[0,1]m}M=\{(a,p(a))\mid a\in[0,1]^{m}\}.

Lemma 8.3.

In the multi-good monopoly problem, every mechanism x𝒳x\in\mathcal{X} can be obtained from a pricing function pp with marginal prices in [0,1][0,1]. If a mechanism x𝒳x\in\mathcal{X} can be obtained from a pricing function pp with marginal prices eip\nabla_{e_{i}}p uniformly bounded away from 0 and 1 for every good i=1,,di=1,\ldots,d, then it is undominated.

Proof.

See Section D.5. ∎

In plain words, a mechanism that, on the margin, prevents low-valuation types from buying additional quantity while enabling high-valuation types to buy additional quantity is undominated. Such a mechanism features “no-distortion at the top” (the highest type receives the efficient allocation) and “exclusion at the bottom” (the lowest type receives nothing), which are well-known properties of optimal mechanisms in screening problems with transfers. In particular, such a mechanism features these two properties separately in each allocation dimension. Not all undominated mechanisms have marginal prices bounded away from zero and one, but the gap to the mechanisms that do admit this bound is negligible.282828For an example, see the mechanism depicted in Figure 2 in [81]. In the bottom-right “market segment,” the marginal price for good one is 1.

Corollary 8.4.

In the multi-good monopoly problem, the set of undominated mechanisms is dense in the set of all (IC) and (IR) mechanisms.

Proof.

See Section D.5. ∎

For a rough intuition for the richness of undominated mechanisms, consider the following trade-off. When the principal increases the price for some allocations, revenue increases from those types who continue to choose these allocations. However, some types that have previously chosen an allocation at the lower price may now opt for a cheaper allocation, decreasing revenue from the types that switch. This trade-off rules out a dominance relationship between many mechanisms.

Given the characterization of undominated mechanisms, the same arguments as in Section 7 can be applied to conclude that extreme points are dense in the set of undominated mechanisms and, therefore, in the set of all mechanisms by Corollary 8.4. In the following result, the first part is well-known (see, for example, [81, Lemma 4 ]).

Theorem 8.5.

Consider the multi-good monopoly problem:

  1. (1)

    With m=1m=1 good, a mechanism x𝒳x\in\mathcal{X} is in ext𝒳\operatorname{ext}\mathcal{X} and undominated if and only if xx is a posted-price mechanism with price p(0,1)p\in(0,1), i.e.,

    x(θ)={(1,p)if θ1p(0,0)otherwise.x(\theta)=\begin{cases}(1,p)&\text{if }\theta_{1}\geq p\\ (0,0)&\text{otherwise.}\end{cases}
  2. (2)

    With m2m\geq 2 goods, the set of mechanisms x𝒳x\in\mathcal{X} that are uniquely optimal for some belief μΔ(Θ)\mu\in\Delta(\Theta) is dense in 𝒳\mathcal{X}.

Proof.

See Section D.5. ∎

Remark.

The proof shows that statement (2) remains true if the belief μ\mu is required to have full support on Θ\Theta.

The second part says that any incentive-compatible and individually rational mechanism can be turned into a mechanism that is uniquely optimal for some belief of the seller by applying an arbitrarily small perturbation. The claim about uniquely optimal mechanisms is not an application of Straszewicz’ theorem upon showing denseness of the extreme points in the set of undominated mechanisms. While Straszewicz’ theorem guarantees that exposed points are arbitrarily close to extreme points, these points may be exposed by linear functionals unrelated to revenue maximization. Our proof modifies the theorem to obtain the desired result.

8.3. Linear Veto Bargaining

We now discuss the following linear veto bargaining problem:

  • A={a+di=1dai1}A=\{a\in\mathbb{R}^{d}_{+}\mid\sum_{i=1}^{d}a_{i}\leq 1\} is the unit simplex, i.e., the allocation space when considering lotteries over m=d+1m=d+1 alternatives or when dividing time or a budget across the alternatives (aa lists the probabilities or shares of the first dd alternatives);

  • Θ=𝕊d1\Theta=\mathbb{S}^{d-1} is the unrestricted domain, i.e., the agent can have all possible von Neumann-Morgenstern preferences over AA;

  • there is a veto alternative a¯extA\underaccent{\bar}{a}\in\operatorname{ext}A for the agent (e.g., the status quo in a political context), and we set a¯=(0,,0)\underaccent{\bar}{a}=(0,\ldots,0) without loss of generality;

  • the principal’s preferences are given by a Bernoulli utility vector v¯\bar{v} independently of the agent’s information, i.e., v(θ)=v¯v(\theta)=\bar{v} for all θΘ\theta\in\Theta, and we assume for simplicity that (1) v¯++\bar{v}\in\mathbb{R}_{++}, i.e., the veto alternative is the principal’s least preferred alternative, and (2) argmaxiv¯i\operatorname*{arg\,max}_{i}\bar{v}_{i} is a singleton, i.e., the principal has a unique favorite alternative.

The problem can be seen as a delegation problem with a state-independent objective and an IR constraint. Therefore, the extreme points of linear veto bargaining are exactly the extreme points of linear delegation that satisfy IR (Lemma A.9). Linear veto bargaining can also be seen as a no-transfers analogue of the monopoly problem since both problems feature state-independent objectives with an IR constraint.

Lemma 8.6.

In the linear veto bargaining problem, a mechanism x𝒳x\in\mathcal{X} is undominated if and only if menu(x)\operatorname{menu}(x) contains the veto alternative and the principal’s most preferred alternative.

Proof.

See Section D.5. ∎

The richness of undominated mechanisms in the veto bargaining problem comes from a trade-off similar to that in the monopoly problem. By adding an alternative to the menu of a mechanism, some types prefer the new alternative over their previous choice. Among those who switch, some types will do so in the principal’s favor, i.e., switch away from alternatives that the principal likes less than the new alternative. Other types will not switch in the principal’s favor, i.e., switch away from alternatives that the principal likes more than the new alternative. A similar trade-off arises when removing an alternative from the menu. These trade-offs prevent a dominance relationship between mechanisms that allocate the principal’s most preferred alternative.

As before, given the characterization of undominated mechanism above, the same arguments as in Section 7 can be applied to conclude that the extreme points are dense in the set of undominated mechanisms whenever there are four or more alternatives. The claim about uniquely optimal mechanisms again requires additional arguments.

Theorem 8.7.

Consider the linear veto bargaining problem:

  1. (1)

    With m=3m=3 alternatives, a mechanism x𝒳x\in\mathcal{X} is undominated and in ext𝒳\operatorname{ext}\mathcal{X} if and only if menu(x)\operatorname{menu}(x) contains the veto alternative, the principal’s most preferred alternative, and at most one other lottery over the alternatives.

  2. (2)

    With m4m\geq 4 alternatives, the set of mechanisms x𝒳x\in\mathcal{X} that are uniquely optimal for some belief μΔ(Θ)\mu\in\Delta(\Theta) is dense in the set of undominated mechanisms.

Proof.

See Section D.5. ∎

9. Related Literature

This paper relates to several areas of research, including multi-dimensional screening, extreme points in mechanism design, delegation and veto bargaining, and the mathematical literature on indecomposability. We will explain the relation to these four areas after first discussing [81], whose work most closely relates to ours.

9.1. [81] (MV)

In the context of the multi-good monopoly problem, MV provide the first—and, prior to this paper, only—analysis of extreme points in multi-dimensional mechanism design, with two main contributions. First, they provide an algebraic characterization of finite-menu extreme points in terms of whether or not a certain linear system associated with a given mechanism has a unique solution. This characterization is based on auxiliary results about the facial structure of the set of incentive-compatible mechanisms. Second, they define the notion of undominated mechanisms and show that every undominated mechanism maximizes expected revenue for some distribution of types.

In comparison to MV, we consider arbitrary linear screening problems with or without transfers and subsume the multi-good monopoly problem as a special case. We contribute explicit, non-algebraic extreme-point characterizations (Section 6). These characterizations reveal the precise structure of the set of extreme points and, therefore, the structure of the possible solutions to linear screening problems. Along the way, we obtain generalizations of MVs results in our more general framework; see Appendix B and Theorem 8.2.

In comparison to MV, we also characterize undominated extreme points and uniquely optimal mechanisms. While MV show for the monopoly problem that all undominated mechanisms are potentially optimal, it has not been known which undominated mechanisms are necessary candidates for optimality, i.e., which extreme points are undominated and uniquely optimal for some type distribution. A priori, one might conjecture that parsimonious candidate sets are significantly smaller than the set of undominated mechanisms. We show that this is not the case: the relevant exposed points are dense in the set of undominated mechanisms. Moreover, we provide new results about undominated mechanisms, showing that these mechanisms are themselves virtually as rich as the set of all IC and IR mechanisms.

Finally, we note that MV have shown for the monopoly problem, modulo minor details, that the extreme points of menu size kk\in\mathbb{N} are relatively open and dense in the IC mechanisms of menu size kk, provided kk is smaller than the number of goods for sale plus one (their Remark 25). We show that this substantial qualifier on kk is not necessary and that the result holds for arbitrary multi-dimensional screening problems with linear utility (Theorem 6.4).

9.2. Multi-dimensional Screening and Mechanism Design

The literature on multi-dimensional screening—and on the multi-good monopoly problem in particular—is much too large to be summarized here in detail. We focus on recent developments and point to a survey by [105] for work up to the early 2000s.292929A sample of important early work includes [1], [114], [87], [122], [8], and [104].

Recent work focuses mostly on the multi-good monopoly problem and can be classified into several approaches for gaining insights into multi-dimensional screening problems or for circumventing the severe difficulties associated with their classical formulations:

  • provide conditions for the optimality of common mechanisms such as separate sales or bundling ([87]; [80]; [41]; [100]; [36]; [90]; [19]; [53]; [47]; [124]);

  • provide duality results that can be used to certify the optimality of a given mechanism ([36]; [69]; [27]; [73]; [68]);

  • identify specific structural properties of optimal mechanisms (e.g., subadditive pricing, monotonicity, no randomization) and show when such structure arises ([87]; [80]; [56]; [13]; [17]; [21]);

  • quantify the worst-case performance (approximation ratio) of common mechanisms or classes of mechanisms ([54]; [55]; [77]; [11]; [109]; [57]; [12]; [17]);

  • identify mechanisms with the optimal worst-case performance for a mechanism designer with Knightian uncertainty over the set of type distributions ([28]; [37]; [33]);

  • derive asymptotic optimality results for a large number of i.i.d. goods ([9]; [14]) or for the speed of convergence to first-best as the principal gains increasingly precise information about the agent’s type ([44]).

Our paper is orthogonal to these developments. We do not focus on specific properties and classes of mechanisms or attempt to escape intractabilities. Instead, we shed light on where these intractabilities originate and identify the limits of the qualitative predictions that can be drawn within the standard Bayesian framework. Moreover, to the best of our knowledge, multi-dimensional screening without transfers has not been studied, with the exception of [68], whose duality approach to a multi-dimensional delegation problem is complementary to our extreme points approach.

Besides the implications for optimal mechanism design, we contribute to the literature on implementability with multi-dimensional type spaces (e.g., [103]; [110]; [20]) by characterizing extreme points of the set of incentive-compatible mechanisms. By Choquet’s theorem, every non-extreme point can be represented as a mixture over extreme points.

In general, little is known about optimal multi-dimensional mechanism design with multiple agents ([99]; [62]; [30]; [61]; [73]); see the conclusion for further discussion.

9.3. Extreme Points in Mechanism Design

A number of papers have approached mechanism design problems by studying the extreme points of the set of incentive-compatible mechanisms. However, aside from the previously discussed work by [81], this approach has only been applied to one-dimensional problems. For instance, [23] uses extreme points—hierarchical allocations—in a characterization of the set of feasible interim allocation rules. Building on Border’s insights, [82] demonstrate the equivalence of Bayesian and dominant strategy incentive-compatibility in standard auction problems. A similar approach is discussed in [121, Chapter 6 ].

[70] present characterizations of the extreme points of certain majorization sets and show how these majorization sets naturally arise as feasible sets in many economic design problems. In the context of mechanism design, their results immediately imply a characterization of the extreme points of the set of feasible and incentive-compatible interim allocation rules in one-dimensional symmetric allocation problems, providing a new perspective on Border’s theorem as well as BIC-DIC equivalence. Their approach is tailored to one-dimensional problems, elegantly handling both the IC constraints (monotonicity for one-dimensional types) and the Maskin-Riley-Matthews-Border feasibility constraints (majorization with respect to the efficient allocation rule).

In subsequent work, [71] characterize certain extreme points of the set of measures defined on a compact convex subset of d\mathbb{R}^{d} that are dominated in the convex order by a given measure. Their result is a multi-dimensional analogue of results obtained in [70] about the set of monotone functions that majorize a given monotone function (see also [7]). These results apply to information design but have no obvious applications to mechanism design.

[96, 97] builds on the majorization approach, allowing for additional constraints on the majorization sets. These constraints may, for example, correspond to fairness or efficiency constraints in a revenue-maximization problem. [125] provide a complementary analysis to [70] based on characterizations of extreme points of sets of distributions characterized by first-order stochastic dominance conditions rather than second-order stochastic dominance conditions (majorization).

Extreme point approaches have also been used in mechanism design without transfers (e.g., [18]; [95]), and several other mechanism design papers use extreme points as a technical tool (e.g., [34]).

9.4. Delegation and Veto Bargaining

Much of the literature on optimal delegation has focused on one-dimensional allocation (action) and type spaces with single-peaked preferences; see [59, 60], [89], [84], [4], [5], [74], and [70].

The applications of our results to delegation differ from the classical literature in two ways. First, allocations in our delegation problem are lotteries over finitely many alternatives.303030Delegation over a finite set of alternatives is also studied in the project selection literature; see [10], [98], [31], and [52]. Second, both the principal and agent have arbitrary vNM preferences over these alternatives; that is, our problem features an unrestricted rather than single-peaked preference domain and therefore multi-dimensional types (and allocations). We can allow more general allocation spaces, provided the agent’s utility remains linear.

A small number of papers consider multi-dimensional type or allocation spaces. [72] study optimal delegation in a setting with a one-dimensional type space and two allocation dimensions across which the principal and the agent have separable quadratic preferences. [42] links multiple independent, one-dimensional delegation problems. Frankel shows that “halfspace delegation,” i.e., imposing a quota on the weighted average of actions across problems, is optimal for normally distributed states and approximately optimal for general distributions as the number of linked problems goes to infinity. See also [43] for a robust mechanism design approach. [68] studies optimal delegation with both multi-dimensional type and allocation spaces. Kleiner’s duality-based approach is complementary to our extreme-point approach.

Veto bargaining is a classical problem in political science, originally studied in [108]. The case with incomplete information about the agent’s (vetoer’s) preferences has only recently been studied using a mechanism design approach by [64].313131See also [2]. Their model features one-dimensional private information. [6] and [113] study related one-dimensional delegation problems with IR constraints where the principal does not necessarily have state-independent preferences.323232See also the ”balanced” delegation problem in [74]. Similarly, our model can nest linear delegation problems with IR constraints.

9.5. Mathematical Foundations

[46] introduced the notion of an indecomposable convex body and announced the first results about indecomposability. Gale’s results were later proven and published in [116], [91]/[117], and [112]. These and other papers have provided many novel results that go beyond Gale’s original presentation. [88, 92, 118] provide algebraic characterizations of indecomposable polytopes. [118] discusses indecomposable polyhedra. Related results characterize extremal convex bodies within a given compact convex set in the plane ([93, 50]); see Theorems 6.1 and C.1 for the application in our paper. Decomposability is related to deformations of polytopes, which we briefly use in Appendix B; [29, Section 2 ] provide a concise treatment. Textbook references on indecomposability include [115, Chapter 3.2 ], [101, Chapter 6 ], and [49, Chapter 15 ].

Characterizations of indecomposable convex bodies can alternatively be seen, via support function duality, as characterizations of the extremal rays of the cone of sublinear (i.e. convex and homogeneous) functions. A subset of the results known in the literature on indecomposable convex bodies have been independently obtained in studies of the extremal rays of the cone of convex functions by [63] (for two-dimensional domains) and [26] (for dd-dimensional domains).333333We thank Andreas Kleiner for pointing us to these references.

We finally mention a result due to [65, Proposition 2.1, Theorem 2.2 ], which shows that for most (in the sense of topological genericity) compact convex subsets of an infinite-dimensional Banach space, the extreme points of the set are dense in the set itself. This follows since such sets have an empty interior, support points are dense in the boundary, hence in the set itself, and since most such sets are strictly convex, so that every support point is an extreme point. However, the set of IC mechanisms is a specific compact convex subset of an infinite-dimensional Banach space, which, in particular, is not strictly convex. The content of our results is that whenever the type space is multi-dimensional, the extreme points are nevertheless dense in a certain part of the set.

10. Conclusion

We have characterized extreme points of the set of incentive-compatible (IC) mechanisms for screening problems with linear utility. For every problem with one-dimensional types, extreme points admit a simple characterization with a tight upper bound on their menu size. In contrast, for every problem with multi-dimensional types, we have identified a large set of IC mechanisms—exhaustive mechanisms—in which the extreme and exposed points lie dense. Consequently, one-dimensional problems allow us to make predictions that are independent of the precise details of the environment, whereas such predictions are largely unattainable for multi-dimensional problems.

One might hope that restricting attention to specific instances of a given multi-dimensional screening problem allows more robust predictions regarding optimality. We have shown that such predictions remain elusive in applications to monopoly and veto bargaining problems, where the principal’s objective is fixed and state-independent and only the principal’s belief about the agent’s type is considered a free parameter.

While our focus has been on screening problems, where there is only a single (representative) agent, one should expect implications of our results for multi-agent settings. In multi-agent settings, Bayesian incentive compatibility of a given multi-agent mechanism is the same as separately requiring incentive compatibility with respect to each agent’s interim-expected mechanism (see, e.g., [24, Chapter 6]). These interim-expected mechanisms, one for each agent, must then be linked towards an ex-post feasible mechanism via an appropriate analogue of the Maskin-Riley-Matthews-Border conditions.343434[85], [86], [23]. Recent treatments include [32], [48], and [120]; see these papers for further references and discussions of potential limitations of the reduced-form approach. Thus, if the extreme points in a multi-agent problem were simpler than the extreme points characterized here for the one-agent case, then this reduction in complexity would have to come from these additional conditions. This is not the case for problems with one-dimensional types (see, e.g., [70]) and is not to be expected for problems with multi-dimensional types.

Our main methodological contribution is to link extreme points of the set of incentive-compatible mechanisms to indecomposable convex bodies studied in convex geometry. This methodology, where we study incentive-compatible mechanisms by analyzing the space of all menus from which the agent could choose, is potentially useful in other areas of economic theory. Examples that come to mind are menu choice à la [38] and the random expected utility (REU) model of [51].

Appendix A Preliminaries & Auxiliary Results

This appendix gathers general tools we use throughout the proofs of our results from the main text. Section A.1 shows that there are bijections between mechanisms, menus, and indirect utility functions that commute with convex combinations (in the sense of Minkowski). The commutativity with convex combinations is essential for our subsequent analysis because we will study extremal menus and then translate back to extremal mechanisms, as explained in Section 7. Section A.2 introduces the relevant topological structure for the three sets of objects. Section A.3 discusses how individual rationality (IR) constraints are incorporated into our analysis.

A.1. Mechanisms, Menus, and Indirect Utility Functions

Recall that we have identified payoff-equivalent mechanisms and that 𝒳\mathcal{X} is the set of payoff-equivalence classes of (IC) and (IR) mechanisms.

Let

𝒰={U:θx(θ)θx𝒳}\mathcal{U}=\{U:\theta\mapsto x(\theta)\cdot\theta\mid x\in\mathcal{X}\}

denote the set of all indirect utility functions induced by the mechanisms in 𝒳\mathcal{X}. It is a direct consequence of (IC) that an indirect utility function is HD1 (homogeneous of degree 1) on coneΘ\operatorname{cone}\Theta because types on the same ray from the origin have the same ordinal preferences. Thus, we extend indirect utility functions U𝒰U\in\mathcal{U} to d\mathbb{R}^{d} by setting U(λθ)=λU(θ)U(\lambda\theta)=\lambda U(\theta) for all θΘ\theta\in\Theta and λ0\lambda\geq 0 and U(z)=U(z)=\infty for all zconeΘz\notin\operatorname{cone}\Theta.

A menu is simply a subset MAM\subset A that the principal offers the agent and from which the agent chooses their favorite allocation. However, different menus can induce payoff-equivalent choice functions, i.e., payoff-equivalent IC mechanisms, for the agent. Thus, we define the notion of an extended menu, which is the inclusion-wise largest representative of a payoff-equivalence class of menus.

To define extended menus, let

Θ={ydθΘ,yθ0}\Theta^{\circ}=\{y\in\mathbb{R}^{d}\mid\forall\theta\in\Theta,\,y\cdot\theta\leq 0\} (5)

denote the polar cone of Θ\Theta. The polar cone of type space is the set of all directions in allocation space AA along which no type’s utility ever strictly improves. If coneΘ=d\operatorname{cone}\Theta=\mathbb{R}^{d}, then the only such direction is the trivial direction 0. By definition, we may add to every menu MAM\subset A the polar cone Θ\Theta^{\circ} and instead offer the agent the Minkowski sum M+ΘM+\Theta^{\circ} without affecting the agent’s indirect utility. We may also take the closed convex hull of MM, which does not affect indirect utility either since utility is linear. By requiring a¯M\underaccent{\bar}{a}\in M, i.e., the veto allocation is in MM, we ensure that the agent does not veto the menu.

Definition A.1.

An extended menu is a closed convex set MdM\subset\mathbb{R}^{d} such that M=convextM+ΘM=\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}, extMA\operatorname{ext}M\subset A, and a¯M\underaccent{\bar}{a}\in M. The set of all extended menus is denoted by \mathcal{M}.

If the type space is unrestricted, i.e., coneΘ=d\operatorname{cone}\Theta=\mathbb{R}^{d}, then extended menus are convex bodies in AA; otherwise, they are unbounded closed convex sets. That extended menus offer infeasible allocations when the type space is restricted has no physical meaning and is merely a convenient way to identify payoff-equivalent menus. We note that extended menus are uniquely pinned down by their extreme points.

v1v_{1}v2v_{2}v3v_{3}v4v_{4}v5v_{5}Θ\Theta^{\circ}coneΘ\operatorname{cone}\ThetaAAMM
Figure 5. An example of an extended menu for a restricted type space. The type cone, coneΘ\operatorname{cone}\Theta, is the 4545^{\circ} cone, shaded in dark gray. The polar cone Θ\Theta^{\circ} is the 135135^{\circ} cone, also shaded in dark gray and with extremal rays that are orthogonal to the extremal rays of coneΘ\operatorname{cone}\Theta. The allocation space AA is the square. An exemplary menu {v1,v2,v3,v4,v5}\{v_{1},v_{2},v_{3},v_{4},v_{5}\} is depicted using dots. Its extension is the polyhedron MM shaded in light gray and with boundary given by the dotted lines. MM is obtained by taking the convex hull of {v1,v2,v3,v4,v5}\{v_{1},v_{2},v_{3},v_{4},v_{5}\} and adding the polar cone Θ\Theta^{\circ}. Here, v2v_{2} and v3v_{3} “vanish” in the polar cone because v2v_{2} and v3v_{3} are dominated by the other three allocations for every type in Θ\Theta.

Figure 5 illustrates the construction of extended menus. The depicted allocation and type spaces fit a one-good monopoly problem. The horizontal allocation dimension a1a_{1} is the probability of sale, and the vertical allocation dimension a2a_{2} is the payment. The type space is Θ=[0,1]×{1}\Theta=[0,1]\times\{-1\}, where the first component is the agent’s valuation for the good. The extremal rays of the polar cone Θ\Theta^{\circ} are allocation directions in which (1) the agent gets the good with lower probability for the same payment, and (2) the agent gets the good with higher probability but for a marginal price that makes the type (1,1)(1,-1), who is willing to pay most, just indifferent. The extended menu MM corresponds to a mechanism where some types never transact (v1v_{1}), some types buy a cheap lottery that sometimes allocates the good (v5v_{5}), and all other types buy the good with probability 1 at a more expensive price (v4v_{4}).

As in Section 7, we equip \mathcal{M} with the operations of Minkowski addition and positive scalar multiplication.

Theorem A.2.

