Rank-Guaranteed Auctions111We thank Mohammad Akbarpour, Martin Bichler, Tilman Börgers, Kim-Sau Chung, Rahul Deb, Jason Hartline, Paul Klemplerer, Fuhito Kojima, Andrew Komo, Paul Milgrom, Michael Ostrovsky, Antonio Penta, Ning Sun, Satoru Takahashi, and Andrzej Skrzypacz for helpful discussions. This manuscript subsumes He et al., (2022).

Wei He Department of Economics, The Chinese University of Hong Kong, [email protected]    Jiangtao Li School of Economics, Singapore Management University, [email protected]    Weijie Zhong Graduate School of Business, Stanford University, [email protected]
Abstract

We propose a combinatorial ascending auction that is “approximately” optimal, requiring minimal rationality to achieve this level of optimality, and is robust to strategic and distributional uncertainties. Specifically, the auction is rank-guaranteed, meaning that for any menu \mathcal{M}caligraphic_M and any valuation profile, the ex-post revenue is guaranteed to be at least as high as the highest revenue achievable from feasible allocations, taking the (||+1)thsuperscript1𝑡(|\mathcal{M}|+1)^{th}( | caligraphic_M | + 1 ) start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest valuation for each bundle as the price. Our analysis highlights a crucial aspect of combinatorial auction design, namely, the design of menus. We provide simple and approximately optimal menus in various settings.

1 Introduction

Auctions play a critical role in economic activities. For example, the online advertising sector generates trillions of dollars annually through the auctioning of advertising “slots”. The Federal Communications Commission (FCC) has collected over 200 billion dollars via auctioning radio spectrum. Despite the critical role these auctions play, there is a surprising lack of theoretical groundwork to navigate the intricacies of auction design. This gap in knowledge stems from a unique challenge: “not all slots are created equal”—bidders typically have complex, combinatorial preferences for different slots. For instance, YouTube intersperses promotional videos at regular intervals within longer content. Here, some advertisers might see value in the repetition of their ads, leveraging the complementarity, while others may fear overexposure could lead to negative perceptions akin to “spamming.” Likewise, in the realm of telecommunications, while larger service providers may pursue nationwide radio spectrum licenses to maximize their coverage, smaller providers often seek only regional licenses, prioritizing local markets over national presence.

The theoretical study of auctions, in contrast to reality, takes aggressive simplifications to an extent that overlooks the nuanced realities of the markets. The game theoretic approach focuses on the incentives of the bidders (incentive compatibility) and auctioneers (optimality), while limiting to very specific environments. Iconic theories, such as Myerson, (1981), made progress by assuming a single item, independent valuations among bidders, full Bayesian rationality, and shared prior beliefs. None of these assumptions hold water in the complex scenarios described earlier, highlighting a clear disconnect. Conversely, the domain of operations research and computer science typically emphasizes the procedural aspects of communicating preferences and determining assignments under complete preferential complexity. This focus often neglects the incentives for bidders to reveal their preferences truthfully and for auctioneers to maximize revenue (see, e.g., a review by Cramton et al., (2006)). Therefore, the development of a comprehensive auction theory that simultaneously addresses both incentives and complexity represents a significant and unfulfilled challenge.

The goal of this study is to address the dilemma between incentives and complexity in auction design. This resonates with what Carroll, (2019) speculates as the “future of economic design”:

“My expectation—and my hope—is that progress in mechanism design over the coming decades will come from developing more useful general models of preferences, information, and actions in complex environments, relatively free of structural assumptions; and developing conceptual tools to argue for why certain kinds of mechanisms will work well in such environments.”

To tackle the complexity of the practical environments, the same article then proposes:

“Making this progress toward more free-form models will require a shift in the criteria by which research in economic theory is evaluated.”

The methodology we develop in this paper hinges on the shift to a novel criterion of approximate optimality, termed the rank guarantee: the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-rank guarantee is the maximal ex-post revenue when each feasible bundle can be sold at the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest value among all bidders. With this approximation, we achieve a near-optimal resolution of the dilemma: it is possible to design a simple auction mechanism to respect the incentives—the auctioneer achieves a rank-guaranteed revenue as long as bidders avoid “obviously” bad strategies—while accommodating the full complexity—fully combinatorial preferences, without the need for any Bayesian prior.

In our model, an auctioneer sells multiple items to several (potentially a large number of) strategic bidders, each with private valuations. We introduce a multi-item variant of the open ascending auction termed the (C)ombinatorial (As)cending (A)uction (CASA). Prior to the auction, the auctioneer curates a menu of item bundles for allocation, denoted by \mathcal{M}caligraphic_M. With the formal game theoretic form of the auction described in Section 2, this auction model distills down to two straightforward principles:

  1. 1.

    Bidders are allowed to place binding bids (increase prices) on any assortment of bundles from the menu, even if these selections overlap.

  2. 2.

    The auction concludes when bid prices stabilize, with the winning bids being those that maximize the total selling price.

Our findings reveal that CASA resolves the dilemma within certain approximation bounds. First, we show that the outcome of CASA respects the incentives of both bidders and the auctioneer:

  • Bidder rationality: All of our results apply to any non-obviously dominated strategy profile, a collection that excludes strategies that are obviously dominated in the sense that even in the most favorable case, they underperform some other strategies in their least favorable case (see Li, (2017) and Li and Dworczak, (2021)). In other words, we allow the bidders to bid fully strategically, while making a weak rationality assumption that all we know is that they avoid obviously bad choices.

  • Approximate auctioneer optimality: In CASA, any non-obviously dominated strategy profile yields an ex-post revenue that is rank-guaranteed — achieving the maximal revenue when each bundle within the menu \mathcal{M}caligraphic_M can be sold at the (||+1)thsuperscript1𝑡(|\mathcal{M}|+1)^{th}( | caligraphic_M | + 1 ) start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest value among all bidders. Since the 1stsuperscript1𝑠𝑡1^{st}1 start_POSTSUPERSCRIPT italic_s italic_t end_POSTSUPERSCRIPT-guarantee is the full ex-post trading surplus, the approximation we take is close in “ordinal distance” from the full surplus.

Second, we show how CASA maintains its performance in complex, unknown environments:

  • Prior-free: The auction format, the selection of strategy profile, and the revenue guarantee does not depend on any Bayesian prior on either the auctioneer’s side or the bidders’ side, rendering the mechanism prior-free and our rank guarantee a prior-free approximation (see, e.g., Chapter 5 and 7 of Hartline, (2013)).

  • Distributional robustness: We quantify the rank-guarantee using canonical robust optimality criteria, i.e., the minimal ex-ante expectation when an adversarial nature chooses the joint distribution of values against the mechanism (see, e.g., Carroll, (2017)). In the worst case, the revenue from CASA approximates the total surplus at the rate of O(||2N)𝑂superscript2𝑁O\left(\frac{|\mathcal{M}|^{2}}{N}\right)italic_O ( divide start_ARG | caligraphic_M | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N end_ARG ), i.e., CASA asymptotically achieves full surplus extraction when the number of bidders is large relative to the menu size.

To our knowledge, this paper is the first to introduce the rank-guarantee approximation. Compared to canonical approximation notions like constant-ratio guarantee and maxmin guarantee, the rank guarantee offers the added benefit of being easy to evaluate across various Bayesian and non-Bayesian environments. On the one hand, the rank guarantee provides an easily computable lower bound on the worst-case revenue, even when the underlying environment changes, such as when there are more bidders, when the menu changes, or when the auctioneer has additional information on the distribution of bidders’ valuations. Such adaptivity allows us to further simplify CASA by studying menu design. On the other hand, the lower bound performance is also straightforward to assess outside adversarial scenarios. For example, in the canonical setting where the values are independent and identically distributed, the rank guarantee is an appealing approximation when N𝑁Nitalic_N is large, as all order statistics converge to the upper bound of the valuation support. More generally, beyond understanding the worst-case scenario, the rank guarantee can be useful to consider the potential upside of a mechanism in best-case scenarios.

Our framework highlights a crucial aspect of combinatorial auction design, namely, the design of menus. Crucially, the rank guarantee we derive reveals a novel trade-off between menu sufficiency and approximation efficiency: a more complete menu achieves a higher benchmark total surplus but increases the rank |||\mathcal{M}|| caligraphic_M |. Therefore, to close the approximation gap, a key exercise is to reduce the menu size while maintaining the allocation efficiency, leveraging further knowledge about the bidders’ preferences. We focus on a specific type of sufficient menus that improve approximation efficiency “for free”—menus that achieve the same benchmark total surplus as the complete menu. Specifically, we show that when the bidder’s preference exhibits canonical preference structures, without loss of the benchmark total surplus, the size of menus can be reduced to be polynomial in the number of items being auctioned and so is the convergence rate of revenue guarantee. The result is summarized in Table 1.

Preference Simple and Sufficient Menu Rank k𝑘kitalic_k
Weak substitutability Individual items O(M)𝑂𝑀O(M)italic_O ( italic_M )
Weak complementarity Grand bundle 2222
Partitional complementarity Partitional bundles O(M)𝑂𝑀O(M)italic_O ( italic_M )
Homogeneous goods Menu of quantities O(M2)𝑂superscript𝑀2O(M^{2})italic_O ( italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT )
Table 1: Simple and sufficient menus

The remainder of the introduction reviews related literature. Section 2 introduces the auction format of CASA and the notion of rank-guarantee, and shows that CASA achieves the rank-guarantee. Section 3 bounds the worst-case performance of rank-guaranteed auctions under distributional uncertainties. Section 4 explores specific preference structures where CASA with simple menus performs as well as the complete menu.

1.1 Related literature

(Approximately) optimal auction design

Beyond the simple environment studied in Myerson, (1981) and Bulow and Klemperer, (1996), solving for the exact optimal mechanism with confounding factors like multiple heterogeneous items, bounded distributional knowledge or bounded rationality is generally intractable. Various alternative optimality notions have been proposed to make progress (see surveys by Roughgarden, (2015) and Hartline, (2013)). Aggarwal and Hartline, (2006) and Goldberg and Hartline, (2001) obtained the “constant fraction” approximation in the auction of sponsored search and digital goods. Following a broader literature on robust mechanism design pioneered by Carroll, (2017), various authors have studied “robustly” optimal auctions that maximize the distributional worst-case revenue.222This research direction complements the large body of papers that focus on the case in which the designer does not have reliable information about the agents’ hierarchies of beliefs about each other while assuming the knowledge of the payoff environment; see, for example, Bergemann and Morris, (2005), Chung and Ely, (2007), Chen and Li, (2018), Du, (2018), Brooks and Du, (2021), Yamashita and Zhu, (2022), and Brooks and Du, (2023). Particularly, He and Li, (2022), Zhang, (2022), and Suzdaltsev, (2022) study robust versions of the single-unit auction problem in the distribution robust framework where the auctioneer has non-Bayesian uncertainty about the joint distribution of the bidders’ valuations. In this paper, we propose the notion of rank-guarantee, a distinct notion of approximate optimality. In Section 3, we apply the distributional robustness analysis similar to that of Carroll, (2017) and show that the rank-guarantee has an appealing worst-case performance.