The following functions are bijections that commute with convex combinations:

  • Φ1:𝒳\Phi_{1}:\mathcal{X}\to\mathcal{M} where x(convmenu(x)+Θ)x\mapsto(\operatorname{conv}\operatorname{menu}(x)+\Theta^{\circ});

  • Φ2:𝒰\Phi_{2}:\mathcal{M}\to\mathcal{U} where M(θsupyMyθ)M\to(\theta\mapsto\sup_{y\in M}y\cdot\theta);

  • Φ3:𝒰𝒳\Phi_{3}:\mathcal{U}\to\mathcal{X} where UθΘ(U(θ)A)U\mapsto\prod_{\theta\in\Theta}(\partial U(\theta)\cap A).353535U(θ)\partial U(\theta) denotes the subdifferential of UU at θΘ\theta\in\Theta. The proof and Corollary A.4 below confirm that the subdifferential of an indirect utility function U𝒰U\in\mathcal{U} is well-defined because UU is convex.

That is, Φ1\Phi_{1} maps (payoff-equivalence classes of) IC and IR mechanisms to the extension of their menus; Φ2\Phi_{2} maps extended menus to their support functions; Φ3\Phi_{3} maps indirect utility functions to their subdifferential.

Proof.

Define an auxiliary map Φ1:x(clconvx(Θ)+Θ)\Phi_{1}^{\prime}:x\mapsto(\operatorname{\operatorname{cl}\operatorname{conv}}x(\Theta)+\Theta^{\circ}), where x:ΘAx:\Theta\to A is an IC and IR mechanism, and claim that Φ1(x)=M\Phi_{1}^{\prime}(x)=M\in\mathcal{M}. We have a¯M\underaccent{\bar}{a}\in M for otherwise there would exist a type that strictly prefers a¯\underaccent{\bar}{a} to x(Θ)x(\Theta) by the linearity of utility and the definitions of the closed convex hull and the polar, which contradicts that xx satisfies (IR). We also have

extM=ext(clconvx(Θ)+Θ)ext(clconvx(Θ))A,\operatorname{ext}M=\operatorname{ext}(\operatorname{\operatorname{cl}\operatorname{conv}}x(\Theta)+\Theta^{\circ})\subseteq\operatorname{ext}(\operatorname{\operatorname{cl}\operatorname{conv}}x(\Theta))\subset A,

where the first inclusion follows since Θ\Theta^{\circ} is a cone and the second inclusion follows since x(Θ)Ax(\Theta)\subseteq A and AA is compact and convex. Finally, MM is closed and convex because it is a sum of a compact convex set and a closed convex set.

The map Φ2:𝒰\Phi_{2}:\mathcal{M}\to\mathcal{U} is well-defined: for each MM\in\mathcal{M}, its support function supaMaθ\sup_{a\in M}a\cdot\theta is an indirect utility function in 𝒰\mathcal{U} because argmaxaMaθ\operatorname*{arg\,max}_{a\in M}a\cdot\theta is non-empty for all θΘ\theta\in\Theta and every selection from the argmax is an IC and IR mechanism.

We next show that Φ2Φ1\Phi_{2}\circ\Phi_{1}^{\prime} is the map that assigns to each IC and IR mechanism xx its indirect utility function U𝒰U\in\mathcal{U}. Let U𝒰U\in\mathcal{U} be the indirect utility function associated with x𝒳x\in\mathcal{X}. As desired, we have

U(θ)={supax(Θ)aθ=supyclconvx(Θ)+Θyθ=supyMyθ if θconeΘ=supyMyθ otherwise. U(\theta)=\begin{cases}\sup_{a\in x(\Theta)}a\cdot\theta=\sup_{y\in\operatorname{\operatorname{cl}\operatorname{conv}}x(\Theta)+\Theta^{\circ}}y\cdot\theta=\sup_{y\in M}y\cdot\theta&\text{ if }\theta\in\operatorname{cone}\Theta\\ \infty=\sup_{y\in M}y\cdot\theta&\text{ otherwise. }\end{cases}

In the first case, the first equality is (IC) and the second equality follows from the definitions of the closed convex hull and the polar cone. The second case also follows by definition of the polar cone.

The map Φ2:𝒰\Phi_{2}:\mathcal{M}\to\mathcal{U} is injective because support functions uniquely determine closed convex sets ([58, Theorem V.2.2.2]).

Thus, if xx and xx^{\prime} are payoff-equivalent IC and IR mechanisms, then Φ1(x)=Φ1(x)\Phi_{1}^{\prime}(x)=\Phi_{1}^{\prime}(x^{\prime}) because Φ2\Phi_{2} is injective and (Φ2Φ1)(x)=(Φ2Φ1)(x)(\Phi_{2}\circ\Phi_{1}^{\prime})(x)=(\Phi_{2}\circ\Phi_{1}^{\prime})(x^{\prime}) (by the definition of payoff-equivalence). Thus, Φ1\Phi_{1}^{\prime} can be defined on the set of payoff-equivalence classes 𝒳\mathcal{X} in the obvious way.

Φ1:𝒳\Phi_{1}^{\prime}:\mathcal{X}\to\mathcal{M} and Φ2:𝒰\Phi_{2}:\mathcal{M}\to\mathcal{U} are bijective because Φ2Φ1\Phi_{2}\circ\Phi_{1}^{\prime} is bijective and Φ2\Phi_{2} is injective.

We next show that Φ1=Φ1\Phi_{1}=\Phi_{1}^{\prime}. Let x=(Φ1)1(M)𝒳x=(\Phi_{1}^{\prime})^{-1}(M)\in\mathcal{X} be any representative mechanism from the payoff-equivalence class associated with MM\in\mathcal{M}. Then, expMmenu(x)\exp M\subseteq\operatorname{menu}(x) since every mechanism xx^{\prime} that is payoff-equivalent to xx must necessarily allocate to each type θΘ\theta\in\Theta with a uniquely preferred option aexpMa\in\exp M that option. Moreover, menu(x)clextM\operatorname{menu}(x)\subseteq\operatorname{cl}\operatorname{ext}M because every type can find a favorite allocation in extM\operatorname{ext}M. By Theorem 2.3 in [65], extMclexpMmenu(x)\operatorname{ext}M\subseteq\operatorname{cl}\exp M\subseteq\operatorname{menu}(x) since menu(x)\operatorname{menu}(x) is compact. Thus, menu(x)=clextM\operatorname{menu}(x)=\operatorname{cl}\operatorname{ext}M. Consequently, M=convmenu(x)+ΘM=\operatorname{conv}\operatorname{menu}(x)+\Theta^{\circ} since MM is closed, as desired.

We next verify that the inverse of the composition Φ2Φ1:𝒳𝒰\Phi_{2}\circ\Phi_{1}:\mathcal{X}\to\mathcal{U} is given by Φ3:𝒰𝒳\Phi_{3}:\mathcal{U}\to\mathcal{X}. Take any (IC) and (IR) mechanism xx with associated extended menu MM\in\mathcal{M} and associated indirect utility function U𝒰U\in\mathcal{U}. For all θΘ\theta\in\Theta, we have

x(θ)argmaxax(Θ)aθargmaxaMaθ=U(θ),x(\theta)\in\operatorname*{arg\,max}_{a\in x(\Theta)}a\cdot\theta\subseteq\operatorname*{arg\,max}_{a\in M}a\cdot\theta=\partial U(\theta),

where the first step is (IC), the second step is immediate from the definition of MM, and the third step is a property of support functions ([106, Corollary 8.2.5]).

It remains to show commutativity with convex combinations. That Φ2:𝒰\Phi_{2}:\mathcal{M}\to\mathcal{U} commutes with convex combinations is a property of support functions [58, Theorem V.3.3.3].363636Remark on the cited theorem: in general, the sum of two closed convex sets need not be closed, but it is always closed if the two sets have the same recession cone, which is here Θ\Theta^{\circ}. That Φ3:𝒰𝒳\Phi_{3}:\mathcal{U}\to\mathcal{X} commutes with convex combinations follows from the linearity of the gradient map, which is almost everywhere well-defined. Thus, Φ1:𝒳\Phi_{1}:\mathcal{X}\to\mathcal{M} must also commute with convex combinations. ∎

Theorem A.2 is fundamental to our approach because the bijections between 𝒳\mathcal{X}, 𝒰\mathcal{U}, and \mathcal{M} map extreme points to extreme points.373737Extreme points are usually only defined for convex subsets of vector spaces, which \mathcal{M} is not. However, we can embed \mathcal{M} into a vector space by Theorem A.2, which justifies the use of the term “extreme point.” We prove our main results by investigating the extreme points of \mathcal{M}. Occasionally, however, we shall work with mechanisms or indirect utility functions, if this simplifies our arguments.

We say that (x,M,U)(x,M,U) are associated if they are isomorphic in the sense of Theorem A.2.

Given Theorem A.2, the definitions of (positive) homothety and exhaustiveness straightforwardly extend from 𝒳\mathcal{X} to \mathcal{M} and 𝒰\mathcal{U}. For example, if x𝒳x\in\mathcal{X} with associated MM\in\mathcal{M}, then

(M):=(x)={HHextM}.\mathcal{F}(M):=\mathcal{F}(x)=\{H\in\mathcal{F}\mid H\cap\operatorname{ext}M\neq\emptyset\}. (6)

We note a few corollaries of Theorem A.2.

Corollary A.3.

Let x𝒳x\in\mathcal{X} and MM\in\mathcal{M} be associated. Then, menu(x)=clextM\operatorname{menu}(x)=\operatorname{cl}\operatorname{ext}M.

We have proven this claim as part of the proof of Theorem A.2. Note that extM\operatorname{ext}M is closed if extM\operatorname{ext}M is finite or d=2d=2.

The following characterization of indirect utility functions is analogous to the one by [103, Proposition 2 ] for settings with transfers.

Corollary A.4.

U𝒰U\in\mathcal{U} if and only if the following conditions are satisfied:

  1. (1)

    UU is sublinear (i.e., convex and HD1).

  2. (2)

    UU is continuous on its effective domain coneΘ={zn:U(z)<}\operatorname{cone}\Theta=\{z\in\mathbb{R}^{n}:\>U(z)<\infty\}.

  3. (3)

    For all θconeΘ\theta\in\operatorname{cone}\Theta, U(θ)a¯θU(\theta)\geq\underaccent{\bar}{a}\cdot\theta.

  4. (4)

    For all θconeΘ\theta\in\operatorname{cone}\Theta, extU(θ)A\operatorname{ext}\partial U(\theta)\subset A.

Proof.

By the previous result, U𝒰U\in\mathcal{U} is the support function of an extended menu MM\in\mathcal{M}. Conversely, every closed sublinear function d{}\mathbb{R}^{d}\to\mathbb{R}\cup\{\infty\} is the support function of a closed convex set. (A sublinear function that is continuous on a closed effective domain is closed.)

It remains to show that the remaining properties hold if and only if MM\in\mathcal{M}. The effective domain of a sublinear function (in our case: coneΘ\operatorname{cone}\Theta) and the recession cone of the associated closed convex set (in our case: Θ\Theta^{\circ}) are mutually polar cones ([58, Proposition V.2.2.4]). Continuity comes for free since coneΘ\operatorname{cone}\Theta is polyhedral ([107, Theorem 10.2]). It is easy to see that (3) holds if and only if a¯M\underaccent{\bar}{a}\in M ([58, Proposition V.2.2.4]). Finally, extMA\operatorname{ext}M\subset A if and only if extU(θ)A\operatorname{ext}\partial U(\theta)\subset A for all θΘ\theta\in\Theta follows from Corollary 8.2.5 in [106]. ∎

We also note the following sanity check that almost everywhere equivalence indeed coincides with payoff-equivalence for (IC) and (IR) mechanisms. This justifies modeling the set 𝒳\mathcal{X} of payoff-equivalence classes of mechanisms in L1L^{1}.

Corollary A.5.

Let xx and xx^{\prime} be mechanisms that satisfy (IC) and (IR). Then, xx and xx^{\prime} are payoff-equivalent if and only if x=xx=x^{\prime} almost everywhere.

Proof.

If xx and xx^{\prime} are payoff-equivalent, then there exists an indirect utility function U𝒰U\in\mathcal{U} such that x,xUx,x^{\prime}\in\partial U by Theorem A.2. Thus, x=xx=x^{\prime} almost everywhere since the subdifferential of a convex function is almost everywhere a singleton. Conversely, suppose x=xx=x^{\prime} almost everywhere. Let xUx\in\partial U and xUx^{\prime}\in\partial U^{\prime}. Then, U=U\nabla U=\nabla U^{\prime} almost everywhere. Thus, U=U+cU=U^{\prime}+c for cc\in\mathbb{R}, and c=0c=0 because UU and UU^{\prime} are sublinear. Thus, xx and xx^{\prime} are payoff-equivalent. ∎

Remark.
v4v_{4}v1v_{1}v2v_{2}v3v_{3}Θ\Theta^{\circ}coneΘ\operatorname{cone}\ThetaAAMM
Figure 6. The extended menu M=conv{v2,v3}+ΘM=\operatorname{conv}\{v_{2},v_{3}\}+\Theta^{\circ} of a mechanism xx that can be decomposed pointwise almost everywhere, i.e., up to payoff-equivalence, but not pointwise everywhere. Define xx as follows: assign to each type in the interior of coneΘ\operatorname{cone}\Theta their favorite allocation between v2v_{2} and v3v_{3} and two the types on the extremal rays of coneΘ\operatorname{cone}\Theta the allocations v1v_{1} and v4v_{4}. MM can be decomposed by translating the vertex v3v_{3} horizontally; thus xx can be decomposed up to payoff-equivalence, i.e., pointwise almost everywhere. However, xx cannot be decomposed pointwise everywhere.

We briefly comment on a subtle difference between extreme points of the set of payoff-equivalence classes of IC and IR mechanisms, i.e., ext𝒳\operatorname{ext}\mathcal{X}, versus extreme points of the set of IC and IR mechanisms themselves. For the former, a mechanism is an extreme point if it does not coincide with a convex combination of two other mechanisms up to payoff-equivalence, i.e., for almost every type. For the latter, a mechanism is an extreme point if it does not coincide with a convex combination of two other mechanisms for every type. For an extremal equivalence class with associated indirect utility function Uext𝒰U\in\operatorname{ext}\mathcal{U}, every element of θΘext(U(θ)A)\prod_{\theta\in\Theta}\operatorname{ext}(\partial U(\theta)\cap A) is an extreme point of the set of (IC) and (IR) mechanisms.

There can exist additional extreme points of the set of (IC) and (IR) mechanisms such that their payoff-equivalence classes are not extreme points of the set of payoff-equivalence classes 𝒳\mathcal{X}. These additional extreme points can only exist if the type space Θ\Theta is restricted and only if types on the boundary of coneΘ\operatorname{cone}\Theta break ties to the boundary of AA; see Figure 6 for an example. Since we assume that the prior distribution μ\mu is absolutely continuous, these additional extreme points are irrelevant for optimality.

A.2. Topologies and Compactness

We now define topologies on the three sets, 𝒳\mathcal{X}, \mathcal{M}, and 𝒰\mathcal{U}, discuss the relation between these topologies, and show that the three sets are compact under their respective topologies.

We equip the set 𝒳\mathcal{X} of payoff-equivalence classes of mechanisms with the L1L_{1}-norm

x=Θx(θ)𝑑θ.||x||=\int_{\Theta}||x(\theta)||\,d\theta. (7)

We equip the set 𝒰\mathcal{U} of indirect utility functions with the sup-norm

U=supθconeΘ:θ1U(θ).||U||=\sup_{\theta\in\operatorname{cone}\Theta:\>||\theta||\leq 1}U(\theta). (8)

We equip the set \mathcal{M} of extended menus with the Hausdorff distance

d(M,M)=inf{ε>0:MM+εB and MM+εB},d(M,M^{\prime})=\inf\left\{\varepsilon>0:\>M\subseteq M^{\prime}+\varepsilon B\text{ and }M^{\prime}\subseteq M+\varepsilon B\right\}, (9)

where B={zd:z1}B=\{z\in\mathbb{R}^{d}:\>||z||\leq 1\} is the unit ball in d\mathbb{R}^{d}.

Thus, (𝒳,||||)(\mathcal{X},||\cdot||) and (𝒰,||||)(\mathcal{U},||\cdot||) are normed spaces.383838For (𝒳,||||)(\mathcal{X},||\cdot||), recall that payoff-equivalent mechanisms are almost everywhere equal by Corollary A.5. We also have:

Lemma A.6.

(,d)(\mathcal{M},d) is a metric space and d(M,M)d(convextM,convextM)d(M,M^{\prime})\leq d(\operatorname{conv}\operatorname{ext}M,\operatorname{conv}\operatorname{ext}M^{\prime}).

Proof.

We have

d(M,M)\displaystyle d(M,M^{\prime}) =inf{ε>0:MM+εB and MM+εB}\displaystyle=\inf\left\{\varepsilon>0:\>M\subseteq M^{\prime}+\varepsilon B\text{ and }M^{\prime}\subseteq M+\varepsilon B\right\}
=inf{ε>0:convextM+ΘconvextM+Θ+εB and convextM+ΘconvextM+Θ+εB}\displaystyle=\inf\left\{\varepsilon>0:\>\begin{multlined}\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}\subseteq\operatorname{conv}\operatorname{ext}M^{\prime}+\Theta^{\circ}+\varepsilon B\text{ and }\\ \operatorname{conv}\operatorname{ext}M^{\prime}+\Theta^{\circ}\subseteq\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}+\varepsilon B\end{multlined}\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}\subseteq\operatorname{conv}\operatorname{ext}M^{\prime}+\Theta^{\circ}+\varepsilon B\text{ and }\\ \operatorname{conv}\operatorname{ext}M^{\prime}+\Theta^{\circ}\subseteq\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}+\varepsilon B\right\}
inf{ε>0:convextMconvextM+εB and convextMconvextM+εB}\displaystyle\leq\inf\left\{\varepsilon>0:\>\begin{multlined}\operatorname{conv}\operatorname{ext}M\subseteq\operatorname{conv}\operatorname{ext}M^{\prime}+\varepsilon B\text{ and }\\ \operatorname{conv}\operatorname{ext}M^{\prime}\subseteq\operatorname{conv}\operatorname{ext}M+\varepsilon B\end{multlined}\operatorname{conv}\operatorname{ext}M\subseteq\operatorname{conv}\operatorname{ext}M^{\prime}+\varepsilon B\text{ and }\\ \operatorname{conv}\operatorname{ext}M^{\prime}\subseteq\operatorname{conv}\operatorname{ext}M+\varepsilon B\right\}
=d(convextM,convextM),\displaystyle=d(\operatorname{conv}\operatorname{ext}M,\operatorname{conv}\operatorname{ext}M^{\prime}),

where the inequality is because Z1Z2Z_{1}\subseteq Z_{2} implies Z1+Z3Z2+Z3Z_{1}+Z_{3}\subseteq Z_{2}+Z_{3} for Z1,Z2,Z3dZ_{1},Z_{2},Z_{3}\subset\mathbb{R}^{d}.

It remains to show that (,d)(\mathcal{M},d) is a metric space. Since convextM,convextMA\operatorname{conv}\operatorname{ext}M,\operatorname{conv}\operatorname{ext}M^{\prime}\subseteq A, we have d(M,M)d(convextM,convextM)<d(M,M^{\prime})\leq d(\operatorname{conv}\operatorname{ext}M,\operatorname{conv}\operatorname{ext}M^{\prime})<\infty. Thus, dd is a metric on \mathcal{M} since extended menus are closed and since the Hausdorff distance is an extended metric on the space of all closed subsets of n\mathbb{R}^{n}. ∎

The topologies on 𝒰\mathcal{U} and \mathcal{M} are equivalent and finer than the topology on 𝒳\mathcal{X}.

Lemma A.7.

Consider sequences (xn)n𝒳(x_{n})_{n\in\mathbb{N}}\subset\mathcal{X}, and (Mn)n(M_{n})_{n\in\mathbb{N}}\subset\mathcal{M}, (Un)n𝒰(U_{n})_{n\in\mathbb{N}}\subset\mathcal{U} such that (xn,Mn,Un)(x_{n},M_{n},U_{n}) are associated for all nn\in\mathbb{N}. Then, the following hold:

  1. (1)

    MnMM_{n}\to M if and only if UnUU_{n}\to U.

  2. (2)

    If UnUU_{n}\to U, then xnxx_{n}\to x.

Proof.

Claim (1) is Theorem 6 in [111]. For claim (2), let DnconeΘD_{n}\subset\operatorname{cone}\Theta be the set of points where UnU_{n} is differentiable and let DconeΘD\subset\operatorname{cone}\Theta be the set of points where UU is differentiable. Let D=DnDnD^{*}=D\cap\bigcap_{n\in\mathbb{N}}D_{n}. Indirect utility functions are convex and therefore almost everywhere differentiable. Moreover, the countable union of nullsets is null; thus coneΘD\operatorname{cone}\Theta\setminus D^{*} is null. Theorem VI.6.2.7 in [58] implies that Un(θ)U(θ)\nabla U_{n}(\theta)\to\nabla U(\theta) for all θconeΘD\theta\in\operatorname{cone}\Theta\setminus D^{*}. Moreover, by Theorem A.2, xnxx_{n}\to x pointwise almost everywhere. The Dominated Convergence Theorem implies convergence in L1L^{1}. ∎

The following lemma is crucial to apply Bauer’s maximum theorem, Choquet’s theorem, and the Straszewicz-Klee theorem, and hence for the interpretation of our results about extreme points.

Lemma A.8.

𝒳\mathcal{X}, \mathcal{M}, and 𝒰\mathcal{U} are compact and convex.

Proof.

Convexity of 𝒳\mathcal{X} is immediate because (IC) and (IR) are linear constraints and because AA is convex. By Theorem A.2, \mathcal{M} and 𝒰\mathcal{U} are also convex.

For compactness, consider any sequence {Mn}n\{M_{n}\}_{n\in\mathbb{N}}\subset\mathcal{M}. By Blaschke’s selection theorem, {clconvextMn}n\{\operatorname{\operatorname{cl}\operatorname{conv}}\operatorname{ext}M_{n}\}_{n\in\mathbb{N}} has a convergent subsequence {clconvextMnk}k\{\operatorname{\operatorname{cl}\operatorname{conv}}\operatorname{ext}M_{n_{k}}\}_{k\in\mathbb{N}} with compact convex limit KAK\subseteq A. Let M=K+ΘM=K+\Theta^{\circ}. It is readily verified that MM\in\mathcal{M}. By Lemma A.6, the subsequence {Mnk}k\{M_{n_{k}}\}_{k\in\mathbb{N}} convergences to MM\in\mathcal{M}. Thus, \mathcal{M} is compact. By Lemma A.7, 𝒳\mathcal{X} and 𝒰\mathcal{U} are also compact. ∎

A.3. Individual Rationality (IR)

The following result is an analogue of the familiar observation in mechanism design with transfers that if IR holds for “the lowest type,” then IR holds for every type. In our setting, however, a “lowest type” need not exist and is instead a type θ¯Θ\underaccent{\bar}{\theta}\in\Theta who likes the veto allocation most, i.e., a¯argmaxaAaθ¯\underaccent{\bar}{a}\in\operatorname*{arg\,max}_{a\in A}a\cdot\underaccent{\bar}{\theta}.

Lemma A.9.

Suppose x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} almost everywhere, where x,x,x′′x,x^{\prime},x^{\prime\prime} are (IC) mechanisms and λ(0,1)\lambda\in(0,1). If xx satisfies (IR), then xx^{\prime} and x′′x^{\prime\prime} satisfy (IR).

Thus, the extreme points of the set of (IC) and (IR) mechanisms are simply the extreme points of the set of (IC) mechanisms that satisfy (IR).

Proof.

Let

Θ={θΘa¯argmaxaAaθ}\Theta^{*}=\{\theta\in\Theta\mid\underaccent{\bar}{a}\in\operatorname*{arg\,max}_{a\in A}a\cdot\theta\}

be the set of types who like the veto allocation most. We have assumed in Section 3 that Θ\Theta^{*} is non-empty.

Let

f=θΘargmaxaAaθ.f^{*}=\bigcap_{\theta\in\Theta^{*}}\operatorname*{arg\,max}_{a\in A}a\cdot\theta.

Since AA is a polytope, ff^{*} is a face of AA. If xx is an (IC) and (IR) mechanism, then x(Θ)fx(\Theta)\cap f^{*}\neq\emptyset.

Since ff^{*} is a face of AA, if x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} for (IC) mechanisms x,x′′x^{\prime},x^{\prime\prime} and λ(0,1)\lambda\in(0,1), then x(Θ)fx^{\prime}(\Theta)\cap f^{*}\neq\emptyset and x′′(Θ)fx^{\prime\prime}(\Theta)\cap f^{*}\neq\emptyset.

For the sake of contradiction, suppose xx^{\prime} does not satisfy (IR). Then, a¯θ>x(θ)θ\underaccent{\bar}{a}\cdot\theta>x^{\prime}(\theta)\cdot\theta for some θΘ\theta\in\Theta. By (IC), x(θ)θaθx^{\prime}(\theta)\cdot\theta\geq a^{*}\cdot\theta for ax(Θ)fa^{*}\in x^{\prime}(\Theta)\cap f^{*}. Thus, a¯θ>aθ\underaccent{\bar}{a}\cdot\theta>a^{*}\cdot\theta, which contradicts the definition of ff^{*}. Thus, xx^{\prime} satisfies (IR). Analogously, x′′x^{\prime\prime} satisfies (IR). ∎

Appendix B Extreme Points of Finite-Menu Mechanisms and Deformations

This appendix characterizes for any given IC mechanism with finite menu size the set of all IC mechanisms that make an inclusion-wise larger set of IC constraints binding. This set is important in our analysis: whenever an IC mechanism can be written as a convex combination of two other IC mechanisms, then these two mechanisms must make at least the same incentive constraints binding as the given mechanism.

We use this characterization to prove Theorem 4.1 and Theorem C.1. Moreover, we can use the characterization to generalize results by [81, Theorems 17, 19, 20, and 24 ] (MV) to arbitrary linear screening problems. In particular, we get an algebraic characterization of finite-menu extreme points (Theorem B.6). We discuss the exact relation to MV at the end of this section.

Throughout this section, we restrict attention to extended menus MM\in\mathcal{M} of finite size, i.e., |extM|<|\operatorname{ext}M|<\infty. Let Fin\mathcal{M}^{\text{Fin}}\subset\mathcal{M} denote the set of all extended menus of finite size. These are polyhedra since they can be written as the convex hull of their extreme points plus the polar of the type space (which is a polyhedral cone). In light of Lemma A.9, we can ignore IR constraints. We make two closely connected definitions.

Definition B.1.

The normal fan 𝒩M\mathcal{N}_{M} of an extended menu MFinM\in\mathcal{M}^{\text{Fin}} is the collection {NCf}\{NC_{f}\} of the normal cones

NCf={θconeΘfargmaxaMaθ}NC_{f}=\{\theta\in\operatorname{cone}\Theta\mid f\subseteq\operatorname*{arg\,max}_{a\in M}a\cdot\theta\}

to the faces ff of MM. The normal fan 𝒩M\mathcal{N}_{M^{\prime}} is coarser than the normal fan 𝒩M\mathcal{N}_{M}, denoted 𝒩M𝒩M\mathcal{N}_{M^{\prime}}\preccurlyeq\mathcal{N}_{M}, if each normal cone in 𝒩M\mathcal{N}_{M^{\prime}} is a union of some set of normal cones in 𝒩M\mathcal{N}_{M}.

Since the agent has linear utility, the set of each type’s most preferred alternatives is a face of MM. The normal fan hence summarizes which types’ most preferred alternatives lie on which faces of the extended menu. The normal fan yields a polyhedral subdivision of the type space; the cells of maximal dimension have been called market segments by MV in the context of the monopoly problem.393939Subdivisions obtained from normal fans of polyhedra have appeared elsewhere in economic design as power diagrams ([45]; [71]) and as regular polyhedral complexes ([15]; [119]; [16]). They are also relevant in the context of the random expected utility model ([51]). See [40] for another recent application.

For the next definition, we define the set of facet-defining hyperplanes of an extended menu MFinM\in\mathcal{M}^{\text{Fin}}, which requires some care when MM is not dd-dimensional. For each facet FiF_{i} of MM, there is a unique outer normal vector ni(affMa)n_{i}\in(\operatorname{aff}M-a), where aMa\in M is arbitrary, and a constant cic_{i}\in\mathbb{R} such that Fi=MHiF_{i}=M\cap H_{i} and MHi,M\subseteq H_{i,-}, where Hi={zd:zni=c}H_{i}=\{z\in\mathbb{R}^{d}:z\cdot n_{i}=c\} is the facet-defining hyperplane and Hi,={zd:znici}H_{i,-}=\{z\in\mathbb{R}^{d}:z\cdot n_{i}\leq c_{i}\} is the facet-defining halfspace. Let M\mathcal{H}_{M} be the union of the set of facet-defining hyperplanes of MM with an arbitrary finite set of hyperplanes with corresponding halfspaces whose intersection is affM\operatorname{aff}M. For brevity, we refer to M\mathcal{H}_{M} as the set of facet-defining hyperplanes of MM (although some of these define the improper face MM).

Definition B.2.

An extended menu MM^{\prime}\in\mathcal{M} is a deformation of MFinM\in\mathcal{M}^{\text{Fin}} with M={H1,,Hk}\mathcal{H}_{M}=\{H_{1},\ldots,H_{k}\} if there exist a deformation vector c=(c1,,ck)kc^{\prime}=(c_{1}^{\prime},\dots,c_{k}^{\prime})\in\mathbb{R}^{k} such that the following two conditions are satisfied:

  1. (1)

    M=i=1kHi,M^{\prime}=\cap_{i=1}^{k}H^{\prime}_{i,-}, where Hi,={zd:znici}H^{\prime}_{i,-}=\{z\in\mathbb{R}^{d}:z\cdot n_{i}\leq c^{\prime}_{i}\}.

  2. (2)

    If iIHi={a}\cap_{i\in I}H_{i}=\{a\} for I{1,,k}I\subseteq\{1,\dots,k\} and aextMa\in\operatorname{ext}M, then there exists aextMa^{\prime}\in\operatorname{ext}M^{\prime} such that iIHi={a}\cap_{i\in I}H^{\prime}_{i}=\{a^{\prime}\}.

Let Def(M)\operatorname{Def}(M)\subset\mathcal{M} denote the set of deformations of MM.

That is, (1) MM^{\prime} can be defined by translates of the facet-defining halfspaces of MM, not all of which necessarily remain facet-defining, and (2) if some subset of the facet-defining hyperplanes of MM defines a vertex of MM, then the translated hyperplanes also define a vertex of MM^{\prime}. See Figure 1 for an illustration, where the right-most facet-defining hyperplane of the menu is translated horizontally, yielding two deformations. (The left panel of Figure 7 in Section D.3 is another illustration). This definition of deformations is due to [29, Definition 2.2 ], except here adapted to polyhedra rather than polytopes.

Remark.

There is a bijection between deformations MDef(M)M^{\prime}\in\operatorname{Def}(M) and deformation vectors cc^{\prime} given by

ci=maxaMnia,i=1,,kc_{i}^{\prime}=\max_{a\in M^{\prime}}n_{i}\cdot a,\quad\forall i=1,\ldots,k

since every hyperplane HiH_{i}^{\prime} in the definition of MM^{\prime} must support MM^{\prime} by condition (2). By condition (1), this bijection commutes with convex combinations.

Lemma B.3.

Let x,x𝒳x,x^{\prime}\in\mathcal{X} be finite menu mechanisms with associated extended menus M,MFinM,M^{\prime}\in\mathcal{M}^{\text{Fin}}. The following are equivalent:

  1. (1)

    𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}).

  2. (2)

    𝒩M𝒩M\mathcal{N}_{M^{\prime}}\preccurlyeq\mathcal{N}_{M}.

  3. (3)

    MM^{\prime} is a deformation of MM.

  4. (4)

    There exists a surjective map φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime} such that for every edge404040One-dimensional face. ab¯=conv{a,b}\overline{ab}=\operatorname{conv}\{a,b\}. ab¯\overline{ab} of MM there exists λab+\lambda_{ab}\in\mathbb{R}_{+} such that λab(ab)=φ(a)φ(b)\lambda_{ab}(a-b)=\varphi(a)-\varphi(b).

The lemma says that coarsening the normal fan is the geometric analogue of making inclusion-wise more incentive constraints binding. Deformations are exactly the operations on extended menus that coarsen the normal fan. The fourth condition is an equivalent formulation of deformations in terms of parallel edges and more readily reveals the algebraic nature of deformations.

Proof.

For the proof, we will need the following basic observation about normal cones. For an extended menu MFinM\in\mathcal{M}^{\text{Fin}} with M={H1,,Hk}\mathcal{H}_{M}=\{H_{1},\ldots,H_{k}\} and a face ff of MM, let If={1ikfHi}I_{f}=\{1\leq i\leq k\mid f\subseteq H_{i}\} denote the set of facet-defining hyperplanes of MM containing the face ff. For every face ff of MM, we have:

NCf=cone{ni}iIf.NC_{f}=\operatorname{cone}\{n_{i}\}_{i\in I_{f}}. (10)

In particular, dimNCf=ddimf\operatorname{dim}NC_{f}=d-\operatorname{dim}f.

We define

NCθ\displaystyle NC_{\theta} :=NCargmaxaMaθ\displaystyle:=NC_{\operatorname*{arg\,max}_{a\in M}a\cdot\theta}
NCθ\displaystyle NC^{\prime}_{\theta} :=NCargmaxaMaθ.\displaystyle:=NC_{\operatorname*{arg\,max}_{a\in M^{\prime}}a\cdot\theta}.

NCθNC_{\theta} and NCθNC^{\prime}_{\theta} are the inclusion-wise smallest normal cones of MM and MM^{\prime}, respectively, to which θ\theta belongs. By definition, 𝒩M={NCθ}θΘ\mathcal{N}_{M}=\{NC_{\theta}\}_{\theta\in\Theta} and 𝒩M={NCθ}θΘ\mathcal{N}_{M^{\prime}}=\{NC_{\theta}^{\prime}\}_{\theta\in\Theta}.

We also make the following preliminary observation: for all θ,θ~intconeΘ\theta,\tilde{\theta}\in\operatorname{int}\operatorname{cone}\Theta,

(θ,θ~)𝒞(x)argmaxaMaθargmaxaMaθNCθNCθ~(\theta,\tilde{\theta})\in\mathcal{IC}(x)\iff\operatorname*{arg\,max}_{a\in M}a\cdot\theta\supseteq\operatorname*{arg\,max}_{a\in M^{\prime}}a\cdot\theta\iff NC_{\theta}\subseteq NC_{\tilde{\theta}} (11)

because menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M and menu(x)=extM\operatorname{menu}(x^{\prime})=\operatorname{ext}M^{\prime} (Corollary A.3), a bounded face of polyhedron is the convex hull of some set of its extreme points, every type θintconeΘ\theta\in\operatorname{int}\operatorname{cone}\Theta is normal to a bounded face of MM and MM^{\prime}, and normal cones are dual to faces and, therefore, reverse the inclusion.

(1) \implies (2). By (10), if 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}) and θ,θ~intconeΘ\theta,\tilde{\theta}\in\operatorname{int}\operatorname{cone}\Theta, then NCθNCθ~NC_{\theta}\subseteq NC_{\tilde{\theta}} implies NCθNCθ~NC^{\prime}_{\theta}\subseteq NC^{\prime}_{\tilde{\theta}}. In particular, θNCθ~\theta\in NC_{\tilde{\theta}} implies θNCθ~\theta\in NC^{\prime}_{\tilde{\theta}}. Thus, every cone in 𝒩M\mathcal{N}_{M} that meets intconeΘ\operatorname{int}\operatorname{cone}\Theta is a subset of a cone in 𝒩M\mathcal{N}_{M^{\prime}}. Every cone in 𝒩M\mathcal{N}_{M} that is contained in the boundary of coneΘ\operatorname{cone}\Theta is also a subset of a cone in 𝒩M\mathcal{N}_{M^{\prime}} because it is a subset of a full-dimensional cone in 𝒩M\mathcal{N}_{M}, which meets intconeΘ\operatorname{int}\operatorname{cone}\Theta. That the cones in 𝒩M\mathcal{N}_{M} are subsets of the cones in 𝒩M\mathcal{N}_{M^{\prime}} implies 𝒩M𝒩M\mathcal{N}_{M^{\prime}}\preccurlyeq\mathcal{N}_{M} ([78, Proposition 2]).

(2) \implies (3). We first show condition (1) in the definition of a deformation, i.e., MM^{\prime} can be defined using translates of the facet-defining halfspaces of MM. First, since every cone in 𝒩M\mathcal{N}_{M} contains the orthogonal complement of affMa\operatorname{aff}M-a, where aMa\in M is arbitrary, the same must be true for the cones in 𝒩M\mathcal{N}_{M^{\prime}}. Thus, affM\operatorname{aff}M^{\prime} must be contained in a translate of affM\operatorname{aff}M and therefore the same normal vectors used to define affM\operatorname{aff}M can be used to define affM\operatorname{aff}M^{\prime}. Second, 𝒩M𝒩M\mathcal{N}_{M^{\prime}}\preccurlyeq\mathcal{N}_{M} implies that the cones in 𝒩M\mathcal{N}_{M^{\prime}} corresponding to the facets of MM^{\prime} are also cones in 𝒩M\mathcal{N}_{M} because these cones can only be written as the trivial union of themselves. Every such cone contains a unique normal vector in affMa\operatorname{aff}M-a, for arbitrary aMa\in M. Thus, the same normal vectors used to define MM can be used to define MM^{\prime}, as desired.

We now show condition (2) in the definition of a deformation. Suppose iIHi={a}\cap_{i\in I}H_{i}=\{a\} for I{1,,k}I\subseteq\{1,\dots,k\} and aextMa\in\operatorname{ext}M. By (10), cone{ni}iINC{a}\operatorname{cone}\{n_{i}\}_{i\in I}\subseteq NC_{\{a\}}. Since NC{a}NC_{\{a\}} is full-dimensional and 𝒩M𝒩M\mathcal{N}_{M^{\prime}}\preccurlyeq\mathcal{N}_{M}, there exists aextMa^{\prime}\in\operatorname{ext}M^{\prime} such that NC{a}NC{a}NC_{\{a\}}\subseteq NC_{\{a^{\prime}\}}. Thus, cone{ni}iINC{a}\operatorname{cone}\{n_{i}\}_{i\in I}\subseteq NC_{\{a^{\prime}\}}. Consequently, for all iIi\in I, there exists ci=maxaMniac_{i}^{\prime}=\max_{a\in M^{\prime}}n_{i}\cdot a such that the hyperplane HiH_{i}^{\prime} with normal nin_{i} and constant cic_{i}^{\prime} supports MM^{\prime} at aa^{\prime}. In particular, iIHi={a}\cap_{i\in I}H_{i}^{\prime}=\{a^{\prime}\}, as desired.

(3) \implies (4). It is immediate from condition (2) in the definition of deformations that there is a surjective map φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime}. Moreover, by condition (2), φ\varphi must map each edge ee of MM either to an edge ee^{\prime} of MM^{\prime} that is parallel to ee or to a vertex of MM^{\prime}. This is because the hyperplanes of MM defining ee must intersect for MM^{\prime} in a translate of the line containing ee.

(4) \implies (1). Suppose M,MFinM,M^{\prime}\in\mathcal{M}^{\text{Fin}} have the properties stated in (4). Recall that menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M and menu(x)=extM\operatorname{menu}(x^{\prime})=\operatorname{ext}M^{\prime} by Corollary A.3. To show that (θ,θ~)𝒞(x)(\theta,\tilde{\theta})\in\mathcal{IC}(x) implies (θ,θ~)𝒞(x)(\theta,\tilde{\theta})\in\mathcal{IC}(x^{\prime}), it suffices to show that

argmaxaextMaθ=φ(argmaxaextMaθ)\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot\theta=\varphi\left(\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot\theta\right) (12)

for all θΘ\theta\in\Theta.

Suppose a~argmaxaextMaθ\tilde{a}\in\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot\theta. Fix any a^extM\hat{a}\in\operatorname{ext}M. By the simplex algorithm, there exists a sequence (a0,a1,,an1,an)(a_{0},a_{1},\ldots,a_{n-1},a_{n}) such that a0=a^a_{0}=\hat{a}, an=a~a_{n}=\tilde{a}, aiai1¯\overline{a_{i}a_{i-1}} is an edge of MM for all i=1,,ni=1,\ldots,n, and (aiai1)θ0(a_{i}-a_{i-1})\cdot\theta\geq 0 for all i=1,,ni=1,\ldots,n. Condition (4) implies that (φ(ai)φ(ai1))θ0(\varphi(a_{i})-\varphi(a_{i-1}))\cdot\theta\geq 0 for all i=1,,ni=1,\ldots,n. Since a^extM\hat{a}\in\operatorname{ext}M was arbitrary, φ(a~)θφ(a)θ\varphi(\tilde{a})\cdot\theta\geq\varphi(a)\cdot\theta for all aAa\in A. That is, φ(a~)argmaxaextMaθ\varphi(\tilde{a})\in\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot\theta.

Suppose a~argmaxaextMaθ\tilde{a}\notin\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot\theta. By the simplex algorithm, there exists a sequence (a0,a1,,an1,an)(a_{0},a_{1},\ldots,a_{n-1},a_{n}) such that a0=a~a_{0}=\tilde{a}, anargmaxaextMaθa_{n}\in\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot\theta, aiai1¯\overline{a_{i}a_{i-1}} is an edge of MM for all i=1,,ni=1,\ldots,n, and (aiai1)θ>0(a_{i}-a_{i-1})\cdot\theta>0 for all i=1,,ni=1,\ldots,n. Condition (4) implies that either φ(a~)θ<φ(an)θ\varphi(\tilde{a})\cdot\theta<\varphi(a_{n})\cdot\theta or φ(a~)=φ(an)\varphi(\tilde{a})=\varphi(a_{n}). In the first case, φ(a~)argmaxaextMaθ\varphi(\tilde{a})\notin\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot\theta. In the second case, repeat the argument with ana_{n} in place of a~\tilde{a}. Since |extM|<|\operatorname{ext}M|<\infty, either the procedure terminates and φ(a~)argmaxaextMaθ\varphi(\tilde{a})\notin\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot\theta or |extM|=1|\operatorname{ext}M^{\prime}|=1, in which case (12) holds trivially. ∎

We note (12) as a separate corollary for later use.

Corollary B.4.

Suppose MDef(M)M^{\prime}\in\operatorname{Def}(M). Then, there exists a surjective function φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime} such that

argmaxaextMaθ=φ(argmaxaextMaθ)\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot\theta=\varphi\left(\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot\theta\right) (13)

for all θΘ\theta\in\Theta.

We can translate the definition of deformations into a polyhedral characterization of Def(M)\operatorname{Def}(M). For each vertex aextMa\in\operatorname{ext}M, let Ia={1ilaHi}I_{a}=\{1\leq i\leq l\mid a\in H_{i}\} denote the set of indices of facet-defining hyperplanes in M={H1,,Hk}\mathcal{H}_{M}=\{H_{1},\ldots,H_{k}\} intersecting aa. Under any feasible deformation and for each aextMa\in\operatorname{ext}M, the hyperplanes in IaI_{a} still need to intersect in a single point φaA\varphi_{a}\in A. Thus, we have the following linear system with variables (φa)aextMd×|extM|(\varphi_{a})_{a\in\operatorname{ext}M}\in\mathbb{R}^{d\times|\operatorname{ext}M|} corresponding to the points in extM\operatorname{ext}M^{\prime} and variables ckc^{\prime}\in\mathbb{R}^{k} corresponding to the deformation vector of MM^{\prime}:

φani\displaystyle\varphi_{a}\cdot n_{i} =ci\displaystyle=c_{i}^{\prime}\quad aextM,iIa\displaystyle\forall a\in\operatorname{ext}M,\,\forall i\in I_{a} (14)
φani\displaystyle\varphi_{a}\cdot n_{i} ci\displaystyle\leq c_{i}^{\prime}\quad aextM,i{1,,k}Ia\displaystyle\forall a\in\operatorname{ext}M,\,\forall i\in\{1,\ldots,k\}\setminus I_{a} (15)
φanH\displaystyle\varphi_{a}\cdot n_{H} cH\displaystyle\leq c_{H}\quad aextM,H.\displaystyle\forall a\in\operatorname{ext}M,\,\forall H\in\mathcal{F}. (16)

Let us parse these (in)equalities. The inequalities in (16) capture the requirement that MM^{\prime} is a feasible extended menu, i.e., extMA\operatorname{ext}M^{\prime}\subset A. (Recall that \mathcal{F} is the set of facet-defining hyperplanes of AA.) The (in)equalities in 14 and 15 are jointly equivalent to condition (2) in the definition of a deformation. (Condition (1) is satisfied by construction: we use the facet-defining hyperplanes of MM to define MM^{\prime}.) 14 ensures that the facet-defining hyperplanes of MM intersecting aextMa\in\operatorname{ext}M still intersect in a single point φaextM\varphi_{a}\in\operatorname{ext}M under the deformation vector cc^{\prime}. 15 ensures that φaextM\varphi_{a}\in\operatorname{ext}M^{\prime}, i.e, the facet-defining halfspaces of MM still contain φa\varphi_{a} under the deformation vector cc^{\prime}. In economic terms, recalling the equivalence between MDef(M)M^{\prime}\in\operatorname{Def}(M) and the corresponding mechanisms xx and xx^{\prime} satisfying 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}) (Lemma B.3), 14 and 15 are tantamount to 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}). If none of the constraints in (15) are binding for MM^{\prime} (which is the case for MM by definition of the index sets IaI_{a}), then 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}).

Lemma B.5.

Def(M)\operatorname{Def}(M) is a polytope and a face of \mathcal{M}. In particular, MextM\in\operatorname{ext}\mathcal{M} if and only if MextDef(M)M\in\operatorname{ext}\operatorname{Def}(M).

Proof.

Note that 14, 15 and 16 define a polytope in d×|extM|×k\mathbb{R}^{d\times|\operatorname{ext}M|}\times\mathbb{R}^{k}: 14, 15 and 16 is a linear system with bounded solutions since AA is bounded. The projection onto the second factor ckc^{\prime}\in\mathbb{R}^{k} is also a polytope. By construction, there is an affine bijection between the projected polytope and Def(M)\operatorname{Def}(M) given by the deformation vectors ckc^{\prime}\in\mathbb{R}^{k}. Thus, Def(M)\operatorname{Def}(M) is a polytope, i.e., the convex hull of finitely many extended menus.

To show that Def(M)\operatorname{Def}(M) is a face of \mathcal{M}, first observe that if MFinM\in\mathcal{M}^{\text{Fin}}, M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} and M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} for some λ(0,1)\lambda\in(0,1), then M,M′′Def(M)M^{\prime},M^{\prime\prime}\in\operatorname{Def}(M). This is because the normal fan of the Minkowski sum of polyhedra is finer than the normal fans of each summand.414141An explicit reference for polyhedra is [79, Equation 2.3.1 ]. It is immediate that MextM\in\operatorname{ext}\mathcal{M} if and only if MextDef(M)M\in\operatorname{ext}\operatorname{Def}(M).

To complete the proof that Def(M)\operatorname{Def}(M) is a face of \mathcal{M}, consider any M~Def(M)\tilde{M}\in\operatorname{Def}(M). If M~=λM+(1λ)M′′\tilde{M}=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} for M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M}, then M,M′′Def(M~)M^{\prime},M^{\prime\prime}\in\operatorname{Def}(\tilde{M}) by the previous paragraph. Observe that “deformation of” is a transitive relation, hence M,M′′Def(M)M^{\prime},M^{\prime\prime}\in\operatorname{Def}(M), as desired. ∎

The polyhedral characterization of Def(M)\operatorname{Def}(M) immediately translates into an algebraic characterization of finite-menu extreme points: by Lemma B.5, MextM\in\operatorname{ext}\mathcal{M} if and only if there is a non-zero direction (t,s)d×|extM|×k(t,s)\in\mathbb{R}^{d\times|\operatorname{ext}M|}\times\mathbb{R}^{k} such that the two candidate solutions ((a±ta)aextM,(c±s))\left((a\pm t_{a})_{a\in\operatorname{ext}M},(c\pm s)\right) solve the linear system 14, 15 and 16.