Multi-item auctions

Beyond the efficient Vickrey auction, few theoretic results have been established regrading multi-item auctions with combinatorial preferences. Jehiel and Moldovanu, 2001a point out the vulnerability of efficiency under multidimensional bidder information. Ausubel and Milgrom, (2002) point out the poor revenue performance and strategic vulnerability of the Vickrey auction and propose simultaneous ascending auctions with package bidding (SAAPB). The multi-item auction design problem has also been extensively studied in the field of combinatorial auctions (see Cramton et al., (2006) for a survey). This literature mainly focuses on (approximately) efficient auction design and their communication/computational complexity, which is orthogonal to our focus on revenue performance and bidder incentives. Compared to other proposals like SAAPB (Ausubel and Milgrom, (2002)) and the Combinatorial Clock Auction (CCA, see Ausubel et al., (2006) and Levin and Skrzypacz, (2016)), CASA uses a simpler “pay-as-bid” rule, as opposed to personalized prices in SAAPB and demand reporting in CCA. Importantly, we assume that the bidders are fully strategic instead of single-minded. The menu design problem we tackle in Section 4 is akin to Rothkopf et al., (1998), which seeks to make a combinatorial auction computationally tractable by restricting the menu. We achieved the same goal while allowing for strategic bidders and maintaining the approximate optimality of the revenue.

Implementation in strategies that are not obviously dominated

We study outcomes when agents are rational in the sense of avoiding obviously dominated strategies. This solution concept draws from the idea of obvious strategy-proof mechanisms (see Li, (2017)) and is systematically studied in Li and Dworczak, (2021). As in Li and Dworczak, (2021), we assume that agents avoid obviously dominated strategies, but refrain from making assumptions regarding how agents select among strategies that are not obviously dominated. This methodology aligns with the spirit of implementation in undominated strategies; see for example Carroll, (2014), Börgers, (1991), Jackson, (1992), and Yamashita, (2015).

2 CASA and rank-guarantee

2.1 The auction environment

There is a set S𝑆Sitalic_S of M𝑀Mitalic_M items to be sold to N𝑁Nitalic_N bidders. Let 𝒩={1,2,,N}𝒩12𝑁\mathcal{N}=\{1,2,\ldots,N\}caligraphic_N = { 1 , 2 , … , italic_N }. We write bS𝑏𝑆b\subseteq Sitalic_b ⊆ italic_S to denote a generic bundle of items. Let 𝒗n={vbn}bSsuperscript𝒗𝑛subscriptsuperscriptsubscript𝑣𝑏𝑛𝑏𝑆\bm{v}^{n}=\{v_{b}^{n}\}_{b\subseteq S}bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = { italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_b ⊆ italic_S end_POSTSUBSCRIPT denote the valuation vector of bidder n𝑛nitalic_n, where vbnsuperscriptsubscript𝑣𝑏𝑛v_{b}^{n}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT is bidder n𝑛nitalic_n’s valuation of bundle b𝑏bitalic_b. Valuations are normalized so that vn=0subscriptsuperscript𝑣𝑛0v^{n}_{\emptyset}=0italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∅ end_POSTSUBSCRIPT = 0 and vbn[v¯,v¯]superscriptsubscript𝑣𝑏𝑛¯𝑣¯𝑣v_{b}^{n}\in[\underline{v},\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}% \mkern 1.5mu]italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∈ [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] (v¯0¯𝑣0\underline{v}\geq 0under¯ start_ARG italic_v end_ARG ≥ 0) for all b𝑏b\neq\emptysetitalic_b ≠ ∅. A generic valuation profile is denoted by 𝒗=(𝒗1,𝒗2,,𝒗N)𝒗superscript𝒗1superscript𝒗2superscript𝒗𝑁\bm{v}=(\bm{v}^{1},\bm{v}^{2},\ldots,\bm{v}^{N})bold_italic_v = ( bold_italic_v start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , bold_italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , bold_italic_v start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ). Let 2Ssuperscript2𝑆\mathcal{M}\subseteq 2^{S}caligraphic_M ⊆ 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT denote a menu of bundles chosen by the auctioneer. While \mathcal{M}caligraphic_M is a choice variable of the auctioneer, for now we take it as exogenously given; we defer the discussion of menu design to Section 4. Assume that N||+1𝑁1N\geq|\mathcal{M}|+1italic_N ≥ | caligraphic_M | + 1. Let

()={X|b,bX,bb=}conditional-set𝑋formulae-sequencefor-all𝑏superscript𝑏𝑋𝑏superscript𝑏\displaystyle\mathcal{B}(\mathcal{M})=\{X\subseteq\mathcal{M}\,|\,\forall b,b^% {\prime}\in X,\,b\cap b^{\prime}=\emptyset\}caligraphic_B ( caligraphic_M ) = { italic_X ⊆ caligraphic_M | ∀ italic_b , italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_X , italic_b ∩ italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ∅ }

denote the set of feasible allocations of bundles within the menu \mathcal{M}caligraphic_M, i.e., all collections consisting of non-overlapping bundles.

2.2 The Combinatorial Ascending Auction

We define the Combinatorial Ascending Auction (CASA) as follows. The auction has an iterative structure, with the “state of the auction” characterized by the identity of the leading bidder and the leading price for each bundle. Initially, the leading price for each bundle is zero and none of the bidders is a leading bidder for any bundle. Bidders take turns raising the bids on the bundles, which determines new leading bidders and leading prices. The process repeats itself until when there are no new bids on any bundle. At that point, the auction stops. The auctioneer chooses a feasible allocation to maximize revenue, taking the leading prices as the prices for the bundles. There is also an activity rule designed to ensure that bidding activity starts out high and declines during the auction as prices rise far enough to discourage some bidders from continuing.

Formally, let P+𝑃superscriptP\subset\mathbb{R}^{+}italic_P ⊂ blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT be a finite grid of feasible bids with grid size ϵitalic-ϵ\epsilonitalic_ϵ and maxP>v¯𝑃¯𝑣\max P>\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5muroman_max italic_P > over¯ start_ARG italic_v end_ARG.

  1. (1)

    Initialization stage t=0𝑡0t=0italic_t = 0. Define

    • the leading bidder vector at stage 00: ϕ0=(ϕb0)b=𝟎superscriptbold-italic-ϕ0subscriptsubscriptsuperscriptitalic-ϕ0𝑏𝑏0\bm{\phi}^{0}=(\phi^{0}_{b})_{b\in\mathcal{M}}=\bm{0}bold_italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = ( italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT = bold_0,

    • the leading price vector at stage 00: 𝒑0=(pb0)b=𝟎superscript𝒑0subscriptsubscriptsuperscript𝑝0𝑏𝑏0\bm{p}^{0}=(p^{0}_{b})_{b\in\mathcal{M}}=\bm{0}bold_italic_p start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = ( italic_p start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT = bold_0,

    • the set of active bidders at stage 00: 𝒩0={1,2,,N}superscript𝒩012𝑁\mathcal{N}^{0}=\{1,2,\ldots,N\}caligraphic_N start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = { 1 , 2 , … , italic_N }.

  2. (2)

    Bidding stage t1𝑡1t\geq 1italic_t ≥ 1. An active bidder n𝒩t1𝑛superscript𝒩𝑡1n\in\mathcal{N}^{t-1}italic_n ∈ caligraphic_N start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT observes (𝒑t1,{b|ϕbt1=n})superscript𝒑𝑡1conditional-set𝑏superscriptsubscriptitalic-ϕ𝑏𝑡1𝑛(\bm{p}^{t-1},\{b\,|\,\phi_{b}^{t-1}=n\})( bold_italic_p start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT , { italic_b | italic_ϕ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT = italic_n } ), and decides whether to quit, which bundles to bid on, and how much to bid.333The bidder selection rule and the observability of history is inconsequential for our analysis. For concreteness, we consider the selection rule that active bidders are cycled in ascending order according to their indices, and the observability of history is minimized to maximally protect privacy.

    • Bidder n𝑛nitalic_n may choose to quit by submitting an empty bid only if {b|ϕbt1=n}=conditional-set𝑏superscriptsubscriptitalic-ϕ𝑏𝑡1𝑛\{b\,|\,\phi_{b}^{t-1}=n\}=\emptyset{ italic_b | italic_ϕ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT = italic_n } = ∅, i.e., bidder n𝑛nitalic_n is not a leading bidder for any bundle in stage t1𝑡1t-1italic_t - 1. Quitting is irreversible, that is, if bidder n𝑛nitalic_n chooses to quit, then bidder n𝑛nitalic_n becomes an inactive bidder and does not participate in future bidding rounds. Update:

      • ϕt=ϕt1superscriptbold-italic-ϕ𝑡superscriptbold-italic-ϕ𝑡1\bm{\phi}^{t}=\bm{\phi}^{t-1}bold_italic_ϕ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = bold_italic_ϕ start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT, 𝒑t=𝒑t1superscript𝒑𝑡superscript𝒑𝑡1\bm{p}^{t}=\bm{p}^{t-1}bold_italic_p start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = bold_italic_p start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT, 𝒩t=𝒩1{n}superscript𝒩𝑡superscript𝒩1𝑛\mathcal{N}^{t}=\mathcal{N}^{-1}\setminus\{n\}caligraphic_N start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = caligraphic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∖ { italic_n }.

    • If bidder n𝑛nitalic_n chooses not to quit, then she submits a bid—a nonempty set of bundle-price pairs {(b,pb)}×P𝑏subscript𝑝𝑏𝑃\{(b,p_{b})\}\subset\mathcal{M}\times P{ ( italic_b , italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) } ⊂ caligraphic_M × italic_P, subject to the requirements that (1) Leading bids are binding: if bidder n𝑛nitalic_n is the leading bidder at some bundle in stage t1𝑡1t-1italic_t - 1, then she must include that bundle in her bid with a bid that is weakly higher than the current leading price for that bundle, and (2) Minimum bid increment: if bidder n𝑛nitalic_n would like to bid on some bundle for which she is not the leading bidder, then her bid for that bundle must be strictly higher than the current leading price for that bundle. Update:

      • ϕbt=nsuperscriptsubscriptitalic-ϕ𝑏𝑡𝑛\phi_{b}^{t}=nitalic_ϕ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = italic_n and pbt=pbsuperscriptsubscript𝑝𝑏𝑡subscript𝑝𝑏p_{b}^{t}=p_{b}italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT for any bundle b𝑏bitalic_b included in her bid,

      • ϕbt=ϕbt1superscriptsubscriptitalic-ϕsuperscript𝑏𝑡superscriptsubscriptitalic-ϕsuperscript𝑏𝑡1\phi_{b^{\prime}}^{t}=\phi_{b^{\prime}}^{t-1}italic_ϕ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT and pbt=pbt1superscriptsubscript𝑝superscript𝑏𝑡superscriptsubscript𝑝superscript𝑏𝑡1p_{b^{\prime}}^{t}=p_{b^{\prime}}^{t-1}italic_p start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = italic_p start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT for any bundle bsuperscript𝑏b^{\prime}italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT not included in her bid,

      • 𝒩t=𝒩t1superscript𝒩𝑡superscript𝒩𝑡1\mathcal{N}^{t}=\mathcal{N}^{t-1}caligraphic_N start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = caligraphic_N start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT.

    Then, move on to the bidding stage t+1𝑡1t+1italic_t + 1.

  3. (3)

    Allocation. The auction ends (in stage T𝑇Titalic_T) when the leading prices stay constant for N𝑁Nitalic_N consecutive periods. The auctioneer chooses a feasible allocation to maximize

    max𝒃()b𝒃pbT.subscript𝒃subscript𝑏𝒃superscriptsubscript𝑝𝑏𝑇\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,p_{b% }^{T}.roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT .

    Denote the maximizer by 𝒃superscript𝒃\bm{b}^{*}bold_italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Each bundle b𝒃𝑏superscript𝒃b\in\bm{b}^{*}italic_b ∈ bold_italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is allocated to ϕbTsuperscriptsubscriptitalic-ϕ𝑏𝑇\phi_{b}^{T}italic_ϕ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT at the price pbTsuperscriptsubscript𝑝𝑏𝑇p_{b}^{T}italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT.

In words, the auction format of CASA runs parallel ascending auctions for each bundle b𝑏b\in\mathcal{M}italic_b ∈ caligraphic_M. Then, the items are allocated to maximize the total price. We discuss possible variants of CASA and its relation to existing auction formats in Section 2.4.

2.3 The rank-guarantee of CASA

In this subsection, we study the strategic behavior of the bidders and establish the rank-guarantee property of CASA. We only assume minimal rationality on the part of the bidders—bidders are rational in the sense of not playing obviously dominated strategies. Formally we adopt the solution concept of implementation in strategies that are not obviously dominated (see Li, (2017) and Li and Dworczak, (2021)). We first sketch the intuition, and then provide the formal arguments.

At any history, consider a non-leading bidder’s choice as to whether to quit. Obviously, as quitting is irreversible, quitting the auction leads to a best possible outcome of a zero payoff. Suppose that there is some bundle for which the bidder’s valuation is higher than the current leading price for that bundle. Consider the following strategy where the bidder raises the price for this particular bundle and never revises her bid afterwards. Clearly, this continuing strategy guarantees a non-negative payoff for the bidder. Thus, at least for the purpose of deciding whether to quit, it is “obviously optimal” not to quit.

More formally, let h=(t,(𝒩0,,𝒩t1),(𝒑0,𝒑t1),(ϕ0,,ϕt1))𝑡superscript𝒩0superscript𝒩𝑡1superscript𝒑0superscript𝒑𝑡1superscriptbold-italic-ϕ0superscriptbold-italic-ϕ𝑡1h=(t,(\mathcal{N}^{0},\ldots,\mathcal{N}^{t-1}),(\bm{p}^{0}\ldots,\bm{p}^{t-1}% ),(\bm{\phi}^{0},\ldots,\bm{\phi}^{t-1}))italic_h = ( italic_t , ( caligraphic_N start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , … , caligraphic_N start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT ) , ( bold_italic_p start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT … , bold_italic_p start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT ) , ( bold_italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , … , bold_italic_ϕ start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT ) ) denote a history of the game in stage t𝑡titalic_t, Htsubscript𝐻𝑡H_{t}italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT the set of such histories in stage t𝑡titalic_t, and H=t0Ht𝐻subscript𝑡0subscript𝐻𝑡H=\cup_{t\geq 0}H_{t}italic_H = ∪ start_POSTSUBSCRIPT italic_t ≥ 0 end_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. Suppose that bidder n𝑛nitalic_n is the active bidder in some stage t𝑡titalic_t and Insubscript𝐼𝑛I_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is bidder n𝑛nitalic_n’s information set. Then the observed prices 𝒑𝒑\bm{p}bold_italic_p and n𝑛nitalic_n’s leading bundles 𝒃𝒃\bm{b}bold_italic_b are the same for all hInsubscript𝐼𝑛h\in I_{n}italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. Let nsubscript𝑛\mathcal{I}_{n}caligraphic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote all information sets of n𝑛nitalic_n. Let sn:n2×+:subscript𝑠𝑛subscript𝑛superscript2superscripts_{n}:\mathcal{I}_{n}\to 2^{\mathcal{M}\times\mathbb{R}^{+}}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : caligraphic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT → 2 start_POSTSUPERSCRIPT caligraphic_M × blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT denote bidder n𝑛nitalic_n’s (pure behavioral) strategy and un(𝒔,𝒗n|h)subscript𝑢𝑛𝒔conditionalsuperscript𝒗𝑛u_{n}(\bm{s},\bm{v}^{n}|h)italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_s , bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT | italic_h ) the payoff to bidder n𝑛nitalic_n given valuation vector 𝒗nsuperscript𝒗𝑛\bm{v}^{n}bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, strategy profile 𝒔𝒔\bm{s}bold_italic_s, conditional on the current history hhitalic_h and n𝑛nitalic_n bidding in period t𝑡titalic_t.

Definition 1.

A bidding strategy sn:n2×P:subscript𝑠𝑛subscript𝑛superscript2𝑃s_{n}:\mathcal{I}_{n}\to 2^{\mathcal{M}\times P}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : caligraphic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT → 2 start_POSTSUPERSCRIPT caligraphic_M × italic_P end_POSTSUPERSCRIPT is obviously dominated if there exists snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT such that at any earliest point of departure Insubscript𝐼𝑛I_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT between snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT,

supsn,hInun(𝒔,𝒗n|h)subscriptsupremumsubscript𝑠𝑛subscript𝐼𝑛subscript𝑢𝑛𝒔conditionalsuperscript𝒗𝑛\displaystyle\sup_{s_{-n},h\in I_{n}}u_{n}(\bm{s},\bm{v}^{n}|h)roman_sup start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_s , bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT | italic_h ) infsn,hInun(sn,𝒔n,𝒗n|h);absentsubscriptinfimumsubscript𝑠𝑛subscript𝐼𝑛subscript𝑢𝑛subscriptsuperscript𝑠𝑛subscript𝒔𝑛conditionalsuperscript𝒗𝑛\displaystyle\leq\inf_{s_{-n},h\in I_{n}}u_{n}(s^{\prime}_{n},\bm{s}_{-n},\bm{% v}^{n}|h);≤ roman_inf start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT | italic_h ) ;
infsn,hInun(𝒔,𝒗n|h)subscriptinfimumsubscript𝑠𝑛subscript𝐼𝑛subscript𝑢𝑛𝒔conditionalsuperscript𝒗𝑛\displaystyle\inf_{s_{-n},h\in I_{n}}u_{n}(\bm{s},\bm{v}^{n}|h)roman_inf start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_s , bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT | italic_h ) <supsn,hInun(sn,𝒔n,𝒗n|h).absentsubscriptsupremumsubscript𝑠𝑛subscript𝐼𝑛subscript𝑢𝑛subscriptsuperscript𝑠𝑛subscript𝒔𝑛conditionalsuperscript𝒗𝑛\displaystyle<\sup_{s_{-n},h\in I_{n}}u_{n}(s^{\prime}_{n},\bm{s}_{-n},\bm{v}^% {n}|h).< roman_sup start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_italic_s start_POSTSUBSCRIPT - italic_n end_POSTSUBSCRIPT , bold_italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT | italic_h ) .

The first inequality is identical to the definition of the obvious dominance relation in Li, (2017), i.e., the best outcome under snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is weakly worse than the worst outcome under snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. In addition, we require the dominated strategy to be non-equivalent in terms of the induced outcome to the strategy that dominates it. The second requirement guarantees that the set of non-obviously dominated strategies is non-empty. The earlier intuition then translates to:

Lemma 1.

If there exists an information set Innsubscript𝐼𝑛subscript𝑛I_{n}\in\mathcal{I}_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ caligraphic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (with observed prices 𝐩𝐩\bm{p}bold_italic_p) such that sn(In)=subscript𝑠𝑛subscript𝐼𝑛s_{n}(I_{n})=\emptysetitalic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∅ (i.e., bidder n𝑛nitalic_n quits) and

  • 𝒑Psuperscript𝒑𝑃\exists\bm{p}^{\prime}\in P∃ bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_P, 𝒃~argmax𝒃()b𝒃pb~𝒃subscript𝒃subscriptsuperscript𝑏𝒃subscriptsuperscript𝑝superscript𝑏\tilde{\bm{b}}\in\arg\max\limits_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b^% {\prime}\in\bm{b}}\,p^{\prime}_{b^{\prime}}over~ start_ARG bold_italic_b end_ARG ∈ roman_arg roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT and b𝒃~𝑏~𝒃\displaystyle b\in\tilde{\bm{b}}italic_b ∈ over~ start_ARG bold_italic_b end_ARG such that 𝒑𝒑superscript𝒑𝒑\bm{p}^{\prime}\geq\bm{p}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≥ bold_italic_p and vbn>pb>pbsubscriptsuperscript𝑣𝑛𝑏subscriptsuperscript𝑝𝑏subscript𝑝𝑏v^{n}_{b}>p^{\prime}_{b}>p_{b}italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT,

then snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is obviously dominated.

  • Proof.

To show that snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is obviously dominated, we explicitly construct another strategy snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT that obviously dominates it. Consider the information set Innsubscript𝐼𝑛subscript𝑛I_{n}\in\mathcal{I}_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ caligraphic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (with observed prices 𝒑𝒑\bm{p}bold_italic_p) such that sn(In)=subscript𝑠𝑛subscript𝐼𝑛s_{n}(I_{n})=\emptysetitalic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∅ and

  • 𝒑Psuperscript𝒑𝑃\exists\bm{p}^{\prime}\in P∃ bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_P, 𝒃~argmax𝒃()b𝒃pb~𝒃subscript𝒃subscriptsuperscript𝑏𝒃subscriptsuperscript𝑝superscript𝑏\tilde{\bm{b}}\in\arg\max\limits_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b^% {\prime}\in\bm{b}}\,p^{\prime}_{b^{\prime}}over~ start_ARG bold_italic_b end_ARG ∈ roman_arg roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT and b𝒃~𝑏~𝒃\displaystyle b\in\tilde{\bm{b}}italic_b ∈ over~ start_ARG bold_italic_b end_ARG such that 𝒑𝒑superscript𝒑𝒑\bm{p}^{\prime}\geq\bm{p}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≥ bold_italic_p and vbn>pb>pbsubscriptsuperscript𝑣𝑛𝑏subscriptsuperscript𝑝𝑏subscript𝑝𝑏v^{n}_{b}>p^{\prime}_{b}>p_{b}italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT.

Obviously, the payoff from snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT conditional on any history hInsubscript𝐼𝑛h\in I_{n}italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is zero. Let snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT be the same as snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT before the information set Insubscript𝐼𝑛I_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and let bidder n𝑛nitalic_n bid sn(In)=(b,pb)superscriptsubscript𝑠𝑛subscript𝐼𝑛𝑏subscriptsuperscript𝑝𝑏s_{n}^{\prime}(I_{n})=(b,p^{\prime}_{b})italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ( italic_b , italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) and never revise her bid afterwards.

We first discuss a special case in which |𝒩t1|=1superscript𝒩𝑡11|\mathcal{N}^{t-1}|=1| caligraphic_N start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT | = 1 and n𝑛nitalic_n is not a current leading bidder (otherwise quitting the auction is not feasible for bidder n𝑛nitalic_n). Then, all the current prices must be 00 and all the other bidders have quit (as this is the unique consistent history). Following the strategy snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, the auction ends with bidder n𝑛nitalic_n bidding (b,pb)𝑏subscriptsuperscript𝑝𝑏(b,p^{\prime}_{b})( italic_b , italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ), leading to a positive payoff of vbnpb>0superscriptsubscript𝑣𝑏𝑛superscriptsubscript𝑝𝑏0v_{b}^{n}-p_{b}^{\prime}>0italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0 for bidder n𝑛nitalic_n.