Using condition (4) in Lemma B.3, we can state an equivalent algebraic characterization of finite-menu extreme points that needs only minimal information about the underlying mechanism. For a mechanism x𝒳x\in\mathcal{X}, let

E={(a,b)menu(x)×menu(x)|θΘ:{a,b}=argmaxa~menu(x)a~θ}E=\left\{(a,b)\in\operatorname{menu}(x)\times\operatorname{menu}(x)\ \middle|\ \exists\theta\in\Theta:\>\{a,b\}=\operatorname*{arg\,max}_{\tilde{a}\in\operatorname{menu}(x)}\tilde{a}\cdot\theta\right\} (17)

denote the set of pairs (a,b)(a,b) of menu items for which there exists a type whose favorite allocations are {a,b}\{a,b\}. These are exactly the edges of the extended menu associated with xx. For an allocation amenu(x)a\in\operatorname{menu}(x), also define

(a)={HaH}.\mathcal{F}(a)=\{H\in\mathcal{F}\mid a\in H\}. (18)
Theorem B.6.

Let x𝒳x\in\mathcal{X} have finite menu size. Then xext𝒳x\in\operatorname{ext}\mathcal{X} if and only if all solutions ((φa)amenu(x),(λab)(a,b)E)d×|menu(x)|×+|E|((\varphi_{a})_{a\in\operatorname{menu}(x)},(\lambda_{ab})_{(a,b)\in E})\in\mathbb{R}^{d\times|\operatorname{menu}(x)|}\times\mathbb{R}_{+}^{|E|} to

λab(ab)\displaystyle\lambda_{ab}(a-b) =φaφb\displaystyle=\varphi_{a}-\varphi_{b} (a,b)E\displaystyle\quad\forall(a,b)\in E (19)
φanH\displaystyle\varphi_{a}\cdot n_{H} =cH\displaystyle=c_{H} amenu(x),H(a)\displaystyle\quad\forall a\in\operatorname{menu}(x),\,H\in\mathcal{F}(a) (20)
φanH\displaystyle\varphi_{a}\cdot n_{H} cH\displaystyle\leq c_{H} amenu(x),H(a)\displaystyle\quad\forall a\in\operatorname{menu}(x),\,H\notin\mathcal{F}(a) (21)

are the trivial solutions where {φa}amenu(x)=extM\{\varphi_{a}\}_{a\in\operatorname{menu}(x)}=\operatorname{ext}M and λab=1\lambda_{ab}=1 for all (a,b)E(a,b)\in E.

Remark.

If ab¯\overline{ab} and ab¯\overline{a^{\prime}b^{\prime}} are not parallel for all (a,b),(a,b)E(a,b),(a^{\prime},b^{\prime})\in E, then the trivial solution is unique.

Proof.

Let MFinM\in\mathcal{M}^{\text{Fin}} be the extended menu associated with the finite-menu extreme point xext𝒳x\in\operatorname{ext}\mathcal{X}. By Corollary A.3, menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M. By Lemma B.5, MextDef(M)M\in\operatorname{ext}\operatorname{Def}(M). An extreme point of a polytope in Euclidean space is uniquely determined from its incident facets, i.e., binding constraints. If φa=a\varphi_{a}=a for all aextMa\in\operatorname{ext}M and c=cc^{\prime}=c, then the constraints (15) are all slack by the definition of the index sets IaI_{a}. Thus, ((a)aextM,c)((a)_{a\in\operatorname{ext}M},c) must be the unique solution to 14, 15, 20 and 21. By Lemma B.3, there exist (λab)(a,b)E|E|(\lambda_{ab})_{(a,b)\in E}\in\mathbb{R}^{|E|} such that ((φa)aextM,(λab)(a,b)E)((\varphi_{a})_{a\in\operatorname{ext}M},(\lambda_{ab})_{(a,b)\in E}) solve (19) if and only if there exists a permutation ξ:extMextM\xi:\operatorname{ext}M\to\operatorname{ext}M and ckc^{\prime}\in\mathbb{R}^{k} such that ((φξ(a))aextM,c)((\varphi_{\xi(a)})_{a\in\operatorname{ext}M},c^{\prime}) solve 14 and 15. ∎

B.1. Relation to MV

We summarize here, for readers of MV, how our results generalize their findings about the facial structure of IC mechanisms and their algebraic characterization of finite-menu extreme points to arbitrary linear screening problems.

Our Lemmas B.3 and B.5 generalize Theorems 17, 19, and 20 in MV. MV show for the multi-good monopoly problem that the decomposing summands of a finite-menu IC mechanism must have a coarser market segmentation (in our language: normal fan) than the mechanism itself (Theorem 17). In their Definition 18, MV then define the set of all IC mechanisms with a coarser market segmentation than a given IC mechanism that also satisfy an analogue of (20), i.e., have at least the same binding feasibility constraints as the given mechanism. This set is the analogue of our deformation polytope Def(M)\operatorname{Def}(M), modulo (20). MV show that the set is a face of the set of IC mechanisms (Theorem 19). We further show that the set is a polytope, which immediately gives us MV’s key technical result (Theorem 20): a finite-menu IC mechanism is an extreme point if and only if it is the singleton element of their set (i.e., Def(M)\operatorname{Def}(M) plus (20)).

Our polyhedral characterization 14, 15 and 16 of Def(M)\operatorname{Def}(M) immediately translates into the algebraic characterization of finite-menu extreme points given in Theorem B.6, generalizing Theorem 24 in MV. In our result, (19) generalizes condition (13) in MV; (20) generalizes condition (14) in MV; (21) generalizes the condition “𝟎za𝟏\mathbf{0}\leq z^{a}\leq\mathbf{1}” in MV (which is feasibility for the monopoly problem); in our model, za=φaz^{a}=\varphi_{a}. Theorem B.6 amends a minor oversight in MV in that multiple solutions of 19, 20 and 21 can correspond to the same extreme point xext𝒳x\in\operatorname{ext}\mathcal{X} because there need not be a unique assignment of the variables (φa)amenu(x)(\varphi_{a})_{a\in\operatorname{menu}(x)} to the menu items whenever the menu has parallel edges.

Appendix C Extreme Points for One-Dimensional Type Spaces

We deduce Theorem 6.1 from a general characterization of the extreme points for one-dimensional linear screening problems. We state the characterization in terms of extended menus. By Theorem A.2, we could equivalently state it in terms of menus of mechanisms. The key concept in the characterization—a flexible chain—requires some notation to be defined. Let us first state the result, then define the concept, and then give the proof.

Theorem C.1.

Let d=2d=2 and MM\in\mathcal{M}. Then MextM\in\operatorname{ext}\mathcal{M} if and only if

  1. (1)

    |extM|2|\operatorname{ext}M|\leq 2 and MM is exhaustive, or

  2. (2)

    3|extM|<3\leq|\operatorname{ext}M|<\infty and MM has no flexible chain.

Remark.

The theorem is an extension of a result due to [93, Theorem 3.1 ]. The result characterizes extremal convex bodies (ext\operatorname{ext}\mathcal{M}) contained in a given convex body in the plane (AA). If coneΘ=2\operatorname{cone}\Theta=\mathbb{R}^{2} is unrestricted, then we can use Mielczarek’s Theorem.424242Specifically, in Mielczarek’s theorem, condition 11^{\circ} is equivalent to condition (1) above; if V(M)V(M)\neq\emptyset, then conditions 22^{\circ} and (i) are equivalent to the absence of a flexible chain; if V(M)=V(M)=\emptyset, then conditions 22^{\circ} and (ii) are equivalent to the absence of a flexible chain. Condition (iii) in Mielczarek’s theorem never applies if AA (QQ in the statement) is a polytope. Otherwise, if coneΘ2\operatorname{cone}\Theta\neq\mathbb{R}^{2} is restricted, we have to make a minor modification to the result because we consider closed convex sets MM\in\mathcal{M} with extreme points in AA. In any case, the original presentation of the result and its proof are notationally tedious, so we have restated and shall reprove most of the result for the reader’s convenience.

To get a first sense of a flexible chain, recall Figure 1 (Section 4). This figure illustrates a non-extreme point that can be deformed by horizontally translating the right-most vertical edge in its menu. The two vertices of this edge form a flexible chain in the sense of Theorem C.1. However, a menu may lack an edge that can be flexibly translated in both normal directions, yet the corresponding mechanism may still not be an extreme point. This is because multiple edges could potentially be translated jointly, which is the idea captured by a flexible chain. The formal definition of a flexible chain requires some new notation.

v1v_{1}v2v_{2}v3v_{3}v4v_{4}v5v_{5}
v2v_{2}v3v_{3}v4v_{4}v1v_{1}β2\beta_{2}α2\alpha_{2}β1\beta_{1}α1\alpha_{1}β4\beta_{4}α4\alpha_{4}α3\alpha_{3}β3\beta_{3}
Figure 7. An illustration of flexible chains and their connection to extreme points. Left: an extended menu MM, depicted as the shaded area, with a flexible chain S=(v2,v3,v4,v5)S=(v_{2},v_{3},v_{4},v_{5}) and two deformations of MM, depicted with dotted lines, that decompose MM. Right: an extended menu MM, depicted as the shaded area, with a chain S=(v1,v2,v3,v4)S=(v_{1},v_{2},v_{3},v_{4}) that is not flexible because it violates the symmetry condition 23 on the angles (αk)k(\alpha_{k})_{k} and (βk)k(\beta_{k})_{k}. Intuitively, the start- and endpoints of a candidate deformation coincide only under the symmetry condition.

For the following definitions, let d=2d=2 and fix an extended menu MM\in\mathcal{M} of finite menu size |extM|<|\operatorname{ext}M|<\infty. Recall that MM is a polyhedron that satisfies M=convextM+ΘM=\operatorname{conv}\operatorname{ext}M+\Theta^{\circ}. The vertices of any polyhedron in the plane can be ordered clockwise and adjacent vertices in the ordering are connected by an edge. If MM\in\mathcal{M} is unbounded, i.e., coneΘ2\operatorname{cone}\Theta\neq\mathbb{R}^{2}, we designate a placeholder * as the first and last vertex in the ordering (which can be thought of as a vertex at infinity).

We define four disjoint subsets V(M),I(M),B1(M),B2(M)extMV(M),I(M),B_{1}(M),B_{2}(M)\subseteq\operatorname{ext}M such that

extM=V(M)I(M)B1(M)B2(M).\operatorname{ext}M=V(M)\cup I(M)\cup B_{1}(M)\cup B_{2}(M). (22)

V(M)=extMextAV(M)=\operatorname{ext}M\cap\operatorname{ext}A. I(M)=extMintAI(M)=\operatorname{ext}M\cap\operatorname{int}A. B1(M)B_{1}(M) is the set of vertices aextM(bndrAextA)a\in\operatorname{ext}M\cap(\operatorname{bndr}A\setminus\operatorname{ext}A) such that there is no other vertex bextMb\in\operatorname{ext}M for which ab¯bndrA\overline{ab}\subset\operatorname{bndr}A. B2(M)=(bndrAextA)B1(M)B_{2}(M)=(\operatorname{bndr}A\setminus\operatorname{ext}A)\setminus B_{1}(M) is the set of vertices aextM(bndrAextM)a\in\operatorname{ext}M\cap(\operatorname{bndr}A\setminus\operatorname{ext}M) for which such a vertex bextMb\in\operatorname{ext}M does exist.

Example C.2.

We illustrate the definition of these subsets with several examples. In the left panel of Figure 7, v1V(M)v_{1}\in V(M), v2B2(M)v_{2}\in B_{2}(M), v3,v4B1(M)v_{3},v_{4}\in B_{1}(M), and v5I(M)v_{5}\in I(M). In the right panel of Figure 7, all vertices are in B1(M)B_{1}(M). In Figure 5, v1V(M)v_{1}\in V(M), v5I(M)v_{5}\in I(M), and v4B1(M)v_{4}\in B_{1}(M).

We define the following angles formed by the edges of MM with the edges of the allocation polytope AA. Let vB1(M)v\in B_{1}(M). Let uu and ww be the vertices preceding and succeeding vv in the clock-wise ordering, respectively. Let ab¯\overline{ab} be the edge of AA on which vv lies, where aa preceeds bb in the clock-wise ordering. Let αk\alpha_{k} be the measure of the angle uva\angle uva, and let βk\beta_{k} be the measure of the angle wvb\angle wvb; see the right panel in Figure 7 for an illustration.434343That is, αk=cos1((uv)(av)uvav)\alpha_{k}=\cos^{-1}\left(\frac{(u-v)\cdot(a-v)}{||u-v||||a-v||}\right) and βk=cos1((wv)(bv)wvbv).\beta_{k}=\cos^{-1}\left(\frac{(w-v)\cdot(b-v)}{||w-v||||b-v||}\right).

Definition C.3.

A sequence S=(v1,,vn)S=(v_{1},\ldots,v_{n}) of vertices of MM that are adjacent in the clock-wise ordering is a flexible chain if SV(M)=S\cap V(M)=\emptyset and one of the following holds:

  1. (1)

    v1,vnI(M)B2(M){}v_{1},v_{n}\in I(M)\cup B_{2}(M)\cup\{*\}, and if n=2n=2, then v1vn¯bndrA\overline{v_{1}v_{n}}\not\subset\operatorname{bndr}A;

  2. (2)

    S=extM=B1(M)S=\operatorname{ext}M=B_{1}(M), coneΘ=2\operatorname{cone}\Theta=\mathbb{R}^{2}, nn is even, and

    k=1nsinαk=k=1nsinβk.\prod_{k=1}^{n}\sin\alpha_{k}=\prod_{k=1}^{n}\sin\beta_{k}. (23)
Example C.4.

We illustrate the definition of a flexible chain with several examples. In the left panel of Figure 7, S=(v2,v3,v4,v5)S=(v_{2},v_{3},v_{4},v_{5}) forms a flexible chain. In the right panel of Figure 7, S=(v1,v2,v3,v4)S=(v_{1},v_{2},v_{3},v_{4}) does not form a flexible chain because the symmetry condition (23) is violated. In contrast, the vertices of a 4545^{\circ} rotation of the allocation square would form a flexible chain. In Figure 5, (,v4,v5)(*,v_{4},v_{5}) forms a flexible chain. Indeed, the extended menu depicted there for the one-good monopoly problem has menu size 3. It is well-known that the corresponding mechanism cannot be an extreme point, i.e., the extended menu must have deformations that decompose it. This observation is generalized in Theorem C.1.

Proof of Theorem C.1.

Suppose |extM|2|\operatorname{ext}M|\leq 2. If extM\operatorname{ext}M is a singleton, then MextM\in\operatorname{ext}\mathcal{M} if and only if MM is exhaustive. Suppose convextM\operatorname{conv}\operatorname{ext}M is a line segment and M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} for M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} and λ(0,1)\lambda\in(0,1). Then, MM^{\prime} and M′′M^{\prime\prime} are homothetic to MM because they must be deformations of MM by Lemma B.5. Using Theorem 5.2, MextM\in\operatorname{ext}\mathcal{M} if and only if MM is exhaustive.

Thus, suppose |extM|3|\operatorname{ext}M|\geq 3. We first show that if MextM\in\operatorname{ext}\mathcal{M}, then |extM|<|\operatorname{ext}M|<\infty. Let U𝒰U\in\mathcal{U} be the indirect utility function associated with MM (i.e., the support function of MM).

For the sake of contradiction, suppose |extM|=|\operatorname{ext}M|=\infty. Since AA has only finitely many edges and on each edge of AA there can be at most two vertices of MM, |extMintA|=|\operatorname{ext}M\cap\operatorname{int}A|=\infty. In particular, there must exist an open cone CconeΘC\subseteq\operatorname{cone}\Theta such that extMCintA\operatorname{ext}M\cap C\subset\operatorname{int}A and |extMC|=|\operatorname{ext}M\cap C|=\infty. Let LL be an open line segment such that coneL=C\operatorname{cone}L=C. Let γ:(0,1)L\gamma:(0,1)\to L be a bijective isometry. Consider the convex function Uγ:(0,1)U\circ\gamma:(0,1)\to\mathbb{R}, which completely determines UU on CC by 1-homogeneity.

Suppose there exist convex functions U1,U2:(0,1)U_{1},U_{2}:(0,1)\to\mathbb{R} such that Uγ=12U1+12U2U\circ\gamma=\frac{1}{2}U_{1}+\frac{1}{2}U_{2}, U1(t)=U2(t)=U(t)U_{1}(t)=U_{2}(t)=U(t) for all t[ε,1ε]t\notin[\varepsilon,1-\varepsilon], where ε>0\varepsilon>0 is sufficiently small, and the right-derivatives of Uγ,U1U\circ\gamma,U_{1}, and U2U_{2}, respectively, are the same at ε\varepsilon, and the left-derivatives of Uγ,U1U\circ\gamma,U_{1}, and U2U_{2}, respectively, are the same at 1ε1-\varepsilon. Then, U1U_{1} and U2U_{2} can be extended to sublinear functions on coneΘ\operatorname{cone}\Theta such that U=12U1+12U2U=\frac{1}{2}U_{1}+\frac{1}{2}U_{2} by first extending the functions to CC by 1-homogeneity and then to coneΘ\operatorname{cone}\Theta by setting U|coneΘC=U1|coneΘC=U2|coneΘCU|_{\operatorname{cone}\Theta\setminus C}=U_{1}|_{\operatorname{cone}\Theta\setminus C}=U_{2}|_{\operatorname{cone}\Theta\setminus C}. If U1U_{1} and U2U_{2} are sufficiently close to UγU\circ\gamma, then their extensions are in 𝒰\mathcal{U} because U(L)intA\partial U(L)\subset\operatorname{int}A.

We now show that the convex functions U1,U2:(0,1)U_{1},U_{2}:(0,1)\to\mathbb{R} from the previous paragraph exist, contradicting that Uext𝒰U\in\operatorname{ext}\mathcal{U}. Consider the set 𝒢\mathcal{G} of convex functions g:[ε,1ε]g:[\varepsilon,1-\varepsilon]\to\mathbb{R} such that (1) g(ε)=(Uγ)(ε)g(\varepsilon)=(U\circ\gamma)(\varepsilon), (2) g(1ε)=(Uγ)(1ε)g(1-\varepsilon)=(U\circ\gamma)(1-\varepsilon), (3) g+(ε)=(Uγ)+(ε)g^{\prime}_{+}(\varepsilon)=(U\circ\gamma)^{\prime}_{+}(\varepsilon), where g+g^{\prime}_{+} is the right-derivative, and (4) g(1ε)=(Uγ)(1ε)g^{\prime}_{-}(1-\varepsilon)=(U\circ\gamma)^{\prime}_{-}(1-\varepsilon), where gg^{\prime}_{-} is the left-derivative. By combining a well-known result due to [22] about extremal convex functions on \mathbb{R} and a result due to [123] about the extreme points of convex sets obtained from a given convex set by imposing finitely many affine restrictions, one can show the the extreme points of 𝒢\mathcal{G} are piecewise-affine with at most three pieces. Thus, Uext𝒰U\notin\operatorname{ext}\mathcal{U} and MextM\notin\operatorname{ext}\mathcal{M} since |extMcone(γ([ε,1ε]))||\operatorname{ext}M\cap\operatorname{cone}(\gamma([\varepsilon,1-\varepsilon]))| can be made arbitrarily large by choosing ε>0\varepsilon>0 small enough.

Suppose coneΘ2\operatorname{cone}\Theta\neq\mathbb{R}^{2} or extMB1\operatorname{ext}M\neq B_{1}. By Theorem 4.1 and Lemma B.5, MextM\notin\operatorname{ext}\mathcal{M} and |extM|<|\operatorname{ext}M|<\infty if and only if MM has a deformation MDef(M)M^{\prime}\in\operatorname{Def}(M) such that (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}). Thus, it suffices to show that MM has a deformation MDef(M)M^{\prime}\in\operatorname{Def}(M) such that (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}) if and only if MM has a flexible chain.

By Lemma B.5, MDef(M)M^{\prime}\in\operatorname{Def}(M) if and only if the facet-defining hyperplanes (lines) of MM^{\prime} are parallel translates of the facet-defining hyperplanes of MM and there is a surjective map φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime}. By taking a convex combination εM+(1ε)M\varepsilon M^{\prime}+(1-\varepsilon)M for ε>0\varepsilon>0 sufficently small, we may assume that φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime} is bijective.

For (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}) to hold, (v,φ(v))(v,\varphi(v)) must lie on the same face of AA. In particular, φ(a)=a\varphi(a)=a for all aV(M)a\in V(M), φ(I(M))=I(M)\varphi(I(M))=I(M^{\prime}), φ(B1(M))=B1(M)\varphi(B_{1}(M))=B_{1}(M^{\prime}), and φ(B2(M))=B2(M)\varphi(B_{2}(M))=B_{2}(M^{\prime}).

We observe that if vB1(M)v\in B_{1}(M) and vφ(v)v\neq\varphi(v), then the two facet-defining hyperplanes of MM intersecting in vv must both be translated in MM^{\prime} for otherwise (v,φ(v))(v,\varphi(v)) cannot lie on a common edge of AA.

Consider a deformation MDef(M)M^{\prime}\in\operatorname{Def}(M) such that (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}); we construct a flexible chain of MM. Find a sequence S=(v1,,vn)S=(v_{1},\ldots,v_{n}) of vertices in extMφ(extM)\operatorname{ext}M\setminus\varphi(\operatorname{ext}M) that are adjacent in the clock-wise ordering and such that no other vertex in extMφ(extM)\operatorname{ext}M\setminus\varphi(\operatorname{ext}M) is adjacent to a vertex in SS. SV(M)=S\cap V(M)=\emptyset follows since φ(a)=a\varphi(a)=a for all aV(M)a\in V(M). If n=2n=2, then v1vn¯bndrA\overline{v_{1}v_{n}}\not\subset\operatorname{bndr}A for otherwise the edge φ(v1)φ(vn)¯\overline{\varphi(v_{1})\varphi(v_{n})} is not in bndrA\operatorname{bndr}A, contradicting (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}). If v1,vnB1(M)v_{1},v_{n}\in B_{1}(M), then, by the previous paragraph, v1v_{1} and vnv_{n} cannot be the first or last vertex in the sequence, contradicting the construction of SS.

Conversely, suppose MM\in\mathcal{M} has a flexible chain (v1,,vn)(v_{1},\ldots,v_{n}). We carry out the construction illustrated in Figure 7. Without loss of generality, we may assume v2,,vn1B1(M)v_{2},\ldots,v_{n-1}\in B_{1}(M) for otherwise, (v1,,vn)(v_{1},\ldots,v_{n}) has a subsequence of adjacent vertices that is a flexible chain with the desired property. Let (H1,,Hn1)(H_{1},\ldots,H_{n-1}) be the hyperplanes such that HiH_{i} defines the facet vivi+1¯\overline{v_{i}v_{i+1}} for all i=1,,n1i=1,\ldots,n-1.

Suppose v1v_{1}\neq*. Let H0H_{0} be the other hyperplane of MM intersecting v1v_{1}. Translate H1H_{1} by a sufficiently small amount, and let φ1\varphi_{1} be the intersection of H1H_{1}^{\prime} and H0H_{0}. Since v1intAB2v_{1}\in\operatorname{int}A\cup B_{2}, φ1\varphi_{1} lies on the same face of AA as v1v_{1}. If v1=v_{1}=*, translate H1H_{1} by a sufficiently small amount to obtain H1H_{1}^{\prime}.

Let φ2\varphi_{2} be the intersection of H1H_{1}^{\prime} with the edge of AA on which v2v_{2} lies. (This intersection is non-empty as long as all translations are sufficiently small.) Let H2H_{2}^{\prime} be the translate of H2H_{2} that intersects φ2\varphi_{2}.

Iterate the construction in the previous paragraphs to obtain a sequence of points (φ1,,φn)(\varphi_{1},\ldots,\varphi_{n}) and hyperplanes (H1,,Hn1)(H_{1}^{\prime},\ldots,H_{n-1}^{\prime}). Since vnintAB2{}v_{n}\in\operatorname{int}A\cup B_{2}\cup\{*\}, the hyperplane HnHn1H_{n}\neq H_{n-1} intersecting vnv_{n} need not be translated to meet φn\varphi_{n}. Define MM^{\prime} as the polyhedron whose edges are defined by (H1,,Hn1)(H_{1}^{\prime},\ldots,H_{n-1}^{\prime}) and by the facet-defining hyperplanes of MM different from (H1,,Hn1)(H_{1},\ldots,H_{n-1}). By construction, MDef(M)M^{\prime}\in\operatorname{Def}(M) and (M)=(M)\mathcal{F}(M)=\mathcal{F}(M^{\prime}).

It remains to consider the case where extM=B1(M)\operatorname{ext}M=B_{1}(M) and coneΘ=2\operatorname{cone}\Theta=\mathbb{R}^{2}. We refer the reader to Lemmas 9, 17, and 18 in [93] for the formal proof that MextM\notin\operatorname{ext}\mathcal{M} if and only if the symmetry condition (23) holds. We illustrate the idea in the right panel of Figure 7: if (23) were to hold, then the dotted chain of line segments would have the same start- and endpoints, i.e., would become a deformation of the depicted extended menu. ∎

Appendix D Proofs

This appendix gathers the proofs for the results in the main text in the order of appearance. By Theorem A.2, we may prove all results either for the set of IC and IR mechanisms 𝒳\mathcal{X}, the set of extended menus \mathcal{M}, or the set of indirect utility functions 𝒰\mathcal{U}.

D.1. Proofs for Section 4

We note the following observation.

Lemma D.1.

Suppose x=λx+(1λ)x′′x=\lambda x^{\prime}+(1-\lambda)x^{\prime\prime} for x,x,x′′𝒳x,x^{\prime},x^{\prime\prime}\in\mathcal{X} of finite menu size and λ(0,1)\lambda\in(0,1). Then, 𝒞(x)=𝒞(x)𝒞(x′′)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime})\cap\mathcal{IC}(x^{\prime\prime}) and (x)=(x)(x′′)\mathcal{F}(x)=\mathcal{F}(x^{\prime})\cap\mathcal{F}(x^{\prime\prime}).

Proof.

Let M,M,M′′M,M^{\prime},M^{\prime\prime}\in\mathcal{M} be the extended menus associated with xx, xx^{\prime}, and x′′x^{\prime\prime}, respectively. M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} by Theorem A.2. For ZdZ\subset\mathbb{R}^{d}, let Top(Z,θ)=argmaxaZaθ\operatorname{Top}(Z,\theta)=\operatorname*{arg\,max}_{a\in Z}a\cdot\theta.

By Corollary A.3, menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M. Thus, (θ,θ)𝒞(x)(\theta,\theta^{\prime})\in\mathcal{IC}(x), i.e., Top(menu(x),θ)Top(menu(x),θ)\operatorname{Top}(\operatorname{menu}(x),\theta^{\prime})\subseteq\operatorname{Top}(\operatorname{menu}(x),\theta), if and only if Top(extM,θ)Top(extM,θ)\operatorname{Top}(\operatorname{ext}M,\theta^{\prime})\subseteq\operatorname{Top}(\operatorname{ext}M,\theta).

We first show 𝒞(x)𝒞(x)𝒞(x′′)\mathcal{IC}(x)\supseteq\mathcal{IC}(x^{\prime})\cap\mathcal{IC}(x^{\prime\prime}). Suppose (θ,θ)𝒞(x)𝒞(x′′)(\theta,\theta^{\prime})\in\mathcal{IC}(x^{\prime})\cap\mathcal{IC}(x^{\prime\prime}). Then,

Top(extM,θ)\displaystyle\operatorname{Top}(\operatorname{ext}M^{\prime},\theta^{\prime}) Top(extM,θ)\displaystyle\subseteq\operatorname{Top}(\operatorname{ext}M^{\prime},\theta)
Top(extM′′,θ)\displaystyle\operatorname{Top}(\operatorname{ext}M^{\prime\prime},\theta^{\prime}) Top(extM′′,θ).\displaystyle\subseteq\operatorname{Top}(\operatorname{ext}M^{\prime\prime},\theta).

Thus,

Top(λextM+(1λ)extM′′,θ)Top(λextM+(1λ)extM′′,θ).\operatorname{Top}(\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime},\theta^{\prime})\subseteq\operatorname{Top}(\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime},\theta).

Since extMλextM+(1λ)extM′′\operatorname{ext}M\subseteq\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime}, we conclude

Top(extM,θ)Top(extM,θ)\operatorname{Top}(\operatorname{ext}M,\theta^{\prime})\subseteq\operatorname{Top}(\operatorname{ext}M,\theta)

or, equivalently, (θ,θ)𝒞(x)(\theta,\theta^{\prime})\in\mathcal{IC}(x).

We next show 𝒞(x)𝒞(x)𝒞(x′′)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime})\cap\mathcal{IC}(x^{\prime\prime}). By interchanging the roles of xx^{\prime} and x′′x^{\prime\prime}, it suffices to show that (θ,θ)𝒞(x)(\theta,\theta^{\prime})\notin\mathcal{IC}(x^{\prime}) implies (θ,θ)𝒞(x)(\theta,\theta^{\prime})\notin\mathcal{IC}(x). Assume Top(extM,θ)Top(extM,θ)\operatorname{Top}(\operatorname{ext}M^{\prime},\theta^{\prime})\setminus\operatorname{Top}(\operatorname{ext}M^{\prime},\theta)\neq\emptyset, i.e., (θ,θ)𝒞(x)(\theta,\theta^{\prime})\notin\mathcal{IC}(x^{\prime}). Then,

Top(λextM+(1λ)extM′′,θ)Top(λextM+(1λ)extM′′,θ).\operatorname{Top}(\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime},\theta^{\prime})\setminus\operatorname{Top}(\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime},\theta)\neq\emptyset.

Since convextM=conv(λextM+(1λ)extM′′)\operatorname{conv}\operatorname{ext}M=\operatorname{conv}(\lambda\operatorname{ext}M^{\prime}+(1-\lambda)\operatorname{ext}M^{\prime\prime}) and utility is linear, we conclude

Top(extM,θ)Top(extM,θ)\operatorname{Top}(\operatorname{ext}M,\theta^{\prime})\setminus\operatorname{Top}(\operatorname{ext}M,\theta)\neq\emptyset

or, equivalently, (θ,θ)𝒞(x)(\theta,\theta^{\prime})\notin\mathcal{IC}(x).

(x)=(x)(x′′)\mathcal{F}(x)=\mathcal{F}(x^{\prime})\cap\mathcal{F}(x^{\prime\prime}) is immediate. If one summand is bounded way from a hyperplane, then the the convex combination must also be bounded away from the hyperplane. Conversely, if both summands make allocations on the same hyperplane, then so does their convex combination. ∎

The following proof uses the polyhedral characterization of Def(M)\operatorname{Def}(M) given by (14), (15), and (16) in Appendix B. The proof idea is described right after the statement in the main text.

Proof of Theorem 4.1.

The remark following Theorem 4.1 is immediate from Lemma D.1: for x𝒳x\in\mathcal{X}, if another x𝒳x^{\prime}\in\mathcal{X} satisfies 𝒞(x)𝒞(x)\mathcal{IC}(x)\subseteq\mathcal{IC}(x^{\prime}) and (x)(x)\mathcal{F}(x)\subseteq\mathcal{F}(x^{\prime}), then x′′=εx+(1ε)xx^{\prime\prime}=\varepsilon x^{\prime}+(1-\varepsilon)x for 0<ε<10<\varepsilon<1 satisfies 𝒞(x)=𝒞(x′′)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime\prime}) and (x)=(x′′)\mathcal{F}(x)=\mathcal{F}(x^{\prime\prime}).

Necessity is also immediate from Lemma D.1: if xext𝒳x\notin\operatorname{ext}\mathcal{X}, then the summands in the decomposition make weakly more constraints binding.

For sufficiency, suppose 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}) for x,x𝒳x,x^{\prime}\in\mathcal{X} of finite menu size. Let M,MM,M^{\prime}\in\mathcal{M} be the associated extended menus. By Lemma B.3, MM and MM^{\prime} are mutual deformations. In particular, there is a bijection φ:extMextM\varphi:\operatorname{ext}M\to\operatorname{ext}M^{\prime}.

As an intermediate observation, we claim that (x)=(x)\mathcal{F}(x)=\mathcal{F}(x^{\prime}) implies, for all amenu(x)a\in\operatorname{menu}(x) and HH\in\mathcal{F}, aHa\in H if and only if φ(a)H\varphi(a)\in H. In words, each menu item of xx makes the same feasibility constraints binding as the corresponding menu item in xx^{\prime}. We have H(M)=(x)H\in\mathcal{F}(M)=\mathcal{F}(x) if and only if maxaextManH=cH\max_{a\in\operatorname{ext}M}a\cdot n_{H}=c_{H}, where nHn_{H} is the normal vector and cHc_{H} the right-hand side constant of the hyperplane HH\in\mathcal{F}. Analogously, H(M)=(x)H\in\mathcal{F}(M^{\prime})=\mathcal{F}(x^{\prime}) if and only if maxaextManH=cH\max\limits_{a\in\operatorname{ext}M^{\prime}}a\cdot n_{H}=c_{H}. The proof of the claim is completed using Corollary B.4, which gives

argmaxaextManH=φ(argmaxaextManH).\operatorname*{arg\,max}_{a\in\operatorname{ext}M^{\prime}}a\cdot n_{H}=\varphi(\operatorname*{arg\,max}_{a\in\operatorname{ext}M}a\cdot n_{H}). (24)

We complete the proof of Theorem 4.1 using the polyhedral characterization of Def(M)\operatorname{Def}(M) given by (14), (15), and (16). Let cc and cc^{\prime} denote the deformation vectors associated with MM and MM^{\prime}, respectively. By the previous paragraph, (x)=(x)\mathcal{F}(x)=\mathcal{F}(x^{\prime}) if and only if the variables (φa=φ(a))aextA(\varphi_{a}=\varphi(a))_{a\in\operatorname{ext}A} make the same constraints in (16) binding as the variables (a)aextA(a)_{a\in\operatorname{ext}A}. 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}) if and only if the variables (c,(a)aextM)(c,(a)_{a\in\operatorname{ext}M}) and (c,(φa)aextM)(c^{\prime},(\varphi_{a})_{a\in\operatorname{ext}M}) both satisfy the constraints in (14) and make none of the constraints in (15) binding. (See the explanation of the constraints in Appendix B.) Thus, (x)=(x)\mathcal{F}(x)=\mathcal{F}(x^{\prime}) and 𝒞(x)=𝒞(x)\mathcal{IC}(x)=\mathcal{IC}(x^{\prime}) if and only if (c,(a)aextM)(c,(a)_{a\in\operatorname{ext}M}) and (c,(φa)aextM)(c^{\prime},(\varphi_{a})_{a\in\operatorname{ext}M}) make the same constraints of Def(M)\operatorname{Def}(M) binding. The latter is equivalent to M,MextDef(M)M,M^{\prime}\notin\operatorname{ext}\operatorname{Def}(M) because Def(M)\operatorname{Def}(M) is a polytope by Lemma B.5. Finally, M,MextDef(M)M,M^{\prime}\notin\operatorname{ext}\operatorname{Def}(M) if and only if x,xext𝒳x,x^{\prime}\notin\operatorname{ext}\mathcal{X}. ∎

D.2. Proofs for Section 5

Recall that by Theorem A.2, the definitions of homothety and exhaustiveness translate straightforwardly to extended menus MM\in\mathcal{M}, where (M)\mathcal{F}(M)\subseteq\mathcal{F} was defined to be the set of facet-defining hyperplanes of AA intersected by extM\operatorname{ext}M. Also recall that M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime} is a homothetic decomposition of MM if λ(0,1)\lambda\in(0,1) and M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} are homothetic to but distinct from MM.

Lemma D.2.

MM\in\mathcal{M} is exhaustive if and only if MM has no homothetic decomposition.

Proof.

Suppose extM={a}\operatorname{ext}M=\{a\} is a singleton. If aextAa\notin\operatorname{ext}A, then there exists aAa^{\prime}\in A on the same faces of AA as aa. Thus, (a+Θ)=(M)\mathcal{F}(a^{\prime}+\Theta^{\circ})=\mathcal{F}(M) and a+Θa^{\prime}+\Theta^{\circ} is homothetic to MM. Thus, MM is not exhaustive. Conversely, if MM is not exhaustive, then there exists MM^{\prime} homothetic to MM, i.e., M=t+λ(M+Θ)=t+λa+ΘM^{\prime}=t+\lambda(M+\Theta^{\circ})=t+\lambda a+\Theta^{\circ}, such that t+λat+\lambda a meets an inclusion-wise larger set of hyperplanes in \mathcal{F} than aa, which implies aextAa\notin\operatorname{ext}A.

Suppose extM={a}\operatorname{ext}M=\{a\} is not a singleton. As an intermediate step, we will show that the set

HC(M)={(λ,t)+×dext(λM+t)A}\operatorname{HC}(M)=\{(\lambda,t)\in\mathbb{R_{+}}\times\mathbb{R}^{d}\mid\operatorname{ext}(\lambda M+t)\subset A\}

of all (parameters of) homotheties of MM is a polytope. HC(M)\operatorname{HC}(M) is bounded because AA is bounded. Therefore, we show that HC(M)\operatorname{HC}(M) is the intersection of finitely many halfspaces. For this, let

HC(M,H)={(λ,t)+×dext(λM+t)H},\operatorname{HC}_{-}(M,H)=\{(\lambda,t)\in\mathbb{R_{+}}\times\mathbb{R}^{d}\mid\operatorname{ext}(\lambda M+t)\subset H_{-}\},

where H={zd:znHcH}H_{-}=\{z\in\mathbb{R}^{d}:z\cdot n_{H}\leq c_{H}\} is the halfspace that contains AA and is bounded by the facet-defining hyperplane HH\in\mathcal{F} of AA. Let HC(M,H)\operatorname{HC}(M,H) denote the associated hyperplane. Equivalently,

HC(M,H)={(λ,t)×d|λmaxaextManH+tnHcH}.\operatorname{HC}_{-}(M,H)=\left\{(\lambda,t)\in\mathbb{R}\times\mathbb{R}^{d}\ \middle|\ \lambda\max_{a\in\operatorname{ext}M}a\cdot n_{H}+t\cdot n_{H}\leq c_{H}\right\}.

That is, HC(M,F)\operatorname{HC}_{-}(M,F) is a halfspace in d+1\mathbb{R}^{d+1} with normal (maxaextManH,nH)(\max_{a\in\operatorname{ext}M}a\cdot n_{H},n_{H}). Thus,

HC(M)=HC+HHC(M,H)\operatorname{HC}(M)=\operatorname{HC}_{+}\cap\bigcap_{H\in\mathcal{F}}\operatorname{HC}_{-}(M,H)

is a polytope, where HC+=+×d\operatorname{HC}_{+}=\mathbb{R}_{+}\times\mathbb{R}^{d}.

We complete the proof by showing that MM is exhaustive if and only if (λ,t)=(1,0)extHC(M)(\lambda,t)=(1,0)\in\operatorname{ext}\operatorname{HC}(M). Note that (1,0)(1,0) does not lie on the boundary of HC+\operatorname{HC}_{+}. Every other halfspace HC(M,H)\operatorname{HC}_{-}(M,H) of HC(M)\operatorname{HC}(M) corresponds to a facet-defining hyperplane HH of AA. Thus, MM is determined by its binding feasibility constraints (M)\mathcal{F}(M) up to homothety, i.e., exhaustive, if and only if (1,0)(1,0) lies on an inclusion-wise maximal set of facet-defining hyperplanes of HC(M)\operatorname{HC}(M). The latter condition is what it means for a point to be an extreme point of a polytope. ∎

Proof of Theorem 5.2.

Immediate from Lemma D.2. ∎

Proof of Theorem 5.3.

By Lemma D.2, MM\in\mathcal{M} is not exhaustive if and only if MM has a homothetic decomposition, i.e., there exist M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} homothetic to MM such that M=12M+12M′′M=\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime}.

Suppose extM={a}\operatorname{ext}M=\{a\} is a singleton. Then MM has a homothetic decomposition if and only if aextAa\notin\operatorname{ext}A. Thus, for the remainder of the proof, assume that extM\operatorname{ext}M is not a singleton.

MM has a homothetic decomposition if and only if one of the following holds:

  1. (1)

    There exists a point znz\in\mathbb{R}^{n} and ε>0\varepsilon>0 such that z+(1+ε)(extMz)z+(1+\varepsilon)(\operatorname{ext}M-z) and z+(1ε)(extMz)z+(1-\varepsilon)(\operatorname{ext}M-z) are both subsets of AA (dilation with center zz).

  2. (2)

    There exists a direction tn{0}t\in\mathbb{R}^{n}\setminus\{0\} such that extM+t\operatorname{ext}M+t and extMt\operatorname{ext}M-t are both subsets of AA (translation).

The reason is that any homothety is itself either a dilation or translation.444444Specifically, suppose M=12M+12M′′M=\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime} and M=z+(1+ε)(Mz)M^{\prime}=z+(1+\varepsilon)(M-z). Plugging in and rearranging for M′′M^{\prime\prime} yields M′′=z+(1ε)(Mz)M^{\prime\prime}=z+(1-\varepsilon)(M-z).

If (1) is true and aHextMa\in H\cap\operatorname{ext}M for some H(M)H\in\mathcal{F}(M), then zHz\in H, for otherwise z+(1+ε)(az)z+(1+\varepsilon)(a-z) or z+(1ε)(az)z+(1-\varepsilon)(a-z) is not in AA. Thus, if (1) is true, H(M)H\bigcap_{H\in\mathcal{F}(M)}H\neq\emptyset. Conversely, if H(M)H\bigcap_{H\in\mathcal{F}(M)}H\neq\emptyset, choose any zH(M)Hz\in\bigcap_{H\in\mathcal{F}(M)}H. For ε>0\varepsilon>0 sufficiently small, z+(1+ε)(extMz)z+(1+\varepsilon)(\operatorname{ext}M-z) and z+(1ε)(extMz)z+(1-\varepsilon)(\operatorname{ext}M-z) are both subsets of AA. This is because extM\operatorname{ext}M is uniformly bounded away from facet-defining hyperplanes H(M)H\notin\mathcal{F}(M) and because aH(M)a\in H\in\mathcal{F}(M) implies (z+(1±ε)(az))H(z+(1\pm\varepsilon)(a-z))\in H by the definition of zz, i.e., all facet-defining inequalities of AA remain satisfied.

If (2) is true, then tt is orthogonal to all the normals of the hyperplanes in (M)\mathcal{F}(M) for otherwise there is a point aextMHa\in\operatorname{ext}M\cap H, for some H(M)H\in\mathcal{F}(M), such that a+tHa+t\notin H or atHa-t\notin H, which contradicts that extM+t\operatorname{ext}M+t and extMt\operatorname{ext}M-t are subsets of AA. Hence the spanning condition span{nH}H(M)=d\operatorname{span}\{n_{H}\}_{H\in\mathcal{F}(M)}=\mathbb{R}^{d} is violated. Conversely, if the spanning condition is violated, there is a direction td{0}t\in\mathbb{R}^{d}\setminus\{0\} such that tt is orthogonal to all the facet normals in (M)\mathcal{F}(M). As in the previous paragraph, extM+t\operatorname{ext}M+t and extMt\operatorname{ext}M-t will still satisfy the facet-defining inequalities of AA for t||t|| sufficiently small, i.e. extM+t,extMtA\operatorname{ext}M+t,\operatorname{ext}M-t\subset A.

The statement of the of Theorem 5.3 is the contraposition of what we have shown. ∎

D.3. Proofs for Section 6

We use Theorem C.1 in Appendix C and the notation introduced for this result in the following proof.

Proof of Theorem 6.1.

Let MextM\in\operatorname{ext}\mathcal{M} be the extended menu associated with a mechanism xext𝒳x\in\operatorname{ext}\mathcal{X}. Recall that menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M by Corollary A.3 (since d=2d=2); thus, we show |extM||||\operatorname{ext}M|\leq|\mathcal{F}|.

If coneΘ2\operatorname{cone}\Theta\neq\mathbb{R}^{2}, then V(M)V(M)\neq\emptyset for otherwise MextM\in\operatorname{ext}\mathcal{M} has a flexible chain. If coneΘ=2\operatorname{cone}\Theta=\mathbb{R}^{2} and V(M)=V(M)=\emptyset, then MM can only not have a flexible chain if extM=B1(M)\operatorname{ext}M=B_{1}(M). In this case, |extM||||\operatorname{ext}M|\leq|\mathcal{F}|.Thus, we assume V(M)V(M)\neq\emptyset going forward.

Consider any vertex vV(M)v\in V(M) such that the sequence of subsequent vertices S=(v1,,vn)S=(v_{1},\ldots,v_{n}) in the clockwise ordering of extM\operatorname{ext}M satisfies SV(M)=S\cap V(M)=\emptyset and such that vnv_{n} is adjacent to a vertex vV(M)v^{\prime}\in V(M). Since v,vextAv,v^{\prime}\in\operatorname{ext}A, let (e1,,ek)(e_{1},\ldots,e_{k}) be the sequence of edges traversed when moving from vv to vv^{\prime} clockwise on the boundary of AA. (If v=vv=v^{\prime}, then all edges are traversed.)

We show that nk1n\leq k-1. Since MextMM\in\operatorname{ext}M, SS does not contain a flexible chain. Thus, |(B2(M)I(M))S|=1|(B_{2}(M)\cup I(M))\cap S|=1. On every edge i=2,k1i=2,\ldots k-1, there lies at most one vertex in extM\operatorname{ext}M, for otherwise |B2(M)|2|B_{2}(M)|\geq 2. Moreover, since vv and vv^{\prime} lie on e1e_{1} and eke_{k}, respectively, there can be at most one vertex in extM{v,v}\operatorname{ext}M\setminus\{v,v^{\prime}\} on e1ene_{1}\cup e_{n}. (This vertex would have to be in B2(M)B_{2}(M)). Thus, nk1n\leq k-1.

By applying the previous argument to every vV(M)v\in V(M), we conclude that |extM||||\operatorname{ext}M|\leq|\mathcal{F}|. ∎

Proof of Theorem 6.2.

Immediate from Theorem 6.6 below. ∎

Proof of Theorem 6.3.

Let x𝒳x\in\mathcal{X} be exhaustive and such that menu(x)\operatorname{menu}(x) is finite and in general position. Let MM\in\mathcal{M} be the associated extended menu. By Corollary A.3, extM=menu(x)\operatorname{ext}M=\operatorname{menu}(x). M=convextM+ΘM=\operatorname{conv}\operatorname{ext}M+\Theta^{\circ} is a polyhedron because extM\operatorname{ext}M is finite and Θ\Theta^{\circ} is a polyhedral cone. Since extM\operatorname{ext}M is in general position, all proper bounded faces of MM are simplices. [118, Theorem 5.1 ] shows that a polyhedron MM of which every bounded face is a simplex cannot be represented as a convex combination of polyhedra with the same recession cone as MM that are not homothetic to MM. Therefore, MM has no non-homothetic decomposition. By Lemma D.2, MM has no homothetic decomposition because MM is exhaustive. Thus MextM\in\operatorname{ext}\mathcal{M}, and xext𝒳x\in\operatorname{ext}\mathcal{X} by Theorem A.2. ∎

We may define exhaustiveness for arbitrary subsets SS of AA: (S)\mathcal{F}(S)\subseteq\mathcal{F} are the facets of AA intersected by SS, and SS is exhaustive if there is no SAS^{\prime}\subset A positively homothetic to SS such that (S)(S)\mathcal{F}(S)\subseteq\mathcal{F}(S^{\prime}). Theorem 5.3 applies as before. Recall that an extended menu MM\in\mathcal{M} is exhaustive if extM\operatorname{ext}M is exhaustive.

We use the following simple consequence of Theorem 5.3 in the proof of Theorem 6.4.

Corollary D.3.

If SAS\subset A is exhaustive, then there exists an exhaustive SSS^{\prime}\subset S such that |(S)|d+1|\mathcal{F}(S^{\prime})|\leq d+1.

Proof.

By Theorem 5.3, span{nH}H(S)=d\operatorname{span}\{n_{H}\}_{H\in\mathcal{F}(S)}=\mathbb{R}^{d}. Thus, there exists a subset (S)\mathcal{F}^{\prime}\subset\mathcal{F}(S) with ||=d|\mathcal{F}^{\prime}|=d such that span{nH}H=d\operatorname{span}\{n_{H}\}_{H\in\mathcal{F}^{\prime}}=\mathbb{R}^{d}. Moreover, by Theorem 5.3, there must exist a hyperplane H(S)H^{\prime}\in\mathcal{F}(S)\setminus\mathcal{F}^{\prime} such that H(S)HH\bigcap_{H\in\mathcal{F}(S)}H\cap H^{\prime}\neq\emptyset. Select SSS^{\prime}\subset S such that (S)={F}\mathcal{F}(S^{\prime})=\mathcal{F}^{\prime}\cup\{F^{\prime}\}. (Clearly, at most d+1d+1 points in SS suffice.) Theorem 5.3 completes the proof. ∎

Proof of Theorem 6.4.