Next, we consider the case in which |𝒩t1|>1superscript𝒩𝑡11|\mathcal{N}^{t-1}|>1| caligraphic_N start_POSTSUPERSCRIPT italic_t - 1 end_POSTSUPERSCRIPT | > 1. It is clear that bidder n𝑛nitalic_n will get a nonnegative payoff regardless of the value profiles and bidding strategies of other bidders. We show that the best possible payoff for bidder n𝑛nitalic_n following the strategy snsuperscriptsubscript𝑠𝑛s_{n}^{\prime}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is positive. It suffices to consider the case in which any other subsequent active bidder bids pbsubscriptsuperscript𝑝superscript𝑏p^{\prime}_{b^{\prime}}italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT for some bundle bsuperscript𝑏b^{\prime}italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT whenever possible, the auction ends with prices 𝒑superscript𝒑\bm{p}^{\prime}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, leading to a positive payoff of vbnpb>0superscriptsubscript𝑣𝑏𝑛superscriptsubscript𝑝𝑏0v_{b}^{n}-p_{b}^{\prime}>0italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0 for bidder n𝑛nitalic_n. ∎

Note that Lemma 1 suggests that our intuition is incomplete as we cannot fully rule out strategies that quit when the bidder’s value for some bundle is above the leading price for that bundle. To rule out all such strategies, it further requires the existence of some scenario under which bidder n𝑛nitalic_n may be pivotal: a strictly profitable bid of n𝑛nitalic_n may ever be selected in the end. Nevertheless, we show below that the non-pivotal bidders are inconsequential for our analysis.

Let SNODn(P)subscriptsuperscript𝑆𝑛𝑁𝑂𝐷𝑃S^{n}_{NOD}(P)italic_S start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N italic_O italic_D end_POSTSUBSCRIPT ( italic_P ) denote the set of non-obviously dominated strategies of n𝑛nitalic_n given price grid P𝑃Pitalic_P. Let R(𝒔,𝒗)𝑅𝒔𝒗R(\bm{s},\bm{v})italic_R ( bold_italic_s , bold_italic_v ) denote the revenue to seller given value profile 𝒗𝒗\bm{v}bold_italic_v and strategy profile 𝒔𝒔\bm{s}bold_italic_s. Define

R¯CASA(𝒗):=lim¯ϵ0infsnSNODn(P)R(𝒔,𝒗),assignsubscript¯𝑅𝐶𝐴𝑆𝐴𝒗subscriptlimit-infimumitalic-ϵ0subscriptinfimumsubscript𝑠𝑛subscriptsuperscript𝑆𝑛𝑁𝑂𝐷𝑃𝑅𝒔𝒗\displaystyle\underline{R}_{CASA}(\bm{v}):=\varliminf_{\epsilon\to 0}\,\inf_{s% _{n}\in S^{n}_{NOD}(P)}\,R(\bm{s},\bm{v}),under¯ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_C italic_A italic_S italic_A end_POSTSUBSCRIPT ( bold_italic_v ) := start_LIMITOP under¯ start_ARG roman_lim end_ARG end_LIMITOP start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT roman_inf start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ italic_S start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N italic_O italic_D end_POSTSUBSCRIPT ( italic_P ) end_POSTSUBSCRIPT italic_R ( bold_italic_s , bold_italic_v ) ,

where the first limit inf is taken over ϵ0italic-ϵ0\epsilon\to 0italic_ϵ → 0 and all P𝑃Pitalic_P with grid size ϵitalic-ϵ\epsilonitalic_ϵ. That is, R¯CASA(𝒗)subscript¯𝑅𝐶𝐴𝑆𝐴𝒗\underline{R}_{CASA}(\bm{v})under¯ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_C italic_A italic_S italic_A end_POSTSUBSCRIPT ( bold_italic_v ) is the worst-case ex-post revenue from CASA under non-obviously dominated strategies in the limit where grid P𝑃Pitalic_P becomes dense.

Given the valuation profile 𝒗𝒗\bm{v}bold_italic_v and menu \mathcal{M}caligraphic_M, the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee is defined as the maximal revenue from feasible allocations within \mathcal{M}caligraphic_M, taking the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest valuations for each bundle as the price.

Definition 2.

The kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee given the menu \mathcal{M}caligraphic_M and the value profile 𝐯𝐯\bm{v}bold_italic_v is

Rk(𝒗):=max𝒃()b𝒃vb(k),assignsubscriptsuperscript𝑅𝑘𝒗subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣𝑘𝑏\displaystyle R^{k}_{\mathcal{M}}(\bm{v}):=\max_{\bm{b}\in\mathcal{B}(\mathcal% {M})}\,\sum_{b\in\bm{b}}\,v^{(k)}_{b},italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) := roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ,

where vb(k)subscriptsuperscript𝑣𝑘𝑏v^{(k)}_{b}italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT denotes the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest value of bundle b𝑏bitalic_b.

Our key observation is that CASA achieves the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee as long as bidders avoid strategies that are obviously dominated. In other words, for CASA to achieve the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee, we only need minimal rationality on the part of the bidders.

Theorem 1.

R¯CASA(𝒗)Rk(𝒗)subscript¯𝑅𝐶𝐴𝑆𝐴𝒗subscriptsuperscript𝑅𝑘𝒗\displaystyle\underline{R}_{CASA}(\bm{v})\geq{R}^{k}_{\mathcal{M}}(\bm{v})under¯ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_C italic_A italic_S italic_A end_POSTSUBSCRIPT ( bold_italic_v ) ≥ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) for k=||+1𝑘1k=|\mathcal{M}|+1italic_k = | caligraphic_M | + 1.

  • Proof.

Consider a grid P𝑃Pitalic_P with any grid size ϵ>0italic-ϵ0\epsilon>0italic_ϵ > 0. We show that as long as bidders avoid obviously dominated strategies, for any value profile 𝒗𝒗\bm{v}bold_italic_v,

max𝒃()b𝒃pbTmax𝒃()b𝒃(vb(k)ϵ).subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑝𝑇𝑏subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣𝑘𝑏italic-ϵ\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,p^{T% }_{b}\geq\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,(v^{(k)}% _{b}-\epsilon).roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≥ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT ( italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_ϵ ) .

Suppose to the contrary, this is not true. Then, there exists strategy profile s𝑠sitalic_s (with snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT being not obviously dominated for each bidder n𝑛nitalic_n), value profile 𝒗𝒗\bm{v}bold_italic_v, and δ>0𝛿0\delta>0italic_δ > 0 such that

max𝒃()b𝒃pbT<max𝒃()b𝒃(vb(k)ϵδ).subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑝𝑇𝑏subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣𝑘𝑏italic-ϵ𝛿\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,p^{T% }_{b}<\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,(v^{(k)}_{b% }-\epsilon-\delta).roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT < roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT ( italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_ϵ - italic_δ ) .

Evidently, there exists some bundle b𝑏bitalic_b such that pbT<vb(k)ϵδsuperscriptsubscript𝑝𝑏𝑇superscriptsubscript𝑣𝑏𝑘italic-ϵ𝛿p_{b}^{T}<v_{b}^{(k)}-\epsilon-\deltaitalic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT < italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT - italic_ϵ - italic_δ, or equivalently, pbT+ϵ<vb(k)δsuperscriptsubscript𝑝𝑏𝑇italic-ϵsuperscriptsubscript𝑣𝑏𝑘𝛿p_{b}^{T}+\epsilon<v_{b}^{(k)}-\deltaitalic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT + italic_ϵ < italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT - italic_δ. For each such bundle b𝑏bitalic_b, raise the price of bundle b𝑏bitalic_b to (the closest price below) vb(k)δsuperscriptsubscript𝑣𝑏𝑘𝛿v_{b}^{(k)}-\deltaitalic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT - italic_δ sequentially until we find the first pivotal bundle b~~𝑏\tilde{b}over~ start_ARG italic_b end_ARG when max𝒃()b𝒃pbsubscript𝒃subscript𝑏𝒃subscript𝑝𝑏\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\sum_{b\in\bm{b}}p_{b}roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is strictly improved by such increase in prices.

Since k=||+1𝑘1k=|\mathcal{M}|+1italic_k = | caligraphic_M | + 1 and pb~T<vb~(k)ϵδsuperscriptsubscript𝑝~𝑏𝑇superscriptsubscript𝑣~𝑏𝑘italic-ϵ𝛿p_{\tilde{b}}^{T}<v_{\tilde{b}}^{(k)}-\epsilon-\deltaitalic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT < italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT - italic_ϵ - italic_δ, there exists a bidder n𝑛nitalic_n with vb~nvb~(k)superscriptsubscript𝑣~𝑏𝑛superscriptsubscript𝑣~𝑏𝑘v_{\tilde{b}}^{n}\geq v_{\tilde{b}}^{(k)}italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ≥ italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT that quits the auction before the auction ends. Let hhitalic_h be the (on-path) history at which n𝑛nitalic_n quits (note that at this history, n𝑛nitalic_n must not be a leading bidder and pb~(h)pb~tvb~nδϵsubscript𝑝~𝑏superscriptsubscript𝑝~𝑏𝑡superscriptsubscript𝑣~𝑏𝑛𝛿italic-ϵp_{\tilde{b}}(h)\leq p_{\tilde{b}}^{t}\leq v_{\tilde{b}}^{n}-\delta-\epsilonitalic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT ( italic_h ) ≤ italic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ≤ italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - italic_δ - italic_ϵ, where pb~(h)subscript𝑝~𝑏p_{\tilde{b}}(h)italic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT ( italic_h ) is the price of bundle b~~𝑏\tilde{b}over~ start_ARG italic_b end_ARG at the history hhitalic_h). Let hInsubscript𝐼𝑛h\in I_{n}italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. Then, sn(In)=subscript𝑠𝑛subscript𝐼𝑛s_{n}(I_{n})=\emptysetitalic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∅. Let 𝒑superscript𝒑\bm{p}^{\prime}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT be the raised prices and 𝒃~argmax𝒃()b𝒃pb~𝒃subscript𝒃subscriptsuperscript𝑏𝒃subscriptsuperscript𝑝superscript𝑏\tilde{\bm{b}}\in\arg\max\limits_{\bm{b}\in\mathcal{B}(\mathcal{M})}\sum_{b^{% \prime}\in\bm{b}}p^{\prime}_{b^{\prime}}over~ start_ARG bold_italic_b end_ARG ∈ roman_arg roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ bold_italic_b end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT. Then b~𝒃~~𝑏~𝒃\tilde{b}\in\tilde{\bm{b}}over~ start_ARG italic_b end_ARG ∈ over~ start_ARG bold_italic_b end_ARG, and 𝒑superscript𝒑\bm{p}^{\prime}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT satisfies the condition in Lemma 1. Therefore, snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is obviously dominated. We arrive at a contradiction. ∎

As we have pointed out following Lemma 1, elimination of obviously dominated strategies does not guarantee the ex-post prices to be above the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-highest values. To establish Theorem 1, we prove in addition that the behaviors of non-pivotal bidders are inconsequential. With Theorem 1, we say that CASA is a rank-guaranteed auction format as long as we are comfortable assuming that bidders avoid strategies that are obviously dominated.