Let x𝒳x\in\mathcal{X} be exhaustive with associated extended menu MM\in\mathcal{M} and such that menu(x)=extM\operatorname{menu}(x)=\operatorname{ext}M is finite. We first construct a menu M~ext\tilde{M}\in\operatorname{ext}\mathcal{M} of finite menu size that is arbitrarily close to MM in the Hausdorff distance and satisfies |extM|=|extM~||\operatorname{ext}M|=|\operatorname{ext}\tilde{M}|. This suffices to show the denseness claim in the statement of Theorem 6.4 by Lemma A.7.

Select an inclusion-wise minimal subset VextMV\subseteq\operatorname{ext}M such that VV is exhaustive and a¯V\underaccent{\bar}{a}\in V. By Corollary D.3, |V|d+1|V|\leq d+1. (If a¯V\underaccent{\bar}{a}\in V, then |V|=2|V|=2 suffices.) If |V|d|V|\leq d, then VV is trivially in general position (i.e., no more than dd points lie on any hyperplane in d\mathbb{R}^{d}). Suppose |V|=d+1|V|=d+1. Then every vertex in VV touches exactly one of the d+1d+1 facets in (V)\mathcal{F}(V) by construction of VV. Select an arbitrary vertex vVv\in V and move vv to a nearby point vv^{\prime} in the same facet of AA touched by vv that is not in the affine hull of V{v}V\setminus\{v\} (which meets the facet of AA touched by vv in a (d2)(d-2)-dimensional convex set). Let W=V{v}{v}W=V\setminus\{v\}\cup\{v^{\prime}\}.

Now consider one-by-one vextMVv\in\operatorname{ext}M\setminus V. Perturb vv to a point vAv^{\prime}\in A arbitrarily close to vv such that vv does not lie in any hyperplane spanned by any subset of dd points in WW. (This is possible since there are only finitely many such hyperplanes.) Update W=W{v}W=W\cup\{v^{\prime}\} and V=V{v}V=V\cup\{v\}. Proceed iteratively until V=extMV=\operatorname{ext}M. The resulting set of points WW is in general position and exhaustive by construction.

Define M~=convW+Θ\tilde{M}=\operatorname{conv}W+\Theta^{\circ}. By construction, M~\tilde{M} is a polyhedron in \mathcal{M}. As long as all of the finitely many perturbations carried out are sufficiently small, WW is in convex position, i.e., no point in WW is in the convex hull of the other points, because extM\operatorname{ext}M was in convex position. Moreover, for all v,vWv,v^{\prime}\in W, vv+Θv\notin v^{\prime}+\Theta^{\circ} because Θ\Theta^{\circ} is closed and the same holds for all v,vextMv,v^{\prime}\in\operatorname{ext}M. Thus, extM~=W\operatorname{ext}\tilde{M}=W and M~\tilde{M} is exhaustive because WW is exhaustive.

M~ext\tilde{M}\in\operatorname{ext}\mathcal{M} by Theorem 6.3 and |extM~|=|W|=|extM||\operatorname{ext}\tilde{M}|=|W|=|\operatorname{ext}M| by construction, proving denseness.

For openness, every polytope in a sufficiently small Hausdorff-ball around a simplicial polytope convextM\operatorname{conv}\operatorname{ext}M is simplicial since the vertices remain in general position (see e.g. [49, Theorems 5.3.1 and 10.1.1]). By Lemmas A.6 and A.7, the claim follows. ∎

Remark.

An alternative statement of Theorem 6.4 is that the set of extreme points of menu size kk is relatively open and dense in the set of exhaustive mechanisms of menu size k\leq k. This is because the set of exhaustive extended menus of menu size kk is relatively open and dense in the set of exhaustive extended menus of menu size k\leq k.

Proof of Corollary 6.5.

Take an arbitrary exhaustive extended menu MM\in\mathcal{M}. Select a finite set of vertices VextMV\subseteq\operatorname{ext}M, including a¯\underaccent{\bar}{a} (if a¯\underaccent{\bar}{a} exists) as well as points on the same facets of AA as extM\operatorname{ext}M, such that for every point of extM\operatorname{ext}M there is a selected point in VV at most ε>0\varepsilon>0 away. By construction, VV is an exhaustive set and a¯V\underaccent{\bar}{a}\in V. Thus, M~=convV+Θ\tilde{M}=\operatorname{conv}V+\Theta^{\circ}\in\mathcal{M} is exhaustive, has finite menu size, and is arbitrarily close to MM for ε\varepsilon sufficiently small. By Theorem 6.4, M~\tilde{M} is arbitrarily close to an element of extM\operatorname{ext}M with finite menu size, which completes the proof. ∎

The proof of Theorem 6.6 proceeds with Baire-category type arguments, for which we need a few definitions:

  • exh\operatorname{exh}\mathcal{M}\subset\mathcal{M} is the set of exhaustive extended menus;

  • 𝒜kexh\mathcal{A}_{k}\subset\operatorname{exh}\mathcal{M} is the set of exhaustive extended menus MM such that M=12M+12M′′M=\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime} for M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M} with d(M,M′′)1kd(M^{\prime},M^{\prime\prime})\geq\frac{1}{k};

  • kexh\mathcal{B}_{k}\subset\operatorname{exh}\mathcal{M} is the set of exhaustive extended menus MM that have a bounded face ff with diam(f)1k\operatorname{diam}(f)\geq\frac{1}{k} and outer unit normal vector nfΘn_{f}\in\Theta such that d(nf,bndrconeΘ)1/kd(n_{f},\operatorname{bndr}\operatorname{cone}\Theta)\geq 1/k (which is satisfied by convention if bndrconeΘ=\operatorname{bndr}\operatorname{cone}\Theta=\emptyset, i.e., coneΘ=d\operatorname{cone}\Theta=\mathbb{R}^{d}).454545The diameter of a set SnS\subseteq\mathbb{R}^{n}, denoted diam(S)\operatorname{diam}(S), is defined as: diam(S)=sup{ab:a,bS}.\operatorname{diam}(S)=\sup\{\|a-b\|:a,b\in S\}.

We note that ext=exhk=1𝒜k\operatorname{ext}\mathcal{M}=\operatorname{exh}\mathcal{M}\setminus\bigcup_{k=1}^{\infty}\mathcal{A}_{k}. Moreover, define exhsc:=exhk=1k\operatorname{exh}\mathcal{M}^{sc}:=\operatorname{exh}\mathcal{M}\setminus\bigcup_{k=1}^{\infty}\mathcal{B}_{k}.

Lemma D.4.

Let x𝒳x\in\mathcal{X} be a mechanism associated with an extended menu MexhscM\in\operatorname{exh}\mathcal{M}^{sc}. Then, x:ΘAx:\Theta\to A is continuous on intconeΘ\operatorname{int}\operatorname{cone}\Theta. In particular, menu(x)\operatorname{menu}(x) is uncountable whenever it is not a singleton.

Proof.

For any MexhscM\in\operatorname{exh}\mathcal{M}^{sc} and θintconeΘ\theta\in\operatorname{int}\operatorname{cone}\Theta, argmaxaMθa\operatorname*{arg\,max}_{a\in M}\theta\cdot a is a singleton for otherwise the boundary of MM would contain a line segment connecting two extreme points of MM. In particular, x(θ)x(\theta) is uniquely determined by MM on intconeΘ\operatorname{int}\operatorname{cone}\Theta. Therefore, the associated indirect utility function UU is differentiable on intconeΘ\operatorname{int}\operatorname{cone}\Theta. By [107, Corollary 25.5.1 ], UU is continuously differentiable and therefore x=Ux=\nabla U is continuous on intconeΘ\operatorname{int}\operatorname{cone}\Theta. ∎

Lemma D.5.

𝒜k\mathcal{A}_{k} and k\mathcal{B}_{k} are closed subsets of exh\operatorname{exh}\mathcal{M} for all kk\in\mathbb{N}.

Proof.

Consider any convergent sequence {Mi}i𝒜k\{M_{i}\}_{i\in\mathbb{N}}\subset\mathcal{A}_{k} with limit MexhM\in\operatorname{exh}\mathcal{M}. We show M𝒜kM\in\mathcal{A}_{k}. Selecting a subsequence, if necessary, we may assume that the associated sequences {Mi}i\{M_{i}^{\prime}\}_{i\in\mathbb{N}}\subset\mathcal{M} and {Mi′′}i\{M_{i}^{\prime\prime}\}_{i\in\mathbb{N}}\subset\mathcal{M}, where Mi=12M+12M′′M_{i}=\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime}, converge in \mathcal{M} by Blaschke’s selection theorem and Lemma A.6. Let MM^{\prime} and M′′M^{\prime\prime} denote the respective limits. We have d(M,M′′)1kd(M^{\prime},M^{\prime\prime})\geq\frac{1}{k} since d(Mi,Mi′′)1kd(M_{i}^{\prime},M_{i}^{\prime\prime})\geq\frac{1}{k} for all ii\in\mathbb{N}. Moreover, M=12M+12M′′M=\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime} since Mi=12Mi+12Mi′′M_{i}=\frac{1}{2}M_{i}^{\prime}+\frac{1}{2}M_{i}^{\prime\prime} for all ii\in\mathbb{N} and \mathcal{M} is convex, so 12M+12M′′\frac{1}{2}M^{\prime}+\frac{1}{2}M^{\prime\prime}\in\mathcal{M}. Thus, M𝒜kM\in\mathcal{A}_{k}.

Consider any convergent sequence {Mi}ik\{M_{i}\}_{i\in\mathbb{N}}\subset\mathcal{B}_{k} with limit MexhM\in\operatorname{exh}\mathcal{M}. We show MkM\in\mathcal{B}_{k}. By definition, for each ii\in\mathbb{N}, there exists a line segment LibndrMiL_{i}\subseteq\operatorname{bndr}M_{i} of length 1k\geq\frac{1}{k} with normal vector niΘn_{i}\in\Theta such that d(ni,bndrconeΘ)1kd(n_{i},\operatorname{bndr}\operatorname{cone}\Theta)\geq\frac{1}{k}. Selecting a subsequence, if necessary, we may assume that the line segments {Li}i\{L_{i}\}_{i\in\mathbb{N}} and the normal vectors {ni}i\{n_{i}\}_{i\in\mathbb{N}} converge to limits LAL^{*}\subset A and nΘn^{*}\in\Theta, respectively, because Θ𝕊d1\Theta\subseteq\mathbb{S}^{d-1} and AA are compact. It is routine to verify that LbndrML^{*}\subseteq\operatorname{bndr}M, LL^{*} has length 1k\geq\frac{1}{k}, nn^{*} is normal to LL^{*} on bndrM\operatorname{bndr}M, and d(n,bndrconeΘ)1kd(n^{*},\operatorname{bndr}\operatorname{cone}\Theta)\geq\frac{1}{k}. Thus, MkM\in\mathcal{B}_{k}. ∎

Proof of Theorem 6.6.

We show that extexhsc\operatorname{ext}\mathcal{M}\cap\operatorname{exh}\mathcal{M}^{sc} is a dense GδG_{\delta} in exh\operatorname{exh}\mathcal{M}. This implies the statement by Lemma D.4.

The proof uses the Baire category theorem. For this, note that exh\operatorname{exh}\mathcal{M} is a compact metric space, hence a Baire space, because exh\operatorname{exh}\mathcal{M} is a closed subset of the compact metric space \mathcal{M} (Lemmas A.6 and A.7). The set exh\operatorname{exh}\mathcal{M} is closed because every extended menu in a sufficiently small neighborhood of a non-exhaustive extended menu MM\in\mathcal{M} intersects a weakly smaller set of facets of AA than MM and is hence also non-exhaustive by Theorem 5.3. Thus, it suffices to show that ext\operatorname{ext}\mathcal{M} and exhsc\operatorname{exh}\mathcal{M}^{sc} are each a dense GδG_{\delta} in exh\operatorname{exh}\mathcal{M}. For ext\operatorname{ext}\mathcal{M}, this follows immediately from Corollary 6.5 and Lemma D.5.

We complete the proof by showing that exhsc\operatorname{exh}\mathcal{M}^{sc} is a dense GδG_{\delta} in exh\operatorname{exh}\mathcal{M}. By Lemma D.5, exhsc\operatorname{exh}\mathcal{M}^{sc} is a GδG_{\delta} in exh\operatorname{exh}\mathcal{M}. To show denseness, consider the set exhk\operatorname{exh}\mathcal{M}\setminus\mathcal{B}_{k} for some arbitrary kk\in\mathbb{N}. By Lemma D.5, exhk\operatorname{exh}\mathcal{M}\setminus\mathcal{B}_{k} is relatively open in exh\operatorname{exh}\mathcal{M}. Moreover, exhk\operatorname{exh}\mathcal{M}\setminus\mathcal{B}_{k} is dense in exh\operatorname{exh}\mathcal{M} because every extended menu MexhM\in\operatorname{exh}\mathcal{M} can be approximated by a polyhedron in exh\operatorname{exh}\mathcal{M} whose bounded faces have diameter <1k<\frac{1}{k}. We have that exhsc=k=1(exhk)\operatorname{exh}\mathcal{M}^{sc}=\bigcap_{k=1}^{\infty}(\operatorname{exh}\mathcal{M}\setminus\mathcal{B}_{k}) is a countable intersection of relatively open and dense sets in a Baire space. Thus, by the Baire category theorem, exhsc\operatorname{exh}\mathcal{M}^{sc} is dense in exh\operatorname{exh}\mathcal{M}. ∎

Proof of Corollary 6.7.

Corollary 6.5 shows that the extreme points of 𝒳\mathcal{X} are dense in the set of exhaustive mechanisms. The Straszewicz-Klee theorem ([66, Theorem 2.1]) implies that the exposed points of 𝒳\mathcal{X} are also dense in the set of exhaustive mechanisms. The Riesz representation theorem ([39, Theorem IV.1]) implies that, for every exposed point xexp𝒳x\in\exp\mathcal{X}, there exists an objective vv and prior μ\mu such that xx is uniquely optimal. ∎

D.4. Proofs for Section 7

Proof of Lemma 7.1.

It remains to show that in the linear delegation problem, the indecomposability of an extended menu MM\in\mathcal{M} is necessary for the non-existence of a non-homothetic decomposition. We show the converse. Assume that there exists an extended menu MM\in\mathcal{M} that is decomposable; that is, there exist convex bodies K,K′′dK^{\prime},K^{\prime\prime}\subset\mathbb{R}^{d}, not homothetic to MM, such that M=K+K′′M=K^{\prime}+K^{\prime\prime}.

We aim to construct from these summands KK^{\prime} and K′′K^{\prime\prime} a non-homothetic decomposition of MM into extended menus. To achieve this, we will identify λ(0,1)\lambda\in(0,1) and tdt\in\mathbb{R}^{d} such that the scaled and translated sets M=1λ(K+t)M^{\prime}=\frac{1}{\lambda}(K^{\prime}+t) and M′′=11λ(K′′t)M^{\prime\prime}=\frac{1}{1-\lambda}(K^{\prime\prime}-t) are extended menus, i.e., subsets of the unit simplex A=ΔA=\Delta. This will complete the proof since M=λM+(1λ)M′′M=\lambda M^{\prime}+(1-\lambda)M^{\prime\prime}.

Since MA=ΔM\subseteq A=\Delta, MM satisfies the following constraints:

  1. (1)

    Positivity: minaMai0\min_{a\in M}a_{i}\geq 0 for all i{1,,d}i\in\{1,\dots,d\};

  2. (2)

    Size: maxaMi=1dai1\max_{a\in M}\sum_{i=1}^{d}a_{i}\leq 1.

We will now define tt and λ\lambda such that the above constraints are binding for MM^{\prime}. This ensures that the constraints are satisfied by M′′M^{\prime\prime} because they are satisfied by MM and MM is a convex combination of MM^{\prime} and M′′M^{\prime\prime}. We set

ti=minaKai.t_{i}=-\min_{a\in K^{\prime}}a_{i}.

This ensures minaK+tai=0\min\limits_{a\in K^{\prime}+t}a_{i}=0 for all i{1,,d}i\in\{1,\dots,d\}; hence MM^{\prime} satisfies the positivity constraint with equality, irrespective of our choice of λ\lambda.

Next, for any convex body KdK\subset\mathbb{R}^{d}, define:

|K|Δ=maxaKi=1daii=1dminaKai.|K|_{\Delta}=\max_{a\in K}\sum_{i=1}^{d}a_{i}-\sum_{i=1}^{d}\min_{a\in K}a_{i}.

Note that ||Δ|\cdot|_{\Delta} commutes with positive scalar multiplication and Minkowski addition; that is, |αK|Δ=α|K|Δ|\alpha K|_{\Delta}=\alpha|K|_{\Delta} and |K1+K2|Δ=|K1|Δ+|K2|Δ|K_{1}+K_{2}|_{\Delta}=|K_{1}|_{\Delta}+|K_{2}|_{\Delta}.

Set

λ=|K|Δ.\lambda=|K^{\prime}|_{\Delta}.

Since KK^{\prime} and K′′K^{\prime\prime} are not singletons (otherwise, the decomposition would be homothetic), we have |K|Δ,|K′′|Δ>0|K^{\prime}|_{\Delta},|K^{\prime\prime}|_{\Delta}>0, hence λ>0\lambda>0. Since |M|Δ1|M|_{\Delta}\leq 1, we have |K|Δ<1|K^{\prime}|_{\Delta}<1 and |K′′|Δ<1|K^{\prime\prime}|_{\Delta}<1, hence λ(0,1)\lambda\in(0,1).

We can now compute

maxaMi=1dai=|M|Δ=1λ|K+t|Δ=1λ|K|Δ=1\max_{a\in M^{\prime}}\sum_{i=1}^{d}a_{i}=|M^{\prime}|_{\Delta}=\frac{1}{\lambda}|K^{\prime}+t|_{\Delta}=\frac{1}{\lambda}|K^{\prime}|_{\Delta}=1

and hence MM^{\prime} satisfies the size constraint with equality. Hence, M′′AM^{\prime\prime}\subseteq A by the earlier argument, which completes the proof. ∎

D.5. Proofs for Section 8

D.5.1. Undominated Mechanisms

We begin by establishing an important result for the proofs of Theorems 8.5 and 8.7, namely that uniquely optimal mechanisms are dense in the undominated extreme points when considering two mechanisms as being “close” when they are “close” with respect to the induced principal’s utility functions. Theorem 8.2 will be proved along the way.

To state the result, let

𝒱={θx(θ)v(θ)xX}\mathcal{V}=\{\theta\mapsto x(\theta)\cdot v(\theta)\mid x\in X\}

denote the set of the principal’s utility functions induced by the set of (IC) and (IR) mechanisms. This set of functions Θ\Theta\to\mathbb{R} is convex and L1L^{1}-compact because it is a continuous image of the compact convex set 𝒳\mathcal{X}.

We say that a principal utility function V𝒱V\in\mathcal{V} is undominated if there exists an undominated mechanism x𝒳x\in\mathcal{X} such that V(θ)=x(θ)v(θ)V(\theta)=x(\theta)\cdot v(\theta).

We also define the following subsets of 𝒱\mathcal{V}:

  • und𝒱𝒱\operatorname{und}\mathcal{V}\subset\mathcal{V} is the set of undominated principal utility functions;

  • und¯𝒱und𝒱\underline{\operatorname{und}}\mathcal{V}\subset\operatorname{und}\mathcal{V} is the set of undominated principal utility functions that are strictly suboptimal for every probability density fL(Θ)f\in L^{\infty}(\Theta) that is uniformly bounded away from 0;

  • exp+𝒱𝒱\exp_{+}\mathcal{V}\subset\mathcal{V} is the set of principal utility functions that are uniquely optimal for some probability density fL(Θ)f\in L^{\infty}(\Theta) that is uniformly bounded away from 0. Note exp+𝒱und𝒱ext𝒱\exp_{+}\mathcal{V}\subset\operatorname{und}\mathcal{V}\cap\operatorname{ext}\mathcal{V}.

As usual, we write V,f=ΘV(θ)f(θ)𝑑θ\langle V,f\rangle=\int_{\Theta}V(\theta)f(\theta)\,d\theta.

Proposition D.6.

exp+𝒱\exp_{+}\mathcal{V} is dense in ext𝒱und𝒱\operatorname{ext}\mathcal{V}\cap\operatorname{und}\mathcal{V}.

We proof the result in three steps. The argument for the first Lemma is inspired by the argument for Theorem 9 in [81]; note the correction in [83].

Lemma D.7.

und¯𝒱clconv(exp+𝒱)\underline{\operatorname{und}}\mathcal{V}\subseteq\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}).

Proof.

Fix any Vund¯𝒱V\in\underline{\operatorname{und}}\mathcal{V}. We show the claim by constructing a convergent sequence of points in 𝒱\mathcal{V} that are convex combinations of points in exp+𝒱\exp_{+}\mathcal{V} with limit VV.

For ε0\varepsilon\geq 0, let

Fε={fL(Θ)εf1}.F_{\varepsilon}=\{f\in L^{\infty}(\Theta)\mid\varepsilon\leq f\leq 1\}.

Up to renormalization, these functions are essentially bounded probability densities that are uniformly bounded away from zero. By the Banach-Alaoglu theorem, FεF_{\varepsilon} is weak*-compact because it is a weak*-closed subset of the dual unit ball.464646Recall that by the Riesz representation theorem, every continuous linear functional on L1(Θ)L^{1}(\Theta) can be represented by a function in L(Θ)L^{\infty}(\Theta).

Recall that Vund¯𝒱V\in\underline{\operatorname{und}}\mathcal{V}, i.e., VV is strictly suboptimal for every density fL(Θ)f\in L^{\infty}(\Theta) that is uniformly bounded away from 0. Thus, for every fFεf\in F_{\varepsilon}, there exists Vf𝒱V_{f}\in\mathcal{V} such that

Vf,f>V,f.\langle V_{f},f\rangle>\langle V,f\rangle.

By the continuity of the evaluation (see e.g. [3, Corollary 6.40]), for every fFεf\in F_{\varepsilon}, there exists a weak*-open neighborhood OfO_{f} of ff such that for all fOff^{\prime}\in O_{f},

Vf,f>V,f.\langle V_{f},f^{\prime}\rangle>\langle V,f^{\prime}\rangle.

Thus, {Of:fFε}\{O_{f}:\>f\in F_{\varepsilon}\} is a weak*-open cover of FεF_{\varepsilon}.

By compactness, the open cover {Of:fFε}\{O_{f}:\>f\in F_{\varepsilon}\} has a finite subcover {Om:m=1,,M}\{O_{m}:\>m=1,\ldots,M\}. The functionals fFεf\in F_{\varepsilon} that expose a point in 𝒱\mathcal{V} are norm-dense in FεF_{\varepsilon} (see e.g. [76] and note that FεF_{\varepsilon} has non-empty interior). Thus, for every m=1,,Mm=1,\ldots,M, there exists fOmf^{\prime}\in O_{m} such that Vm:=Vfexp𝒱V_{m}:=V_{f^{\prime}}\in\exp\mathcal{V}.

Let

G={(V1V,f,,VmV,f)fFε}M.G=\left\{\left(\langle V_{1}-V,f\rangle,\ldots,\langle V_{m}-V,f\rangle\right)\mid f\in F_{\varepsilon}\right\}\subset\mathbb{R}^{M}.

The set GG is

  • convex (because FεF_{\varepsilon} is convex);

  • compact (because it is the continuous image of a weak*-compact set);

  • and satisfies GM=G\cap\mathbb{R}^{M}_{-}=\emptyset (by construction of the open cover {Om:m=1,,M}\{O_{m}:\>m=1,\ldots,M\}), where M\mathbb{R}^{M}_{-} is the negative orthant.

By the Separating Hyperplane Theorem, there exists a vector α+M{0}\alpha\in\mathbb{R}^{M}_{+}\setminus\{0\}, such that αy>0\alpha\cdot y>0 for all yGy\in G. Renormalize i=1Mαi=1\sum_{i=1}^{M}\alpha_{i}=1.

Define

V~ε=i=1MαiVi.\tilde{V}_{\varepsilon}=\sum_{i=1}^{M}\alpha_{i}V_{i}.