CASA and its rank-guarantee become extremely simple when ||=11|\mathcal{M}|=1| caligraphic_M | = 1, where CASA reduces to the canonical English auction of the unique feasible bundle and the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee becomes the second highest value. In this special case, Theorem 1 shares the insight from Li, (2017) that English auction achieves the 2ndsuperscript2𝑛𝑑2^{nd}2 start_POSTSUPERSCRIPT italic_n italic_d end_POSTSUPERSCRIPT-guarantee via an obviously strategy-proof outcome. To understand our general result, we provide an alternative interpretation of this special case. The English auction made it obvious that a bidder should avoid losing the auction when there is a remaining surplus. Therefore, it maximally utilizes the competition among losers of the auction, pushing the price to the highest value among them. Meanwhile, it does not further screen the winner at all. CASA is exactly the multi-item generalization of this philosophy: it maximally invokes competition among losers by making it obviously suboptimal to lose with a remaining surplus. Meanwhile, it does not “care” at all which winner gets which bundle and how much extra he pays. As a result, the auctioneer extracts at least as much surplus as that from the losers only, which is exactly the rank-guarantee.

2.4 Discussions

Alternative formats:

As we have discussed above, CASA is an intuitive extension of the canonical single-item English auction. The idea of generalizing the English auction to accommodate multiple items has been extensively explored. However, crucial to the design of CASA is a set of unique properties that cope with the incentives of the players. In comparison to the simultaneous ascending auction (SAA, Milgrom, (2000)), allowing for bidding on bundles (package bidding) has the advantage of mitigating the demand reduction problem or the exposure problem, leading to sufficient competition. Compared to the combinatorial variants of SAA like SAAPB (Ausubel and Milgrom, (2002)) and CCA (Ausubel et al., (2006)), CASA uses a simpler ”pay-as-bid” rule so that the bidders find it straightforward to determine the remaining surplus from the auction, leading to the obvious strategy-proof implementation.444The SAAPB is indeed kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guaranteed under non-strategic “straightforward bidding” strategies (Theorem 1 of Ausubel and Milgrom, (2002)). However, whether SAAPB is kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guaranteed with fully strategic bidders is yet unknown to us. Compared to the majority of iterative combinatorial auctions (Chapter 2, Cramton et al., (2006)), which seek to replicate VCG (known not to be rank-guaranteed555Imagine the case ={(a),(b),(a,b)}𝑎𝑏𝑎𝑏\mathcal{M}=\{(a),(b),(a,b)\}caligraphic_M = { ( italic_a ) , ( italic_b ) , ( italic_a , italic_b ) }. v(a,b)=1subscript𝑣𝑎𝑏1v_{(a,b)}=1italic_v start_POSTSUBSCRIPT ( italic_a , italic_b ) end_POSTSUBSCRIPT = 1 for all bidders. va=vb=0subscript𝑣𝑎subscript𝑣𝑏0v_{a}=v_{b}=0italic_v start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 0 for all bidders except for two, whose value for a𝑎aitalic_a and b𝑏bitalic_b are 1111. The VCG revenue is 00, while the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guarantee is 1111 for any k2𝑘2k\geq 2italic_k ≥ 2.) and maximize efficiency, CASA has much better revenue performance. The design of CASA accommodates fully strategic bidders, as opposed to single-minded or myopic bidders assumed in the cited papers.

Except for the several crucial features, the exact format of CASA can be flexibly tailored to the auctioneer’s needs and practical considerations without losing its various desirable properties. Notably, the menu \mathcal{M}caligraphic_M is a design variable of the auctioneer, which we examine in detail in Section 4. Additionally, instead of allowing bidders to bid on multiple bundles, we could restrict them to bid on a single bundle when it is their turn to move (if a bidder is already leading for some bundle when it is her turn to move, she could not bid on any other bundle but she would remain an active bidder). Such restriction prevents the type of collusive communication documented in Grimm et al., (2003); Jehiel and Moldovanu, 2001b . Moreover, we could also allow the bidders to observe the entire history if that is considered desirable for transparency purposes.

Robustness to collusion / irrationality:

Klemperer, (2002) identifies collusion as the first concern that “really matters in auction design”. We illustrate below that our framework allows us to quantify the impact of collusive or irrational behavior on the performance of CASA. Recall that the key force that drives the revenue to the rank-guarantee is the competition among the losing bidders. Therefore, when there are non-strategic bidders, one can simply consider the competition among the subset of rational losing bidders.

Formally, when the number of non-strategic bidders is bounded, then Theorem 1 still holds when k𝑘kitalic_k is relaxed by the number of non-strategic bidders.

Proposition 1.

Suppose there are j𝑗jitalic_j non-strategic bidders, then R¯CASA(𝐯)Rk(𝐯)subscript¯𝑅𝐶𝐴𝑆𝐴𝐯superscriptsubscript𝑅𝑘𝐯\underline{R}_{CASA}(\bm{v})\geq R_{\mathcal{M}}^{k}(\bm{v})under¯ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_C italic_A italic_S italic_A end_POSTSUBSCRIPT ( bold_italic_v ) ≥ italic_R start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( bold_italic_v ) for k=||+1+j𝑘1𝑗k=|\mathcal{M}|+1+jitalic_k = | caligraphic_M | + 1 + italic_j.

  • Proof.

    Observe that in the proof of Theorem 1, since k=||+1+j𝑘1𝑗k=|\mathcal{M}|+1+jitalic_k = | caligraphic_M | + 1 + italic_j, there exists at least one strategic player n𝑛nitalic_n that quits the auction before period t𝑡titalic_t and vb~nvb~(k)superscriptsubscript𝑣~𝑏𝑛superscriptsubscript𝑣~𝑏𝑘v_{\tilde{b}}^{n}\geq v_{\tilde{b}}^{(k)}italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ≥ italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT. The rest of the proof follows. ∎

Proposition 2 applies when a relatively small number of bidders play collusive/irrational strategies. Then, they reduce the rank of the guarantee by a small amount. Nevertheless, CASA is still rank-guaranteed.

What if all bidders are collusive? When the bidders form coalitions, and they strategically maximize group-level payoffs666Each group can freely shift allocations within the group and maximize the total payoff., Theorem 1 still holds when k𝑘kitalic_k is scaled by the coalition sizes.

Proposition 2.

Suppose bidders are partitioned into strategic coalitions {ci}iIsubscriptsubscript𝑐𝑖𝑖𝐼\{c_{i}\}_{i\in I}{ italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT, where the index is chosen such that |ci|subscript𝑐𝑖|c_{i}|| italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | decreases in i𝑖iitalic_i. Then, R¯CASA(𝐯)Rk(𝐯)subscript¯𝑅𝐶𝐴𝑆𝐴𝐯superscriptsubscript𝑅𝑘𝐯\underline{R}_{CASA}(\bm{v})\geq R_{\mathcal{M}}^{k}(\bm{v})under¯ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_C italic_A italic_S italic_A end_POSTSUBSCRIPT ( bold_italic_v ) ≥ italic_R start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ( bold_italic_v ) for k=i|||ci|+1𝑘subscript𝑖subscript𝑐𝑖1k=\sum_{i\leq|\mathcal{M}|}|c_{i}|+1italic_k = ∑ start_POSTSUBSCRIPT italic_i ≤ | caligraphic_M | end_POSTSUBSCRIPT | italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | + 1.

  • Proof.

    Observe that in the proof of Theorem 1, since k=i|||ci|+1𝑘subscript𝑖subscript𝑐𝑖1k=\sum_{i\leq|\mathcal{M}|}|c_{i}|+1italic_k = ∑ start_POSTSUBSCRIPT italic_i ≤ | caligraphic_M | end_POSTSUBSCRIPT | italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | + 1, there exists at least one coalition of players c𝑐citalic_c that all quit the auction before period t𝑡titalic_t and maxnc{vb~n}vb~(k)subscript𝑛𝑐superscriptsubscript𝑣~𝑏𝑛superscriptsubscript𝑣~𝑏𝑘\max_{n\in c}\{v_{\tilde{b}}^{n}\}\geq v_{\tilde{b}}^{(k)}roman_max start_POSTSUBSCRIPT italic_n ∈ italic_c end_POSTSUBSCRIPT { italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT } ≥ italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT. Let hhitalic_h be the (on-path) history at which the last member n𝑛nitalic_n in the coalition quits (note that at this history, n𝑛nitalic_n must not be a leading bidder and pb~pb~tmaxn{vb~n}δϵsubscript𝑝~𝑏superscriptsubscript𝑝~𝑏𝑡subscript𝑛superscriptsubscript𝑣~𝑏𝑛𝛿italic-ϵp_{\tilde{b}}\leq p_{\tilde{b}}^{t}\leq\max_{n}\{v_{\tilde{b}}^{n}\}-\delta-\epsilonitalic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT ≤ italic_p start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ≤ roman_max start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT { italic_v start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT } - italic_δ - italic_ϵ). Let hInsubscript𝐼𝑛h\in I_{n}italic_h ∈ italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. Then, sn(In)=subscript𝑠𝑛subscript𝐼𝑛s_{n}(I_{n})=\emptysetitalic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∅. Obviously, quitting gives the entire group zero payoff while bidding pb~subscriptsuperscript𝑝~𝑏p^{\prime}_{\tilde{b}}italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT over~ start_ARG italic_b end_ARG end_POSTSUBSCRIPT guarantees a non-negative payoff. Suppose all other bidders bid up to pbsuperscriptsubscript𝑝𝑏p_{b}^{\prime}italic_p start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT when it is their turn, the auction ends with 𝒑superscript𝒑\bm{p}^{\prime}bold_italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and the group obtains a strictly positive payoff. Therefore, snsubscript𝑠𝑛s_{n}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is obviously dominated for coalition c𝑐citalic_c. ∎

The intuition behind the two extensions is exactly the competition among losing strategic bidders. The price of each bundle must be higher than the value of any losing strategic bidder or any losing coalition group as otherwise they will outbid the price. Of course, Proposition 2 has no bite when k𝑘kitalic_k is large compared to N𝑁Nitalic_N; hence, it should be interpreted as the strategic robustness of CASA only in relatively thick markets. Nevertheless, CASA is also aligned with the philosophy of anti-collusion design, even in thin markets, because CASA permits the minimum transmission of information. For example, anonymity prevents the reciprocity behavior documented in Cramton and Schwartz, (2000).

Efficiency:

In CASA, although bidders can fully avoid the exposure problem by bidding only on bundles with a positive surplus, they may strategically expose themselves. It is not obviously dominated to bid strictly above the true valuation for a bundle.777Consider, for example, the case with three bidders 1,2,31231,2,31 , 2 , 3 and three items a,b,c𝑎𝑏𝑐a,b,citalic_a , italic_b , italic_c . Bidder 1111 only wants a𝑎aitalic_a, bidder 2222 only wants b𝑏bitalic_b, and bidder 3333 only wants the grand bundle {a,b,c}𝑎𝑏𝑐\{a,b,c\}{ italic_a , italic_b , italic_c }. The valuation of each bidder for the desired bundle is 1111. By strategically bidding up item b𝑏bitalic_b even though bidder 1111 gets zero value from it, bidder 1111 can reduce the bid required for him to win item a𝑎aitalic_a, creating an exposure problem for 1111. Therefore, CASA might not satisfy ex-post IR; hence kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guaranteed revenue does not implie kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT-guaranteed surplus. Of course, CASA still satisfies ex-ante IR (assuming bidders have correct Bayesian priors) since quitting at the beginning is always an option; hence, the ex-ante bounds we derive in the following sections on the revenue of CASA also apply to surplus.

3 Rank-guarantee as a desideratum

Section 2 shows that CASA achieves the rank-guarantee as long as bidders avoid strategies that are obviously dominated. In this section, we explore the concept of rank-guarantee as a desideratum in auction design.

Clearly, if the auctioneer knows that the bidders’ valuations are independent and identically distributed, then rank-guarantee is an appealing approximation when N𝑁Nitalic_N is large, as all order statistics converge to the upper bound of the valuation support (at the rate of 1N1𝑁\frac{1}{N}divide start_ARG 1 end_ARG start_ARG italic_N end_ARG). In what follows, we show that even when the auctioneer has non-Bayesian uncertainty about the joint distribution of the bidders’ valuations and maximizes the revenue-guarantee (the worst-case expected revenue where the worst case is taken over all joint distributions that are perceived to be plausible), in many settings, rank-guarantee remains an appealing approximation.

Let 𝔾Δ([v¯,v¯]2S)𝔾Δsuperscript¯𝑣¯𝑣superscript2𝑆\mathbb{G}\subset\Delta([\underline{v},\mkern 1.5mu\overline{\mkern-1.5muv% \mkern-1.5mu}\mkern 1.5mu]^{2^{S}})blackboard_G ⊂ roman_Δ ( [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] start_POSTSUPERSCRIPT 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) be an arbitrary subset of distributions of valuation vector. We interpret 𝔾𝔾\mathbb{G}blackboard_G as the auctioneer’s estimate of a representative bidder’s valuation. Then, the joint distributions of the bidders’ valuations that are considered possible by the auctioneer are

𝔽={FΔ([v¯,v¯]N×2S)|1NFn𝔾},𝔽conditional-set𝐹Δsuperscript¯𝑣¯𝑣𝑁superscript2𝑆1𝑁subscript𝐹𝑛𝔾\displaystyle\mathbb{F}=\Big{\{}F\in\Delta([\underline{v},\mkern 1.5mu% \overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu]^{N\times 2^{S}})\,\Big{|}\,% \frac{1}{N}\sum F_{n}\in\mathbb{G}\Big{\}},blackboard_F = { italic_F ∈ roman_Δ ( [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] start_POSTSUPERSCRIPT italic_N × 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) | divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_G } ,

where Fnsubscript𝐹𝑛F_{n}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the marginal distribution of bidder n𝑛nitalic_n’s valuation. Thus, 1NFn1𝑁subscript𝐹𝑛\frac{1}{N}\sum F_{n}divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the cumulative distribution function of the valuation of a uniformly randomly selected bidder in the population. We call 𝔽𝔽\mathbb{F}blackboard_F an ambiguity set. Such an ambiguity set 𝔽𝔽\mathbb{F}blackboard_F could come from the statistical estimation of F𝐹Fitalic_F based on a “sanitized” dataset about valuations, that is, past bidders’ valuations with identity information removed. The ambiguity set 𝔽𝔽\mathbb{F}blackboard_F captures the type of distributional uncertainty introduced by Carroll, (2017), while further generalizing it to capture realistic knowledge structures stemming from statistical inference.888This setup covers a wide range of scenarios, as 𝔾𝔾\mathbb{G}blackboard_G is completely general. 𝔾𝔾\mathbb{G}blackboard_G could be a singleton set capturing the case in which the auctioneer has no uncertainty about the distribution of a representative bidder’s valuation. 𝔾𝔾\mathbb{G}blackboard_G could also be the set of distributions satisfying certain statistical properties (say moment conditions), capturing scenarios in which the auctioneer also has some non-Bayesian uncertainty about the distribution of a representative bidder’s valuation.

For any FΔ([v¯,v¯]N×2S)𝐹Δsuperscript¯𝑣¯𝑣𝑁superscript2𝑆F\in\Delta([\underline{v},\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}% \mkern 1.5mu]^{N\times 2^{S}})italic_F ∈ roman_Δ ( [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] start_POSTSUPERSCRIPT italic_N × 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) and menu 2Ssuperscript2𝑆\mathcal{M}\subseteq 2^{S}caligraphic_M ⊆ 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT, define the ex-ante efficient surplus

V(F):=𝔼F[max𝒃()maxι:𝒃𝒩b𝒃vbι(b)],assignsubscript𝑉𝐹subscript𝔼𝐹delimited-[]subscript𝒃subscript:𝜄𝒃𝒩subscript𝑏𝒃subscriptsuperscript𝑣𝜄𝑏𝑏\displaystyle V_{\mathcal{M}}(F):=\mathbb{E}_{F}\left[\max_{\bm{b}\in\mathcal{% B}(\mathcal{M})}\,\max_{\iota:\bm{b}\rightarrow\mathcal{N}}\,\sum_{b\in\bm{b}}% \,v^{\iota(b)}_{b}\right],italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) := blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT roman_max start_POSTSUBSCRIPT italic_ι : bold_italic_b → caligraphic_N end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_ι ( italic_b ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] ,

where ι𝜄\iotaitalic_ι, the assignment function satisfies ι(b)ι(b)𝜄𝑏𝜄superscript𝑏\iota(b)\neq\iota(b^{\prime})italic_ι ( italic_b ) ≠ italic_ι ( italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) for any bb𝑏superscript𝑏b\neq b^{\prime}italic_b ≠ italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. The ex-ante efficient surplus V2S(F)subscript𝑉superscript2𝑆𝐹V_{2^{S}}(F)italic_V start_POSTSUBSCRIPT 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_F ) with respect to the complete menu 2Ssuperscript2𝑆2^{S}2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT is denoted by V(F)superscript𝑉𝐹V^{*}(F)italic_V start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_F ) for simplicity.

Theorem 2.

For any menu \mathcal{M}caligraphic_M,

infF𝔽𝔼F[Rk(𝒗)]infF𝔽V(F)(k1)||v¯N.subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗subscriptinfimum𝐹𝔽subscript𝑉𝐹𝑘1¯𝑣𝑁\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}[{R}^{k}_{\mathcal{M}}(\bm{v% })]\geq\inf_{F\in\mathbb{F}}\,V_{\mathcal{M}}(F)-\frac{(k-1)|\mathcal{M}|% \mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}.roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG .
  • Proof.

Let ®®\circledR® be a uniform random element of 𝒩𝒩\mathcal{N}caligraphic_N.

Rk(𝒗)=subscriptsuperscript𝑅𝑘𝒗absent\displaystyle{R}^{k}_{\mathcal{M}}(\bm{v})=italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) = max𝒃()b𝒃vb(k)subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣𝑘𝑏\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v^{(% k)}_{b}roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT
\displaystyle\geq max𝒃()(b𝒃vb®b𝒃max{vb®vb(k),0})subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏0\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\Big{(}\sum_{b\in\bm{b}% }\,v^{\circledR}_{b}-\sum_{b\in\bm{b}}\,\max\,\{v^{\circledR}_{b}-v^{(k)}_{b},% 0\}\Big{)}roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT roman_max { italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , 0 } )
\displaystyle\geq max𝒃()(b𝒃vb®)bmax{vb®vb(k),0}subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscript𝑏subscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏0\displaystyle\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\Big{(}\sum_{b\in\bm{b}% }\,v^{\circledR}_{b}\Big{)}-\sum_{b\in\mathcal{M}}\,\max\,\{v^{\circledR}_{b}-% v^{(k)}_{b},0\}roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ( ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) - ∑ start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT roman_max { italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , 0 }
𝔼F[Rk(𝒗)]absentsubscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗absent\displaystyle\implies\mathbb{E}_{F}[{R}^{k}_{\mathcal{M}}(\bm{v})]\geq⟹ blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ 𝔼F[max𝒃()b𝒃vb®]b𝔼F[vb®vb(k)|vb®>vb(k)] Prob(vb®>vb(k))subscript𝔼𝐹delimited-[]subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscript𝑏subscript𝔼𝐹delimited-[]subscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏ketsubscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏 Probsubscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏\displaystyle\mathbb{E}_{F}\left[\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,% \sum_{b\in\bm{b}}\,v^{\circledR}_{b}\right]-\sum_{b\in\mathcal{M}}\,\mathbb{E}% _{F}\left[v^{\circledR}_{b}-v^{(k)}_{b}|v^{\circledR}_{b}>v^{(k)}_{b}\right]% \text{ Prob}(v^{\circledR}_{b}>v^{(k)}_{b})blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - ∑ start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] Prob ( italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT )
\displaystyle\geq 𝔼F[max𝒃()b𝒃vb®]bv¯ Prob(vb®>vb(k))subscript𝔼𝐹delimited-[]subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscript𝑏¯𝑣 Probsubscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏\displaystyle\mathbb{E}_{F}\left[\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,% \sum_{b\in\bm{b}}\,v^{\circledR}_{b}\right]-\sum_{b\in\mathcal{M}}\,\mkern 1.5% mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu\text{ Prob}(v^{\circledR}_{% b}>v^{(k)}_{b})blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - ∑ start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG Prob ( italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT )
\displaystyle\geq 𝔼F[max𝒃()b𝒃vb®](k1)||v¯N.subscript𝔼𝐹delimited-[]subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏𝑘1¯𝑣𝑁\displaystyle\mathbb{E}_{F}\left[\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,% \sum_{b\in\bm{b}}\,v^{\circledR}_{b}\right]-\frac{(k-1)|\mathcal{M}|\mkern 1.5% mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}.blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG .

This further implies that

infF𝔽𝔼F[Rk(𝒗)]subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗absent\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}\left[{R}^{k}_{\mathcal{M}}(% \bm{v})\right]\geqroman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ infF𝔽𝔼F[max𝒃()b𝒃vb®](k1)||v¯Nsubscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏𝑘1¯𝑣𝑁\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}\left[\max_{\bm{b}\in% \mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v^{\circledR}_{b}\right]-\frac{(% k-1)|\mathcal{M}|\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}% {N}roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG
=\displaystyle== infG𝔾𝔼G[max𝒃()b𝒃vb](k1)||v¯Nsubscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃subscript𝑏𝒃subscript𝑣𝑏𝑘1¯𝑣𝑁\displaystyle\inf_{G\in\mathbb{G}}\,\mathbb{E}_{G}\left[\max_{\bm{b}\in% \mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v_{b}\right]-\frac{(k-1)|% \mathcal{M}|\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG
\displaystyle\geq infF𝔽V(F)(k1)||v¯N,subscriptinfimum𝐹𝔽subscript𝑉𝐹𝑘1¯𝑣𝑁\displaystyle\inf_{F\in\mathbb{F}}\,V_{\mathcal{M}}(F)-\frac{(k-1)|\mathcal{M}% |\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N},roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG ,

where the equality follows from the definition of the set 𝔽𝔽\mathbb{F}blackboard_F. For the last inequality, observe that for any G𝔾𝐺𝔾G\in\mathbb{G}italic_G ∈ blackboard_G, the joint distribution where each bidder’s value is distributed according to G𝐺Gitalic_G and all bidders’ values are maximally positively correlated is contained in 𝔽𝔽\mathbb{F}blackboard_F. Therefore,

infF𝔽V(F)infG𝔾𝔼G[max𝒃()b𝒃vb].subscriptinfimum𝐹𝔽subscript𝑉𝐹subscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃subscript𝑏𝒃subscript𝑣𝑏\inf_{F\in\mathbb{F}}\,V_{\mathcal{M}}(F)\leq\inf_{G\in\mathbb{G}}\mathbb{E}_{% G}\left[\max_{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v_{b}% \right].roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) ≤ roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] .