Note V~ε𝒱\tilde{V}_{\varepsilon}\in\mathcal{V} since 𝒱\mathcal{V} is convex. For all fFεf\in F_{\varepsilon},

V~ε,fV,f=α(V1V,f,,VmV,f)>0.\langle\tilde{V}_{\varepsilon},f\rangle-\langle V,f\rangle=\alpha\cdot\left(\langle V_{1}-V,f\rangle,\ldots,\langle V_{m}-V,f\rangle\right)>0.

Now consider a sequence εn0\varepsilon_{n}\to 0 and the corresponding sequence of V~εn\tilde{V}_{\varepsilon_{n}} constructed above. Since 𝒱\mathcal{V} is norm-compact, a subsequence of (V~εn)(\tilde{V}_{\varepsilon_{n}}) converges to some V~𝒱\tilde{V}\in\mathcal{V}.

We show V~=V\tilde{V}=V, which proves the claim. Recall that VV is undominated and suppose V~V\tilde{V}\neq V. Then there exists a set Θ~Θ\tilde{\Theta}\subset\Theta of non-zero (spherical) measure such that V(θ)>V~(θ)V(\theta)>\tilde{V}(\theta) for all θΘ~\theta\in\tilde{\Theta}. Thus, any density ff concentrated on Θ~\tilde{\Theta} is such that

V,f>V~,f.\langle V,f\rangle>\langle\tilde{V},f\rangle.

By norm-norm continuity of the evaluation, there exists a strictly positive density ff^{\prime} and some V~εn\tilde{V}_{\varepsilon_{n}} for nn large enough such that

V,f>V~εn,f,\langle V,f\rangle>\langle\tilde{V}_{\varepsilon_{n}},f^{\prime}\rangle,

a contradiction. ∎

Proof of Theorem 8.2.

Follows from the proof for Lemma D.7 with ε=0\varepsilon=0. ∎

We now extend Lemma D.7 to cover all undominated mechanisms.

Lemma D.8.

und𝒱clconv(exp+𝒱)\operatorname{und}\mathcal{V}\subseteq\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}).

Proof.

Suppose not, i.e., Vund𝒱clconv(exp+𝒱)V\in\operatorname{und}\mathcal{V}\setminus\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}). By Lemma D.7, Vund¯𝒱V\notin\underline{\operatorname{und}}\mathcal{V}, i.e., VV is optimal for some density fL(Θ)f^{*}\in L^{\infty}(\Theta) that is uniformly bounded away from 0.

Since clconv(exp+𝒱)\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}) is a closed subset of the norm-compact set 𝒱\mathcal{V}, it is norm-compact. By the Hahn-Banach Separation Theorem, there exists fL(Θ)f\in L^{\infty}(\Theta) such that

V,f>maxVclconv(exp+𝒱)V,f.\langle V,f\rangle>\max_{V^{\prime}\in\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V})}\langle V^{\prime},f\rangle.

f~=εf+(1ε)f\tilde{f}=\varepsilon f+(1-\varepsilon)f^{*} is still uniformly bounded away from 0 for ε(0,1)\varepsilon\in(0,1) small enough and, moreover, for ε(0,1)\varepsilon\in(0,1) small enough

V,f~>maxVclconv(exp+𝒱)V,f~.\langle V,\tilde{f}\rangle>\max_{V^{\prime}\in\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V})}\langle V^{\prime},\tilde{f}\rangle.

by norm-norm continuity of the evaluation and Berge’s maximum theorem (for the RHS). By the result of [76] used in Lemma D.7, there is another density f^\hat{f} arbitrarily close to f~\tilde{f} and therefore also uniformly bounded away from ε\varepsilon that exposes a point V^𝒱\hat{V}\in\mathcal{V}. Again by continuity and Berge’s maximum theorem,

V,f^>maxVclconv(exp+𝒱)V,f^.\langle V,\hat{f}\rangle>\max_{V^{\prime}\in\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V})}\langle V^{\prime},\hat{f}\rangle.

By definition, V^,f^>V,f^\langle\hat{V},\hat{f}\rangle>\langle V,\hat{f}\rangle. Thus, the point V^\hat{V} exposed by f^\hat{f} cannot be in exp+𝒱\exp_{+}\mathcal{V}, a contradiction. ∎

We complete the proof of Proposition D.6.

Proof of Proposition D.6.

The claim is a consequence of Milman’s theorem (see e.g. [66], Theorem 1.1.). The theorem implies that extclconv(exp+𝒱)clexp+𝒱\operatorname{ext}\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V})\subseteq\operatorname{cl}\exp_{+}\mathcal{V} since clconv(exp+𝒱)\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}) is compact and convex. In particular, by Lemma D.8, every undominated extreme point of 𝒱\mathcal{V} must be in clconv(exp+𝒱)\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}). But since clconv(exp+𝒱)\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}) is a convex subset of 𝒱\mathcal{V}, every undominated extreme point of 𝒱\mathcal{V} must also be in extclconv(exp+𝒱)\operatorname{ext}\operatorname{\operatorname{cl}\operatorname{conv}}(\exp_{+}\mathcal{V}) and therefore arbitrarily close to a point in exp+𝒱\exp_{+}\mathcal{V}. ∎

D.5.2. Multi-Good Monopoly

We proceed with the multi-good monopoly problem. To follow the standard terminology in mechanism design with transfers, we abuse language and refer to elements of [0,1]m[0,1]^{m} as types and allocations, and consider mechanisms and indirect utility functions as functions defined on [0,1]m[0,1]^{m}. In line with standard notation, we also write (x,t)𝒳(x,t)\in\mathcal{X} to separate the “allocation component” of a mechanism from the “transfer component.”

We use the following lemma about undominated mechanisms in the upcoming arguments.

Lemma D.9 ([81], Lemma 11).

Suppose (x,t)𝒳(x^{\prime},t^{\prime})\in\mathcal{X} and (x,t)𝒳(x,t)\in\mathcal{X} with indirect utility functions UU^{\prime} and UU, respectively, are such that ttt^{\prime}\geq t almost everywhere. Then, for all θ[0,1]m\theta\in[0,1]^{m} and λθ[0,1]m\lambda\theta\in[0,1]^{m} with λ>1\lambda>1,

  1. (1)

    U(θ)>U(θ)U(λθ)>U(λθ)U^{\prime}(\theta)>U(\theta)\implies U^{\prime}(\lambda\theta)>U(\lambda\theta);

  2. (2)

    U(θ)U(θ)U(λθ)U(λθ)U^{\prime}(\theta)\geq U(\theta)\implies U^{\prime}(\lambda\theta)\geq U(\lambda\theta).

Proof of Lemma 8.3.

Is is without loss of generality to consider only pricing functions with marginal prices in [0,1][0,1] because types are in [0,1]m[0,1]^{m}.

Now consider a pricing function pp with marginal prices in [δ,1δ][\delta,1-\delta] for some δ>0\delta>0. Let (x,t)𝒳(x,t)\in\mathcal{X} be the mechanism obtained from pp, and let UU be the associated indirect utility function. For the sake of contradiction, suppose (x,t)𝒳(x^{\prime},t^{\prime})\in\mathcal{X} dominates (x,t)(x,t), and let UU^{\prime} be the associated indirect utility function.

Since marginal prices are in [δ,1δ][\delta,1-\delta], we have U(θ)U(θ)=0U^{\prime}(\theta)\geq U(\theta)=0 for all θ(δ,,δ)\theta\leq(\delta,\ldots,\delta). By Lemma D.9, we have UUU^{\prime}\geq U. Thus, there is a type θ[0,1]m\theta\in[0,1]^{m} such that U(θ)>U(θ)U^{\prime}(\theta)>U(\theta) (otherwise xx and xx^{\prime} are payoff-equivalent). By the continuity of indirect utility functions, we may assume θ(0,1)m\theta\in(0,1)^{m}.

Let λ1>1\lambda^{1}>1 be the largest scalar such that θ1=λ1θ[0,1δ2]m\theta^{1}=\lambda^{1}\theta\in[0,1-\frac{\delta}{2}]^{m}, and let i=1,,mi=1,\ldots,m be such that θi1=1δ2\theta_{i}^{1}=1-\frac{\delta}{2}. Without loss of generality, suppose i=1i=1. By Lemma D.9, we have U(θ1)>U(θ1)U^{\prime}(\theta^{1})>U(\theta^{1}).

Consider the subspace H1={θ[0,1]mθ1=1δ2}H^{1}=\{\theta\in[0,1]^{m}\mid\theta_{1}=1-\frac{\delta}{2}\}. Up to an arbitrarily small translation of H1H^{1} in coordinate direction ±e1\pm e_{1}, we may assume by Fubini’s Theorem that x(θ)θx(θ)θx^{\prime}(\theta)\cdot\theta\geq x(\theta)\cdot\theta for almost every θH1\theta\in H^{1} (with respect to m1m-1-dimensional Lebesgue measure) because UUU^{\prime}\geq U and ttt^{\prime}\geq t almost everywhere. For all θH1\theta\in H^{1}, we have xi(θ)=1x_{i}(\theta)=1 since θi>1δ\theta_{i}>1-\delta and marginal prices are in [δ,1δ][\delta,1-\delta]. Together with dominance, we have x(θ)(0,θ1)x(θ)(0,θ1)x^{\prime}(\theta)\cdot(0,\theta_{-1})\geq x(\theta)\cdot(0,\theta_{-1}) for almost every θH1\theta\in H^{1}.

Now let λ2>1\lambda^{2}>1 be the largest scalar such that θ2=(θ11,λ2θ11)[0,1δ2]m\theta^{2}=(\theta^{1}_{1},\lambda^{2}\theta_{-1}^{1})\in[0,1-\frac{\delta}{2}]^{m}, and let i=2,,mi=2,\ldots,m be such that θi2=1δ2\theta_{i}^{2}=1-\frac{\delta}{2}. Without loss of generality, suppose i=2i=2. By the same arguments as for the proof of Lemma D.9, we have U(θ2)>U(θ2)U^{\prime}(\theta^{2})>U(\theta^{2}).

Iteratively proceed with this argument, constructing a sequence of affine subspaces (H1,,Hm)(H^{1},\ldots,H^{m}) and types (θ1,,θm)(\theta^{1},\ldots,\theta^{m}), where θki=1δ2\theta^{i}_{k}=1-\frac{\delta}{2} for all i=1,,mi=1,\ldots,m and kik\leq i, such that U(θi)>U(θi)U^{\prime}(\theta^{i})>U(\theta^{i}) for all i=1,,mi=1,\ldots,m and xki=1x_{k}^{i}=1 for all i=1,,mi=1,\ldots,m and kik\leq i.

Finally, U(θm)>U(θ)mU^{\prime}(\theta^{m})>U(\theta)^{m} implies U(θ)=x(θ)θt(θ)>x(θ)θt(θ)=U(θ)U^{\prime}(\theta)=x^{\prime}(\theta)\cdot\theta-t^{\prime}(\theta)>x(\theta)\cdot\theta-t(\theta)=U(\theta) for all θB(θm)\theta\in B(\theta^{m}) by continuity, where B(θm)B(\theta^{m}) is a sufficiently small ball around θm\theta^{m}. We also have x(θ)=(1,,1)x(\theta)=(1,\ldots,1) for all θ>(1δ,,1δ)\theta>(1-\delta,\ldots,1-\delta) since marginal prices are in [δ,1δ][\delta,1-\delta]. Thus, t(θ)<t(θ)t^{\prime}(\theta)<t(\theta) for all θB(θm)\theta\in B(\theta^{m}), a contradiction with dominance. ∎

Proof of Corollary 8.4 .

Take any pricing function pp with marginal prices in [0,1][0,1]. Then, for ε>δ>0\varepsilon>\delta>0 small enough, the pricing function p(a)=(1ε)p(a)+δap^{\prime}(a)=(1-\varepsilon)p(a)+\delta a has marginal prices uniformly bounded away from 0 and 1. Moreover, the epigraphs epip\operatorname{epi}p and epip\operatorname{epi}p^{\prime} of pp and pp^{\prime}, respectively, are arbitrarily close in the Hausdorff distance. Thus, the extended menus M=epip+ΘM=\operatorname{epi}p+\Theta^{\circ} and M=epip+ΘM^{\prime}=\operatorname{epi}p^{\prime}+\Theta^{\circ} are arbitrarily close in the Hausdorff distance (Lemma A.6). By Lemma A.7, the associated mechanisms xx and xx^{\prime}, respectively, are arbitrarily close in L1L^{1}. By Lemma 8.3, xx^{\prime} is undominated. ∎

Proof of Theorem 8.5.

The argument for why the undominated extreme points are dense in the set of (IC) and (IR) mechanisms when m2m\geq 2 is analogous to the arguments in the proofs for Section 6 in Section D.3. By Corollary 8.4, for every (x,t)𝒳(x,t)\in\mathcal{X}, find (x,t)𝒳(x^{\prime},t^{\prime})\in\mathcal{X} arbitrarily close to (x,t)(x,t) with marginal prices bounded away from 0 and 1. Then follow the construction for Corollary 6.5 and then the construction for Theorem 6.4. As long as all perturbations are small enough, the constructed extreme point still has marginal prices bounded away from 0 and 1 and is hence undominated.

We complete the the proof by showing that the mechanisms that are uniquely optimal for some type distribution are dense in the undominated mechanisms.

We first show that if (x,t),(x,t)𝒳(x,t),(x^{\prime},t^{\prime})\in\mathcal{X} are undominated and t=tt=t^{\prime} almost everywhere, then x=xx=x^{\prime} almost everywhere. It is easy to show (e.g., using Euler’s homogenous function theorem) that xxx-x^{\prime} is constant on almost every ray from the origin, i.e., HD0 up to tie-breaking. It therefore suffices to show that for every undominated mechanism (x,t)𝒳(x,t)\in\mathcal{X}, limθ0x(θ)=0\lim_{\theta\to 0}x(\theta)=0 (independently of the choice of sequence). Let pp denote the pricing function associated with xx and assume, for the sake of contradiction, that limθ0x(θ)=a0\lim_{\theta\to 0}x(\theta)=a^{*}\neq 0. Then, p(a)=0p(a)=0 for all aaa\leq a^{*}. Define a mechanism (x,t)𝒳(x^{\prime},t^{\prime})\in\mathcal{X} by letting the agent buy from another pricing schedule

p(a)={p(a+a)if a+a[0,1]m;maxamenu(x)p(a)+εotherwise,p^{\prime}(a)=\begin{cases}p(a+a^{*})&\text{if }a+a^{*}\in[0,1]^{m};\\ \max_{a\in\operatorname{menu}(x)}p(a)+\varepsilon&\text{otherwise},\end{cases}

where ε>0\varepsilon>0. By construction, p(0)=0p^{\prime}(0)=0; thus, (x,t)(x^{\prime},t^{\prime}) is IC and IR. Since pp^{\prime} is obtained from pp by translation of the graph of pp in direction a-a^{*} with a new price maxa[0,1]mp(a)+ε\max_{a\in[0,1]^{m}}p(a)+\varepsilon for the grand bundle a=1a=1, almost every type either buys the same allocation as under pp translated by a-a^{*} or the grand bundle. Thus, ttt^{\prime}\geq t almost everywhere. For all sufficiently small ε>0\varepsilon>0, a positive measure of types will buy the grand bundle. Thus, (x,t)(x^{\prime},t^{\prime}) dominates (x,t)(x,t), a contradiction.

We next claim that ext𝒱und𝒱\operatorname{ext}\mathcal{V}\cap\operatorname{und}\mathcal{V} is dense in und𝒱\operatorname{und}\mathcal{V}, where 𝒱\mathcal{V} is the set of IC transfer functions. If (x,t)ext𝒳(x,t)\in\operatorname{ext}\mathcal{X} is undominated, then text𝒱t\in\operatorname{ext}\mathcal{V}. To see this, suppose t=12t+12t′′ext𝒱t=\frac{1}{2}t^{\prime}+\frac{1}{2}t^{\prime\prime}\notin\operatorname{ext}\mathcal{V} for t,t′′𝒱t^{\prime},t^{\prime\prime}\in\mathcal{V}. Define x~=12x+12x′′𝒳\tilde{x}=\frac{1}{2}x^{\prime}+\frac{1}{2}x^{\prime\prime}\in\mathcal{X}, where x,x′′𝒳x^{\prime},x^{\prime\prime}\in\mathcal{X} induce transfer tt^{\prime} and t′′t^{\prime\prime}, respectively. By definition, xx and x~\tilde{x} both induce tt. Thus, x=x~x=\tilde{x} by the previous paragraph, so xext𝒳x\notin\operatorname{ext}\mathcal{X}. The claim now follows because the undominated mechanisms in ext𝒳\operatorname{ext}\mathcal{X} are dense in the undominated mechanisms in 𝒳\mathcal{X}.

Fix any undominated mechanism (x,t)𝒳(x,t)\in\mathcal{X}. By Proposition D.6 and the previous paragraph, there exists a sequence of transfer functions (tn)nexp+𝒱(t_{n})_{n\in\mathbb{N}}\subset\exp_{+}\mathcal{V}, each uniquely optimal for some type distribution μ\mu, converging to tt in L1L^{1}. We have shown above that the associated sequence of allocation rules (xn)n(x_{n})_{n\in\mathbb{N}} is uniquely determined. Since 𝒳\mathcal{X} is compact, up to taking a subsequence, (xn,tn)n(x_{n},t_{n})_{n\in\mathbb{N}} converges in L1L^{1} to some (x,t)𝒳(x^{\prime},t)\in\mathcal{X}. But (x,t)(x,t) is undominated, hence x=xx=x^{\prime}, as desired. ∎

D.5.3. Linear Veto Bargaining

We proceed with the linear veto bargaining problem.

Proof of Lemma 8.6.

We first show that the conditions given in the statement are necessary. For this, fix any mechanism x𝒳x\in\mathcal{X}. It is clear that a¯menu(x)\underaccent{\bar}{a}\in\operatorname{menu}(x) for otherwise xx does not satisfy (IR) since there is a type θ¯\underaccent{\bar}{\theta} for which a¯\underaccent{\bar}{a} is their (unique) most preferred alternative in AA. Next suppose menu(x)\operatorname{menu}(x) does not contain the principal’s (unique) favorite alternative aextAa^{*}\in\operatorname{ext}A. Obtain a new mechanism xXx^{\prime}\in X by letting the agent choose from menu(x){a}\operatorname{menu}(x)\cup\{a^{*}\}. Thus, for all θΘ\theta\in\Theta, either x(θ)=x(θ)x^{\prime}(\theta)=x(\theta) or x(θ)=ax^{\prime}(\theta)=a^{*}. Since aextAa^{*}\in\operatorname{ext}A, there is a positive measure of types for which aa^{*} is their most preferred allocation in AA. Thus, xx^{\prime} dominates xx.

For sufficiency, let aextAa^{*}\in\operatorname{ext}A be the principal’s favorite alternative and suppose x,x𝒳x,x^{\prime}\in\mathcal{X} are such that amenu(x),menu(x)a^{*}\in\operatorname{menu}(x),\operatorname{menu}(x^{\prime}) and x(θ)v¯x(θ)v¯x(\theta)\cdot\bar{v}\geq x^{\prime}(\theta)\cdot\bar{v} for almost all θΘ\theta\in\Theta. We show that x=xx=x^{\prime} almost everywhere. We extend both mechanisms to coneΘ=d\operatorname{cone}\Theta=\mathbb{R}^{d} by letting each type chose their favorite allocation in menu(x)\operatorname{menu}(x) and menu(x)\operatorname{menu}(x^{\prime}) (xx and xx^{\prime} are constant along almost every ray from the origin). Let UU and UU^{\prime} be the agent’s indirect utility functions associated with xx and xx^{\prime}, respectively.

We claim that for all θconeΘ\theta\in\operatorname{cone}\Theta and λ\lambda\in\mathbb{R},

ϕθ(λ):=U(θ+λv¯)=U(θ)+0λx(θ+zv¯),v¯𝑑z\phi_{\theta}(\lambda):=U(\theta+\lambda\bar{v})=U(\theta)+\int_{0}^{\lambda}\langle x(\theta+z\bar{v}),\bar{v}\rangle\,dz

and analogously for ϕθ\phi_{\theta}^{\prime}, xx^{\prime}, and UU^{\prime}. Recall that x(θ)=U(θ)x(\theta)=\partial U(\theta) and x(θ)=U(θ)x^{\prime}(\theta)=\partial U^{\prime}(\theta) (Theorem A.2). ϕθ(λ)\phi_{\theta}(\lambda) is the restriction of a continuous convex function to a line, hence continuous and convex. It is easy to verify that x(θ+λv¯),v¯\langle x(\theta+\lambda\bar{v}),\bar{v}\rangle as a function of λ\lambda is a subgradient of ϕθ(λ)\phi_{\theta}(\lambda). Hence the envelope formula follows ([107, Theorem 24.2]).

By Fubini’s theorem, ϕθ(λ)ϕθ(λ)\phi_{\theta}(\lambda)-\phi_{\theta}^{\prime}(\lambda) is non-decreasing for almost all θconeΘ\theta\in\operatorname{cone}\Theta because x(θ+λv¯)v¯x(θ+λv¯)v¯x(\theta+\lambda\bar{v})\cdot\bar{v}\geq x^{\prime}(\theta+\lambda\bar{v})\cdot\bar{v} for almost all θconeΘ\theta\in\operatorname{cone}\Theta and all λ\lambda\in\mathbb{R}.

For all sufficiently large λ>0\lambda>0, we have x(θ+λv¯)=x(θ+λv¯)=ax(\theta+\lambda\bar{v})=x^{\prime}(\theta+\lambda\bar{v})=a^{*} since aextAa^{*}\in\operatorname{ext}A is the principal’s, i.e., type v¯\bar{v}’s, (unique) favorite alternative in AA and thus the favorite alternative of type θ+λv¯\theta+\lambda\bar{v}. Thus, ϕθ(λ)ϕθ(λ)=0\phi_{\theta}(\lambda)-\phi_{\theta}^{\prime}(\lambda)=0 for all sufficiently large λ>0\lambda>0.

Similarly, for all sufficiently small λ<0\lambda<0, we have x(θ+λv¯)=x(θ+λv¯)=a¯x(\theta+\lambda\bar{v})=x^{\prime}(\theta+\lambda\bar{v})=\underaccent{\bar}{a} since a¯\underaccent{\bar}{a} is, by assumption, the principal’s (unique) least preferred alternative and the principal’s and agent’s preferences are sufficiently aligned.

Thus, for almost every θconeΘ\theta\in\operatorname{cone}\Theta and every λ\lambda\in\mathbb{R}, we have ϕθ(λ)=ϕθ(λ)\phi_{\theta}(\lambda)=\phi^{\prime}_{\theta}(\lambda) since ϕθ(λ)ϕθ(λ)\phi_{\theta}(\lambda)-\phi_{\theta}^{\prime}(\lambda) is non-decreasing. Put differently, U=UU=U^{\prime} almost everywhere. By continuity, U=UU=U^{\prime}. Consequently, x=xx=x^{\prime} almost everywhere. ∎

Proof of Theorem 8.7.

The argument for statement (1) is immediate from Theorem 7.2 and Lemma 8.6. We proceed with statement (2).

The argument for why the undominated extreme points are dense in the undominated mechanisms when m4m\geq 4 is completely analogous to the proofs of Corollary 6.5 and Theorem 6.4 when making sure that a¯,aV\underaccent{\bar}{a},a^{*}\in V, where aa^{*} is the principal’s favorite alternative and VV is the set of vertices constructed in the proof of Theorem 6.4.

We complete the the proof by showing that the mechanisms that are uniquely optimal for some type distribution are dense in the undominated mechanisms.

The proof of Lemma 8.6 shows that if x,x𝒳x,x^{\prime}\in\mathcal{X} are undominated and such that x(θ)v¯=x(θ)v¯x(\theta)\cdot\bar{v}=x(\theta)\cdot\bar{v} for almost every θΘ\theta\in\Theta, then x=xx=x^{\prime} almost everywhere. Thus, an undominated principal utility function uniquely determines an undominated mechanism.