Theorem 2 highlights a key trade-off between menu sufficiency and approximation efficiency. Evidently, presenting the bidders with a larger menu has the potentially of increasing allocation efficiency. Particularly, \mathcal{M}caligraphic_M can be naively chosen to be the complete menu 2Ssuperscript2𝑆2^{S}2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT to guarantee full menu sufficiency, i.e., V(F)=V(F)subscript𝑉𝐹superscript𝑉𝐹V_{\mathcal{M}}(F)=V^{*}(F)italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) = italic_V start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_F ). However, this leads to |||\mathcal{M}|| caligraphic_M | growing exponentially in M𝑀Mitalic_M, causing both complex auction process and slow convergence. On the other hand, choosing a small menu achieves approximation efficiency but sacrifices allocation efficiency. Although such trade-off is generally non-trivial under general combinatorial preferences, we show in the next section that under canonical preference structures, menu sufficiency and approximation efficiency can often be achieved simultaneously.

3.1 Discussions

Alternative ambiguity sets:

The proof of Theorem 2 goes through for a general 𝔽𝔽\mathbb{F}blackboard_F if the following equality holds.

infF𝔽𝔼F[max𝒃()b𝒃vb®]=infF𝔽V(F).subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscript𝒃subscript𝑏𝒃subscriptsuperscript𝑣®𝑏subscriptinfimum𝐹𝔽subscript𝑉𝐹\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}\left[\max_{\bm{b}\in% \mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v^{\circledR}_{b}\right]=\inf_{F% \in\mathbb{F}}\,V_{\mathcal{M}}(F).roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] = roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) .

That is, in the worst case, randomly selecting bidders performs as well as optimally selecting bidders. Therefore, it is straightforward that Theorem 2 holds under an alternative ambiguity set:

𝔽^={FΔ([v¯,v¯]N×2S)|n,Fn𝔾}.^𝔽conditional-set𝐹Δsuperscript¯𝑣¯𝑣𝑁superscript2𝑆for-all𝑛subscript𝐹𝑛𝔾\displaystyle\widehat{\mathbb{F}}=\left\{F\in\Delta([\underline{v},\mkern 1.5% mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu]^{N\times 2^{S}})\,\Big{|}% \forall n,\ F_{n}\in\mathbb{G}\right\}.over^ start_ARG blackboard_F end_ARG = { italic_F ∈ roman_Δ ( [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] start_POSTSUPERSCRIPT italic_N × 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) | ∀ italic_n , italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_G } .

Under 𝔽^^𝔽\widehat{\mathbb{F}}over^ start_ARG blackboard_F end_ARG, the auctioneer knows that the marginal distributions of bidders are contained in 𝔾𝔾\mathbb{G}blackboard_G, but nothing beyond that. This ambiguity set exactly captures the correlational uncertainty studied in Carroll, (2017), He and Li, (2022), Zhang, (2022), and Suzdaltsev, (2022).

Distributions with unbounded support:

Suppose the support of distributions in 𝔾𝔾\mathbb{G}blackboard_G has unbounded support, i.e., v¯=¯𝑣\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu=\inftyover¯ start_ARG italic_v end_ARG = ∞, then the bound derived in Theorem 2 has no bite. However, observe that

𝔼F[vb®|vb®>vb(k)] Prob(vb®>vb(k))QF¯b(k1)/NvbdF¯(v),subscript𝔼𝐹delimited-[]subscriptsuperscript𝑣®𝑏ketsubscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏 Probsubscriptsuperscript𝑣®𝑏subscriptsuperscript𝑣𝑘𝑏superscriptsubscriptsuperscriptsubscript𝑄subscript¯𝐹𝑏𝑘1𝑁subscript𝑣𝑏differential-d¯𝐹𝑣\displaystyle\mathbb{E}_{F}\left[v^{\circledR}_{b}|v^{\circledR}_{b}>v^{(k)}_{% b}\right]\text{ Prob}(v^{\circledR}_{b}>v^{(k)}_{b})\leq\int_{Q_{\bar{F}_{b}}^% {(k-1)/N}}^{\infty}v_{b}\mathrm{d}\bar{F}(v),blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT | italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] Prob ( italic_v start_POSTSUPERSCRIPT ® end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > italic_v start_POSTSUPERSCRIPT ( italic_k ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) ≤ ∫ start_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT roman_d over¯ start_ARG italic_F end_ARG ( italic_v ) ,

where F¯=FnN¯𝐹subscript𝐹𝑛𝑁\bar{F}=\frac{\sum F_{n}}{N}over¯ start_ARG italic_F end_ARG = divide start_ARG ∑ italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_N end_ARG and QF¯b(k1)/Nsuperscriptsubscript𝑄subscript¯𝐹𝑏𝑘1𝑁Q_{\bar{F}_{b}}^{(k-1)/N}italic_Q start_POSTSUBSCRIPT over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT is the top k1N𝑘1𝑁\frac{k-1}{N}divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG quantile of the marginal of F¯¯𝐹\bar{F}over¯ start_ARG italic_F end_ARG for vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. To see why the inequality holds, the LHS is the constrained expectation of vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT given F¯¯𝐹\bar{F}over¯ start_ARG italic_F end_ARG on some event of probability at most k1N𝑘1𝑁\frac{k-1}{N}divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG. The RHS is the constrained expectation of vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT given F¯¯𝐹\bar{F}over¯ start_ARG italic_F end_ARG on the probability-k1N𝑘1𝑁\frac{k-1}{N}divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG event that maximizes its value. Therefore,

infF𝔽𝔼F[Rk(𝒗)]infF𝔽V(F)supG𝔾bQGb(k1)/NvbdG(v).subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗subscriptinfimum𝐹𝔽subscript𝑉𝐹subscriptsupremum𝐺𝔾subscript𝑏superscriptsubscriptsuperscriptsubscript𝑄subscript𝐺𝑏𝑘1𝑁subscript𝑣𝑏differential-d𝐺𝑣\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}[{R}^{k}_{\mathcal{M}}(\bm{v% })]\geq\inf_{F\in\mathbb{F}}\,V_{\mathcal{M}}(F)-\sup_{G\in\mathbb{G}}\sum_{b% \in\mathcal{M}}\int_{Q_{G_{b}}^{(k-1)/N}}^{\infty}v_{b}\mathrm{d}G(v).roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) - roman_sup start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ caligraphic_M end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT roman_d italic_G ( italic_v ) .

For any G𝐺Gitalic_G, the concentration inequality implies QGb(k1)/NvbdG(v)𝔼G[vb2]QGb(k1)/Nsuperscriptsubscriptsuperscriptsubscript𝑄subscript𝐺𝑏𝑘1𝑁subscript𝑣𝑏differential-d𝐺𝑣subscript𝔼𝐺delimited-[]superscriptsubscript𝑣𝑏2superscriptsubscript𝑄subscript𝐺𝑏𝑘1𝑁\int_{Q_{G_{b}}^{(k-1)/N}}^{\infty}v_{b}\mathrm{d}G(v)\leq\frac{\mathbb{E}_{G}% [v_{b}^{2}]}{Q_{G_{b}}^{(k-1)/N}}∫ start_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT roman_d italic_G ( italic_v ) ≤ divide start_ARG blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG start_ARG italic_Q start_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT end_ARG. If G𝐺Gitalic_G has unbounded support, so is the quantile QGb(k1)/Nsuperscriptsubscript𝑄subscript𝐺𝑏𝑘1𝑁Q_{G_{b}}^{(k-1)/N}italic_Q start_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT. As a result, when distributions in 𝔾𝔾\mathbb{G}blackboard_G have unbounded support but uniformly bounded second moment, the worst-case rank-guarantee converges to the full surplus at the rate of O(||infG𝔾QGb(k1)/N)𝑂subscriptinfimum𝐺𝔾superscriptsubscript𝑄subscript𝐺𝑏𝑘1𝑁O\Big{(}\frac{|\mathcal{M}|}{\inf_{G\in\mathbb{G}}Q_{G_{b}}^{(k-1)/N}}\Big{)}italic_O ( divide start_ARG | caligraphic_M | end_ARG start_ARG roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_k - 1 ) / italic_N end_POSTSUPERSCRIPT end_ARG ).

Tightness of the bound:

The coefficient k||𝑘k|\mathcal{M}|italic_k | caligraphic_M | in Theorem 2 consists of two parts. The coefficient k𝑘kitalic_k comes from the kthsuperscript𝑘𝑡k^{th}italic_k start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT highest value approximation. The coefficient |||\mathcal{M}|| caligraphic_M | comes from the total number of bundles in the menu \mathcal{M}caligraphic_M. Proposition 3 below shows that the dependence on k𝑘kitalic_k is tight.

Proposition 3.

For any M,N,,k𝑀𝑁𝑘M,N,\mathcal{M},kitalic_M , italic_N , caligraphic_M , italic_k, there exists some 𝔾𝔾\mathbb{G}blackboard_G such that

infF𝔽𝔼F[Rk(𝒗)]infF𝔽V(F)O(kN).subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗subscriptinfimum𝐹𝔽subscript𝑉𝐹𝑂𝑘𝑁\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}[{R}^{k}_{\mathcal{M}}(\bm{v% })]\leq\inf_{F\in\mathbb{F}}\,V_{\mathcal{M}}(F)-\textstyle O\left(\frac{k}{N}% \right).roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≤ roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) - italic_O ( divide start_ARG italic_k end_ARG start_ARG italic_N end_ARG ) .

The dependence of the approximation gap on |||\mathcal{M}|| caligraphic_M |, however, might not be tight as we take a very loose upper bound in the proof of Theorem 2: we bound the revenue loss from the allocated bundles (up to M𝑀Mitalic_M of them) by the revenue loss from all bundles (|||\mathcal{M}|| caligraphic_M | of them). While we speculate that the coefficient |||\mathcal{M}|| caligraphic_M | can be improved, a formal proof is yet unknown to us.

4 Menu design and simple menus

In this section, we examine several canonical classes of preference structures where there exist menus that are both sufficient, ensuring full allocation efficiency, and simple, with menu size growing at a polynomial rate as M𝑀Mitalic_M increases.

Definition 3.

Menu \mathcal{M}caligraphic_M is 𝔾𝔾\mathbb{G}blackboard_G-sufficient if:

infG𝔾𝔼G[max𝒃()b𝒃vb]=infG𝔾𝔼G[max𝒃(2S)b𝒃vb].subscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃subscript𝑏𝒃subscript𝑣𝑏subscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃superscript2𝑆subscript𝑏𝒃subscript𝑣𝑏\displaystyle\inf_{G\in\mathbb{G}}\,\mathbb{E}_{G}\left[\max_{\bm{b}\in% \mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v_{b}\right]=\inf_{G\in\mathbb{G% }}\,\mathbb{E}_{G}\left[\max_{\bm{b}\in\mathcal{B}(2^{S})}\,\sum_{b\in\bm{b}}% \,v_{b}\right].roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] = roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] .

In words, a menu \mathcal{M}caligraphic_M is 𝔾𝔾\mathbb{G}blackboard_G-sufficient if the worst-case surplus from allocating to (hypothetically) identical bidders with valuation distribution from 𝔾𝔾\mathbb{G}blackboard_G is the same as that under the complete menu 2Ssuperscript2𝑆2^{S}2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT. Importantly, 𝔾𝔾\mathbb{G}blackboard_G-sufficiency is defined with respect to the preference of a single bidder instead of all bidders. It is much weaker than assuming that restricting to allocations within \mathcal{M}caligraphic_M is without loss for ex-post efficiency.999Consider for instance two items and two bidders, where for each bidder the sum of the value for each individual item is more than her value for the grand bundle. Then, menu of individual items is “sufficient” per Definition 3, but not necessarily ex-post efficient when the two bidder’s values are highly asymmetric. Nevertheless, sufficiency guarantees full allocation efficiency:

Theorem 3.