We claim that ext𝒱und𝒱\operatorname{ext}\mathcal{V}\cap\operatorname{und}\mathcal{V} is dense in und𝒱\operatorname{und}\mathcal{V}. If xext𝒳x\in\operatorname{ext}\mathcal{X} is undominated, then the induced principal utility function V𝒱V\in\mathcal{V} is in ext𝒱\operatorname{ext}\mathcal{V}. To see this, suppose V=12V+12V′′ext𝒱V=\frac{1}{2}V^{\prime}+\frac{1}{2}V^{\prime\prime}\notin\operatorname{ext}\mathcal{V} for V,V′′𝒱V^{\prime},V^{\prime\prime}\in\mathcal{V}. Define x~=12x+12x′′𝒳\tilde{x}=\frac{1}{2}x^{\prime}+\frac{1}{2}x^{\prime\prime}\in\mathcal{X}, where x,x′′𝒳x^{\prime},x^{\prime\prime}\in\mathcal{X} induce VV^{\prime} and V′′V^{\prime\prime}, respectively. By definition, xx and x~\tilde{x} both induce vv. Thus, x=x~x=\tilde{x} by the previous paragraph, so xext𝒳x\notin\operatorname{ext}\mathcal{X}. The claim now follows because the undominated mechanisms in ext𝒳\operatorname{ext}\mathcal{X} are dense in the undominated mechanisms in 𝒳\mathcal{X}.

Fix any undominated mechanism x𝒳x\in\mathcal{X}. Let V𝒱V\in\mathcal{V} be the associated principal utility function. By Proposition D.6 and the previous paragraph, there is a sequence (Vn)nexp+𝒱(V_{n})_{n\in\mathbb{N}}\subset\exp_{+}\mathcal{V} of uniquely optimal undominated principal utility functions converging to VV. Let (xn)next𝒳(x_{n})_{n\in\mathbb{N}}\subset\operatorname{ext}\mathcal{X} be the sequence of mechanisms that is uniquely determined by (Vn)n(V_{n})_{n\in\mathbb{N}}. (By definition, each mechanism in the sequence is uniquely optimal for some type distribution.) By compactness of 𝒳\mathcal{X}, up to taking a subsequence, xnx𝒳x_{n}\to x^{\prime}\in\mathcal{X}. By continuity of the map that assigns to each mechanism in 𝒳\mathcal{X} a principal utility function in 𝒱\mathcal{V}, xx^{\prime} must induce VV since VnVV_{n}\to V. Therefore, xx^{\prime} is undominated. Thus, x=xx=x^{\prime} almost everywhere, which completes the proof. ∎

References

  • [1] William James Adams and Janet L Yellen “Commodity Bundling and the Burden of Monopoly” In The Quarterly Journal of Economics 90.3 MIT Press, 1976, pp. 475–498
  • [2] S Nageeb Ali, Navin Kartik and Andreas Kleiner “Sequential veto bargaining with incomplete information” In Econometrica 91.4 Wiley Online Library, 2023, pp. 1527–1562
  • [3] Charalambos D Aliprantis and Kim C Border “Infinite Dimensional Analysis: A Hitchhiker’s Guide” Springer Science & Business Media, 2007
  • [4] Ricardo Alonso and Niko Matouschek “Optimal delegation” In The Review of Economic Studies 75.1 Wiley-Blackwell, 2008, pp. 259–293
  • [5] Manuel Amador and Kyle Bagwell “The theory of optimal delegation with an application to tariff caps” In Econometrica 81.4 Wiley Online Library, 2013, pp. 1541–1599
  • [6] Manuel Amador and Kyle Bagwell “Regulating a monopolist with uncertain costs without transfers” In Theoretical Economics 17.4 Wiley Online Library, 2022, pp. 1719–1760
  • [7] Itai Arieli, Yakov Babichenko, Rann Smorodinsky and Takuro Yamashita “Optimal persuasion via bi-pooling” In Theoretical Economics 18.1 Wiley Online Library, 2023, pp. 15–36
  • [8] Mark Armstrong “Multiproduct nonlinear pricing” In Econometrica: Journal of the Econometric Society JSTOR, 1996, pp. 51–75
  • [9] Mark Armstrong “Price discrimination by a many-product firm” In The Review of Economic Studies 66.1 Wiley-Blackwell, 1999, pp. 151–168
  • [10] Mark Armstrong and John Vickers “A model of delegated project choice” In Econometrica 78.1 Wiley Online Library, 2010, pp. 213–244
  • [11] Moshe Babaioff, Yannai A Gonczarowski and Noam Nisan “The menu-size complexity of revenue approximation” In Proceedings of the 49th Annual ACM SIGACT Symposium on Theory of Computing, 2017, pp. 869–877
  • [12] Moshe Babaioff, Nicole Immorlica, Brendan Lucier and S Matthew Weinberg “A simple and approximately optimal mechanism for an additive buyer” In Journal of the ACM (JACM) 67.4 ACM New York, NY, USA, 2020, pp. 1–40
  • [13] Moshe Babaioff, Noam Nisan and Aviad Rubinstein “Optimal Deterministic Mechanisms for an Additive Buyer” In arXiv preprint arXiv:1804.06867, 2018 URL: https://arxiv.org/abs/1804.06867
  • [14] Yannis Bakos and Erik Brynjolfsson “Bundling information goods: Pricing, profits, and efficiency” In Management Science 45.12 INFORMS, 1999, pp. 1613–1630
  • [15] Elizabeth Baldwin and Paul Klemperer “Understanding Preferences: “Demand Types”, and the Existence of Equilibrium with Indivisibilities” In Econometrica 87.3 Wiley Online Library, 2019, pp. 867–932
  • [16] Nicholas C. Bedard and Jacob K. Goeree “Tropical Analysis: With an Application to Indivisible Goods” In arXiv preprint arXiv:2308.04593, 2023 URL: https://arxiv.org/abs/2308.04593
  • [17] Ran Ben-Moshe, Sergiu Hart and Noam Nisan “Monotonic Mechanisms for Selling Multiple Goods” In arXiv preprint arXiv:2210.17150, 2022
  • [18] Elchanan Ben-Porath, Eddie Dekel and Barton L Lipman “Optimal Allocation with Costly Verification” In American Economic Review 104.12 American Economic Association, 2014, pp. 3779–3813
  • [19] Dirk Bergemann, Alessandro Bonatti, Andreas Haupt and Alex Smolin “The optimality of upgrade pricing” In International Conference on Web and Internet Economics, 2021, pp. 41–58 Springer
  • [20] Sushil Bikhchandani, Shurojit Chatterji, Ron Lavi, Ahuva Mu’alem, Noam Nisan and Arunava Sen “Weak monotonicity characterizes deterministic dominant-strategy implementation” In Econometrica 74.4 Wiley Online Library, 2006, pp. 1109–1132
  • [21] Sushil Bikhchandani and Debasis Mishra “Selling two identical objects” In Journal of Economic Theory 200 Elsevier, 2022, pp. 105397
  • [22] Wilhelm Blaschke and Georg Pick “Distanzschätzungen im Funktionenraum II” In Mathematische Annalen 77.2 Springer, 1916, pp. 277–300
  • [23] Kim C. Border “Implementation of reduced form auctions: A geometric approach” In Econometrica: Journal of the Econometric Society JSTOR, 1991, pp. 1175–1187
  • [24] Tilman Börgers “An Introduction to the Theory of Mechanism Design” Oxford University Press, USA, 2015
  • [25] Tilman Börgers and Peter Postl “Efficient compromising” In Journal of Economic Theory 144.5 Elsevier, 2009, pp. 2057–2076
  • [26] Efim Mikhailovich Bronshtein “Extremal convex functions” In Siberian Mathematical Journal 19.1 Springer, 1978, pp. 6–12
  • [27] Yang Cai, Nikhil R Devanur and S Matthew Weinberg “A duality-based unified approach to bayesian mechanism design” In SIAM Journal on Computing 50.3 SIAM, 2019, pp. STOC16–160
  • [28] Gabriel Carroll “Robustness and separation in multidimensional screening” In Econometrica 85.2 Wiley Online Library, 2017, pp. 453–488
  • [29] Federico Castillo and Fu Liu “Deformation cones of nested braid fans” In International Mathematics Research Notices 2022.3 Oxford University Press, 2022, pp. 1973–2026
  • [30] Indranil Chakraborty “Bundling decisions for selling multiple objects” In Economic Theory 13.3 Springer, 1999, pp. 723–733
  • [31] Yeon-Koo Che, Wouter Dessein and Navin Kartik “Pandering to persuade” In American Economic Review 103.1, 2013, pp. 47–79
  • [32] Yeon-Koo Che, Jinwoo Kim and Konrad Mierendorff “Generalized Reduced-Form Auctions: A Network-Flow Approach” In Econometrica 81.6 Wiley Online Library, 2013, pp. 2487–2520
  • [33] Yeon-Koo Che and Weijie Zhong “Robustly Optimal Mechanisms for Selling Multiple Goods” In arXiv preprint, 2023
  • [34] Yi-Chun Chen, Wei He, Jiangtao Li and Yeneng Sun “Equivalence of stochastic and deterministic mechanisms” In Econometrica 87.4 Wiley Online Library, 2019, pp. 1367–1390
  • [35] Constantinos Daskalakis, Alan Deckelbaum and Christos Tzamos “The complexity of optimal mechanism design” In Proceedings of the twenty-fifth annual ACM-SIAM symposium on Discrete algorithms, 2014, pp. 1302–1318 SIAM
  • [36] Constantinos Daskalakis, Alan Deckelbaum and Christos Tzamos “Strong Duality for a Multiple-Good Monopolist” In Econometrica 85 Econometric Society, 2017, pp. 735–767
  • [37] Rahul Deb and Anne-Katrin Roesler “Multi-Dimensional Screening: Buyer-Optimal Learning and Informational Robustness” In The Review of Economic Studies Oxford University Press, 2023, pp. rdad100
  • [38] Eddie Dekel, Barton L Lipman and Aldo Rustichini “Representing preferences with a unique subjective state space” In Econometrica 69.4 Wiley Online Library, 2001, pp. 891–934
  • [39] J. Diestel and J… Uhl “Vector Measures”, Math. Surveys 15 Providence, R.I.: American Mathematical Society, 1977
  • [40] Laura Doval, Ran Eilat, Tianhao Liu and Yangfan Zhou “Revealed Information” In arXiv preprint arXiv:2411.13293, 2024
  • [41] Hanming Fang and Peter Norman “To bundle or not to bundle” In The RAND Journal of Economics 37.4 Wiley Online Library, 2006, pp. 946–963
  • [42] Alex Frankel “Delegating multiple decisions” In American Economic Journal: Microeconomics 8.4, 2016, pp. 16–53
  • [43] Alexander Frankel “Aligned delegation” In American Economic Review 104.1, 2014, pp. 66–83
  • [44] Mira Frick, Ryota Iijima and Yuhta Ishii “Multidimensional Screening with Rich Consumer Data”, 2024
  • [45] Rafael M Frongillo and Ian A Kash “General truthfulness characterizations via convex analysis” In Games and Economic Behavior 130 Elsevier, 2021, pp. 636–662
  • [46] David Gale “Irreducible convex sets” In Proceedings of the International Congress of Mathematicians, Amsterdam 2, 1954, pp. 217–218
  • [47] Soheil Ghili “A Characterization for Optimal Bundling of Products with Nonadditive Values” In American Economic Review: Insights 5.3 American Economic Association, 2023, pp. 311–326
  • [48] Parikshit Gopalan, Noam Nisan and Tim Roughgarden “Public projects, Boolean functions, and the borders of Border’s theorem” In ACM Transactions on Economics and Computation (TEAC) 6.3-4 ACM New York, NY, USA, 2018, pp. 1–21
  • [49] Branko Grünbaum, Victor Klee, Micha A Perles and Geoffrey Colin Shephard “Convex polytopes” Springer, 1967
  • [50] Ryszard Grzaślewicz “Extreme convex sets in 2\mathbb{R}^{2} In Archiv der Mathematik 43 Springer, 1984, pp. 377–380
  • [51] Faruk Gul and Wolfgang Pesendorfer “Random expected utility” In Econometrica 74.1 Wiley Online Library, 2006, pp. 121–146
  • [52] Yingni Guo and Eran Shmaya “Regret-Minimizing Project Choice” In Econometrica 91.5 Wiley Online Library, 2023, pp. 1567–1593
  • [53] Nima Haghpanah and Jason Hartline “When is pure bundling optimal?” In The Review of Economic Studies 88.3 Oxford University Press, 2021, pp. 1127–1156
  • [54] Sergiu Hart and Noam Nisan “Approximate revenue maximization with multiple items” In Journal of Economic Theory 172 Elsevier, 2017, pp. 313–347
  • [55] Sergiu Hart and Noam Nisan “Selling multiple correlated goods: Revenue maximization and menu-size complexity” In Journal of Economic Theory 183 Elsevier, 2019, pp. 991–1029
  • [56] Sergiu Hart and Philip J Reny “Maximal revenue with multiple goods: Nonmonotonicity and other observations” In Theoretical Economics 10.3 Wiley Online Library, 2015, pp. 893–922
  • [57] Sergiu Hart and Philip J Reny “The better half of selling separately” In ACM Transactions on Economics and Computation (TEAC) 7.4 ACM New York, NY, USA, 2019, pp. 1–18
  • [58] Jean-Baptiste Hiriart-Urruty and Claude Lemaréchal “Convex Analysis and Minimization Algorithms I: Fundamentals” Springer Science & Business Media, 1996
  • [59] Bengt Holmström “On Incentives and Control in Organizations” Stanford University, 1977
  • [60] Bengt Holmström “On the theory of delegation” In Bayesian Models in Economic Theory New York: North-Holland, 1984
  • [61] Philippe Jehiel, Moritz Meyer-Ter-Vehn and Benny Moldovanu “Mixed bundling auctions” In Journal of Economic Theory 134.1 Elsevier, 2007, pp. 494–512
  • [62] Philippe Jehiel, Benny Moldovanu and Ennio Stacchetti “Multidimensional mechanism design for auctions with externalities” In Journal of economic theory 85.2 Elsevier, 1999, pp. 258–293
  • [63] Søren Johansen “The extremal convex functions” In Mathematica Scandinavica 34.1 JSTOR, 1974, pp. 61–68
  • [64] Navin Kartik, Andreas Kleiner and Richard Van Weelden “Delegation in veto bargaining” In American Economic Review 111.12 American Economic Association 2014 Broadway, Suite 305, Nashville, TN 37203, 2021, pp. 4046–4087
  • [65] Victor Klee “Some new results on smoothness and rotundity in normed linear spaces” In Mathematische Annalen 139.1 Springer, 1959, pp. 51–63
  • [66] Victor L Klee Jr “Extremal structure of convex sets” In Archiv der Mathematik 8.3 Springer, 1957, pp. 234–240
  • [67] Victor L Klee Jr “Extremal structure of convex sets. II” In Mathematische Zeitschrift 69.1 Springer, 1958, pp. 90–104
  • [68] Andreas Kleiner “Optimal Delegation in a Multidimensional World” In arXiv preprint arXiv:2208.11835, 2022
  • [69] Andreas Kleiner and Alejandro Manelli “Strong Duality in Monopoly Pricing” In Econometrica 87.4 Wiley Online Library, 2019, pp. 1391–1396
  • [70] Andreas Kleiner, Benny Moldovanu and Philipp Strack “Extreme points and majorization: Economic applications” In Econometrica 89.4 Wiley Online Library, 2021, pp. 1557–1593
  • [71] Andreas Kleiner, Benny Moldovanu, Philipp Strack and Mark Whitmeyer “The Extreme Points of Fusions” In arXiv preprint arXiv:2409.10779, 2024 URL: https://arxiv.org/abs/2409.10779
  • [72] Frédéric Koessler and David Martimort “Optimal delegation with multi-dimensional decisions” In Journal of Economic Theory 147.5 Elsevier, 2012, pp. 1850–1881
  • [73] Alexander V. Kolesnikov, Fedor Sandomirskiy, Aleh Tsyvinski and Alexander P. Zimin “Beckmann’s Approach to Multi-Item Multi-Bidder Auctions” In arXiv preprint arXiv:2203.06837, 2022 URL: https://arxiv.org/abs/2203.06837
  • [74] Anton Kolotilin and Andriy Zapechelnyuk “Persuasion Meets Delegation” In arXiv preprint arXiv:1902.02628, 2019 URL: https://arxiv.org/abs/1902.02628
  • [75] Eugen Kováč and Tymofiy Mylovanov “Stochastic mechanisms in settings without monetary transfers: The regular case” In Journal of Economic Theory 144.4 Elsevier, 2009, pp. 1373–1395
  • [76] Ka-Sing Lau “On strongly exposing functionals” In Journal of the Australian Mathematical Society 21.3 Cambridge University Press, 1976, pp. 362–367
  • [77] Xinye Li and Andrew Chi-Chih Yao “On revenue maximization for selling multiple independently distributed items” In Proceedings of the National Academy of Sciences 110.28 National Acad Sciences, 2013, pp. 11232–11237
  • [78] Shu Lu and Stephen M Robinson “Normal fans of polyhedral convex sets: Structures and connections” In Set-Valued Analysis 16.2 Springer, 2008, pp. 281–305
  • [79] Diane Maclagan and Bernd Sturmfels “Introduction to tropical geometry” American Mathematical Society, 2021
  • [80] Alejandro M Manelli and Daniel R Vincent “Bundling as an Optimal Selling Mechanism for a Multiple-Good Monopolist” In Journal of Economic Theory 127.1 Elsevier, 2006, pp. 1–35
  • [81] Alejandro M Manelli and Daniel R Vincent “Multidimensional mechanism design: Revenue maximization and the multiple-good monopoly” In Journal of Economic Theory 137.1 Elsevier, 2007, pp. 153–185
  • [82] Alejandro M Manelli and Daniel R Vincent “Bayesian and Dominant-Strategy Implementation in the Independent Private-Values Model” In Econometrica 78.6 Wiley Online Library, 2010, pp. 1905–1938
  • [83] Alejandro M Manelli and Daniel R Vincent “Multidimensional Mechanism Design: Revenue Maximization and the Multiple-Good Monopoly. A Corrigendum” In Journal of Economic Theory 148.6 Elsevier, 2012, pp. 2492–2493
  • [84] David Martimort and Aggey Semenov “Continuity in mechanism design without transfers” In Economics Letters 93.2 Elsevier, 2006, pp. 182–189
  • [85] Eric Maskin and John Riley “Optimal auctions with risk averse buyers” In Econometrica: Journal of the Econometric Society JSTOR, 1984, pp. 1473–1518
  • [86] Steven A Matthews “On the implementability of reduced form auctions” In Econometrica: Journal of the Econometric Society JSTOR, 1984, pp. 1519–1522
  • [87] R Preston McAfee, John McMillan and Michael D Whinston “Multiproduct monopoly, commodity bundling, and correlation of values” In The Quarterly Journal of Economics 104.2 MIT Press, 1989, pp. 371–383
  • [88] Peter McMullen “Representations of polytopes and polyhedral sets” In Geometriae dedicata 2.1 Springer, 1973, pp. 83–99
  • [89] Nahum D Melumad and Toshiyuki Shibano “Communication in settings with no transfers” In The RAND Journal of Economics JSTOR, 1991, pp. 173–198
  • [90] Domenico Menicucci, Sjaak Hurkens and Doh-Shin Jeon “On the optimality of pure bundling for a monopolist” In Journal of Mathematical Economics 60 Elsevier, 2015, pp. 33–42
  • [91] Walter Meyer “Decomposing plane convex bodies” In Archiv der Mathematik 23.1 Springer, 1972, pp. 534–536
  • [92] Walter Meyer “Indecomposable Polytopes” In Transactions of the American Mathematical Society 190 American Mathematical Society, 1974, pp. 77–86
  • [93] G Mielczarek “On extreme convex subsets of the plane” In Acta Mathematica Hungarica 78.3 Springer, 1998, pp. 213–226
  • [94] Roger B Myerson “Optimal Auction Design” In Mathematics of Operations Research 6.1 INFORMS, 1981, pp. 58–73
  • [95] Axel Niemeyer and Justus Preusser “Optimal Allocation with Peer Information” In arXiv preprint arXiv:2410.08954, 2024 URL: https://arxiv.org/abs/2410.08954
  • [96] Afshin Nikzad “Constrained majorization: Applications in mechanism design” In Proceedings of the 23rd ACM Conference on Economics and Computation, 2022, pp. 330–331
  • [97] Afshin Nikzad “Multi-Criteria Allocation Mechanisms: Constraints and Comparative Statics” In Available at SSRN, 2024 URL: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=4695785
  • [98] Volker Nocke and Michael D. Whinston “Merger policy with merger choice” In American Economic Review 103.2, 2013, pp. 1006–1033
  • [99] Thomas R Palfrey “Bundling decisions by a multiproduct monopolist with incomplete information” In Econometrica: Journal of the Econometric Society JSTOR, 1983, pp. 463–483
  • [100] Gregory Pavlov “Optimal mechanism for selling two goods” In The BE Journal of Theoretical Economics 11.1 De Gruyter, 2011
  • [101] Guillermo Pineda Villavicencio “Polytopes and Graphs”, Cambridge Studies in Advanced Mathematics Cambridge University Press, 2024
  • [102] John Riley and Richard Zeckhauser “Optimal selling strategies: When to haggle, when to hold firm” In The Quarterly Journal of Economics 98.2 MIT Press, 1983, pp. 267–289
  • [103] Jean-Charles Rochet “A necessary and sufficient condition for rationalizability in a quasi-linear context” In Journal of Mathematical Economics 16.2 Elsevier, 1987, pp. 191–200
  • [104] Jean-Charles Rochet and Philippe Choné “Ironing, sweeping, and multidimensional screening” In Econometrica JSTOR, 1998, pp. 783–826
  • [105] Jean-Charles Rochet and Lars A Stole “The economics of multidimensional screening” In Econometric Society Monographs 35 Cambridge University Press, 2003, pp. 150–197
  • [106] R Tyrrell Rockafellar and Roger J-B Wets “Variational Analysis” Springer Science & Business Media, 2009
  • [107] R. Rockafellar “Convex Analysis” Princeton University Press, 1997
  • [108] Thomas Romer and Howard Rosenthal “Political Resource Allocation, Controlled Agendas, and the Status Quo” In Public Choice JSTOR, 1978, pp. 27–43
  • [109] Aviad Rubinstein and S Matthew Weinberg “Simple Mechanisms for a Subadditive Buyer and Applications to Revenue Monotonicity” In ACM Transactions on Economics and Computation (TEAC) 6.3-4 ACM, 2018, pp. 1–25
  • [110] Michael Saks and Lan Yu “Weak monotonicity suffices for truthfulness on convex domains” In Proceedings of the 6th ACM Conference on Electronic Commerce, 2005, pp. 286–293
  • [111] Gabriella Salinetti and Roger J-B Wets “On the convergence of sequences of convex sets in finite dimensions” In SIAM Review 21.1 SIAM, 1979, pp. 18–33
  • [112] GT Sallee “Minkowski decomposition of convex sets” In Israel Journal of Mathematics 12 Springer, 1972, pp. 266–276
  • [113] Rene Saran “A Dynamic Optimization Approach to Delegation with an Application to Volunteer Contracts” In Available at SSRN, 2022 URL: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=4703145
  • [114] Richard Schmalensee “Gaussian demand and commodity bundling” In Journal of Business JSTOR, 1984, pp. S211–S230
  • [115] Rolf Schneider “Convex Bodies: The Brunn–Minkowski Theory” Cambridge University Press, 2014
  • [116] Geoffrey C. Shephard “Decomposable convex polyhedra” In Mathematika 10.2 London Mathematical Society, 1963, pp. 89–95
  • [117] Ruth Silverman “Decomposition of plane convex sets I” In Pacific Journal of Mathematics 47.2 Mathematical Sciences Publishers, 1973, pp. 521–530
  • [118] Zeev Smilansky “Decomposability of polytopes and polyhedra” In Geometriae dedicata 24.1 Springer, 1987, pp. 29–49
  • [119] Ngoc Mai Tran and Josephine Yu “Product-mix auctions and tropical geometry” In Mathematics of Operations Research 44.4 INFORMS, 2019, pp. 1396–1411
  • [120] Quitzé Valenzuela-Stookey “Interim Allocations and Greedy Algorithms” Unpublished manuscript, 2023
  • [121] Rakesh V Vohra “Mechanism Design: A Linear Programming Approach” Cambridge University Press, 2011
  • [122] Robert B Wilson “Nonlinear Pricing” Oxford University Press, USA, 1993
  • [123] Gerhard Winkler “Extreme points of moment sets” In Mathematics of Operations Research 13.4 INFORMS, 1988, pp. 581–587
  • [124] Frank Yang “Nested bundling”, https://web.stanford.edu/~shuny/papers/bundling.pdf, 2023
  • [125] Kai Hao Yang and Alexander K Zentefis “Monotone Function Intervals: Theory and Applications” In American Economic Review 114.8 American Economic Association, 2024, pp. 2239–2270