If menu \mathcal{M}caligraphic_M is 𝔾𝔾\mathbb{G}blackboard_G-sufficient, then

infF𝔽𝔼F[Rk(𝒗)]infF𝔽V(F)(k1)||v¯N.subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗subscriptinfimum𝐹𝔽superscript𝑉𝐹𝑘1¯𝑣𝑁\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}\left[{R}^{k}_{\mathcal{M}}(% \bm{v})\right]\geq\inf_{F\in\mathbb{F}}\,V^{*}(F)-\frac{(k-1)|\mathcal{M}|% \mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}.roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_F ) - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG .
  • Proof.
    infF𝔽𝔼F[Rk(𝒗)]subscriptinfimum𝐹𝔽subscript𝔼𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗absent\displaystyle\inf_{F\in\mathbb{F}}\,\mathbb{E}_{F}\left[{R}^{k}_{\mathcal{M}}(% \bm{v})\right]\geqroman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] ≥ infG𝔾𝔼G[max𝒃()b𝒃vb](k1)||v¯Nsubscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃subscript𝑏𝒃subscript𝑣𝑏𝑘1¯𝑣𝑁\displaystyle\inf_{G\in\mathbb{G}}\,\mathbb{E}_{G}\left[\max_{\bm{b}\in% \mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v_{b}\right]-\frac{(k-1)|% \mathcal{M}|\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG
    =\displaystyle== infG𝔾𝔼G[max𝒃(2S)b𝒃vb](k1)||v¯Nsubscriptinfimum𝐺𝔾subscript𝔼𝐺delimited-[]subscript𝒃superscript2𝑆subscript𝑏𝒃subscript𝑣𝑏𝑘1¯𝑣𝑁\displaystyle\inf_{G\in\mathbb{G}}\,\mathbb{E}_{G}\left[\max_{\bm{b}\in% \mathcal{B}(2^{S})}\,\sum_{b\in\bm{b}}\,v_{b}\right]-\frac{(k-1)|\mathcal{M}|% \mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N}roman_inf start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT blackboard_E start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT [ roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ] - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG
    \displaystyle\geq infF𝔽V(F)(k1)||v¯N,subscriptinfimum𝐹𝔽superscript𝑉𝐹𝑘1¯𝑣𝑁\displaystyle\inf_{F\in\mathbb{F}}\,V^{*}(F)-\frac{(k-1)|\mathcal{M}|\mkern 1.% 5mu\overline{\mkern-1.5muv\mkern-1.5mu}\mkern 1.5mu}{N},roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_F ) - divide start_ARG ( italic_k - 1 ) | caligraphic_M | over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_N end_ARG ,

    where the equality follows from the 𝔾𝔾\mathbb{G}blackboard_G-sufficiency of menu \mathcal{M}caligraphic_M, and the two inequalities have been established in the proof of Theorem 2. ∎

A simple sufficient condition for the 𝔾𝔾\mathbb{G}blackboard_G-sufficiency of menu \mathcal{M}caligraphic_M is that

max𝒃(2S)b𝒃vb=max𝒃()b𝒃vbsubscript𝒃superscript2𝑆subscript𝑏𝒃subscript𝑣𝑏subscript𝒃subscript𝑏𝒃subscript𝑣𝑏\displaystyle\max_{\bm{b}\in\mathcal{B}(2^{S})}\,\sum_{b\in\bm{b}}\,v_{b}=\max% _{\bm{b}\in\mathcal{B}(\mathcal{M})}\,\sum_{b\in\bm{b}}\,v_{b}roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = roman_max start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B ( caligraphic_M ) end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT (1)

holds ex-post, i.e. 𝒗Supp(𝔾):=G𝔾Supp(G)for-all𝒗Supp𝔾assignsubscript𝐺𝔾Supp𝐺\forall\bm{v}\in\mathrm{Supp}(\mathbb{G}):=\cup_{G\in\mathbb{G}}\,\mathrm{Supp% }(G)∀ bold_italic_v ∈ roman_Supp ( blackboard_G ) := ∪ start_POSTSUBSCRIPT italic_G ∈ blackboard_G end_POSTSUBSCRIPT roman_Supp ( italic_G ). This condition allows us to convert combinatorial preferences into sufficiency. Theorem 3, as well as Equation 1, states that one only needs to verify the sufficiency of a menu based on individual bidder’s preference, as opposed to the distribution of valuations among all bidders. This is a consequence of the robustness concern. Recall from our analysis in Section 3 that the adversarial nature minimizes the rank-guarantee by making all losing bidders identical. Then, in the worst-case, a 𝔾𝔾\mathbb{G}blackboard_G-sufficient menu performs as good as the complete menu. In the alternative cases where losing bidders are asymmetric, even though the 𝔾𝔾\mathbb{G}blackboard_G-sufficient menu under-performs the complete menu, the extra surplus from asymmetry leads to an even higher rank-guarantee. With Theorem 3 and Equation 1, we derive simple sufficient menus for several canonical preference structures.

Weak substitutability and itemized ascending auction

Definition 4.

We say that bidder preferences exhibit weak substitutability if for any 𝐯Supp(𝔾)𝐯Supp𝔾\bm{v}\in\mathrm{Supp}(\mathbb{G})bold_italic_v ∈ roman_Supp ( blackboard_G ) and bS𝑏𝑆b\subseteq Sitalic_b ⊆ italic_S,

sbv{s}vb.subscript𝑠𝑏subscript𝑣𝑠subscript𝑣𝑏\displaystyle\sum_{s\in b}v_{\{s\}}\geq v_{b}.∑ start_POSTSUBSCRIPT italic_s ∈ italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT { italic_s } end_POSTSUBSCRIPT ≥ italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT .

In words, a representative bidder finds the value of any bundle weakly lower than the sum of her value for each item in the bundle. Weak substitutability is a necessary condition for various substitutability notions studied in the literature.

Proposition 4.

If bidder preferences exhibit weak substitutability, then the menu =S𝑆\mathcal{M}=Scaligraphic_M = italic_S is 𝔾𝔾\mathbb{G}blackboard_G-sufficient and k=M+1𝑘𝑀1k=M+1italic_k = italic_M + 1.

When =S𝑆\mathcal{M}=Scaligraphic_M = italic_S, CASA reduces to a simple itemized ascending auction, where the allocation is determined jointly in the end. Weak substitutability is one of the most widely studied preference assumptions in the literature as it captures a natural diminishing return to scale. Our analysis shows that under such preference structures, CASA exhibits extreme simplicity while achieving both allocation and approximation efficiency. Intriguingly, under weak substitutability, the canonical Vickery auction performs as well as CASA, despite its much worse performance under more general preference structures.

Proposition 5.

If bidder preferences exhibit weak substitutability, then the Vickery auction achieves a revenue guarantee of RSM+1(𝐯)subscriptsuperscript𝑅𝑀1𝑆𝐯{R}^{M+1}_{S}(\bm{v})italic_R start_POSTSUPERSCRIPT italic_M + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( bold_italic_v ).

An even more special case of weak substitutability is the sponsored search auction, where valuations of items are constant (and common) ratios of a one-dimensional private type. As shown in Edelman et al., (2007), the clock auction version of generalized second price (GSP) auction is outcome equivalent to the Vickery auction; hence achieving the same rank-guarantee.

Weak complementarity and the second-price auction

Definition 5.

Bidder preferences exhibit weak complementarity if for any 𝐯Supp(𝔾)𝐯Supp𝔾\bm{v}\in\mathrm{Supp}(\mathbb{G})bold_italic_v ∈ roman_Supp ( blackboard_G ) and 𝐛(2S)𝐛superscript2𝑆\bm{b}\in\mathcal{B}(2^{S})bold_italic_b ∈ caligraphic_B ( 2 start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT ),

b𝒃vbvS.subscript𝑏𝒃subscript𝑣𝑏subscript𝑣𝑆\displaystyle\sum_{b\in\bm{b}}v_{b}\leq v_{S}.∑ start_POSTSUBSCRIPT italic_b ∈ bold_italic_b end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT .

In words, a representative bidder finds the value of the grand bundle weakly higher than the total value of any feasible collection of bundles. Weak complementarity is a necessary condition for various complementarity notions studied in the literature.

Proposition 6.

If bidder preferences exhibit weak substitutability, then the menu ={S}𝑆\mathcal{M}=\{S\}caligraphic_M = { italic_S } is 𝔾𝔾\mathbb{G}blackboard_G-sufficient and k=2𝑘2k=2italic_k = 2.

When ={S}𝑆\mathcal{M}=\{S\}caligraphic_M = { italic_S }, CASA reduces to a simple ascending auction for only the grand bundle. Evidently, in this case, the standard second-price auction is second-guaranteed and outcome-equivalent to CASA.

“Partitional” complementarity

A hybrid case of substitutability and complementarity is the partitional complementarity which we define below, described by a partition 𝒦𝒦\mathcal{K}caligraphic_K of S𝑆Sitalic_S.

Definition 6.

Let 𝒦𝒦\mathcal{K}caligraphic_K be a partition of S𝑆Sitalic_S. Bidder preferences exhibit 𝒦𝒦\mathcal{K}caligraphic_K-partitioned complementarity if for any 𝐯Supp(𝔾)𝐯Supp𝔾\bm{v}\in\mathrm{Supp}(\mathbb{G})bold_italic_v ∈ roman_Supp ( blackboard_G ),

for any b𝒦 and partition κ of b,bκvbvb;formulae-sequencefor any 𝑏𝒦 and partition κ of bsubscriptsuperscript𝑏𝜅subscript𝑣superscript𝑏subscript𝑣𝑏\displaystyle\text{for any }b\in\mathcal{K}\text{ and partition $\kappa$ of $b% $},\ \sum_{b^{\prime}\in\kappa}\,v_{b^{\prime}}\leq v_{b};for any italic_b ∈ caligraphic_K and partition italic_κ of italic_b , ∑ start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ italic_κ end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ;
for any bS,b𝒦vbbvb.formulae-sequencefor any superscript𝑏𝑆subscript𝑏𝒦subscript𝑣𝑏superscript𝑏subscript𝑣superscript𝑏\displaystyle\text{for any }b^{\prime}\subseteq S,\ \sum_{b\in\mathcal{K}}v_{b% \cap b^{\prime}}\geq v_{b^{\prime}}.for any italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ⊆ italic_S , ∑ start_POSTSUBSCRIPT italic_b ∈ caligraphic_K end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_b ∩ italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ≥ italic_v start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT .

In words, 𝒦𝒦\mathcal{K}caligraphic_K-partitioned complementarity structure means there is weak complementarity within each b𝒦𝑏𝒦b\in\mathcal{K}italic_b ∈ caligraphic_K and weak substitutability across each b𝒦𝑏𝒦b\in\mathcal{K}italic_b ∈ caligraphic_K.

Proposition 7.

If bidder preferences exhibit 𝒦𝒦\mathcal{K}caligraphic_K-partitioned complementarity, then the menu =𝒦𝒦\mathcal{M}=\mathcal{K}caligraphic_M = caligraphic_K is 𝔾𝔾\mathbb{G}blackboard_G-sufficient and k=|𝒦|+1𝑘𝒦1k=|\mathcal{K}|+1italic_k = | caligraphic_K | + 1.

In some cases, the auctioneer may understand that bidder preferences exhibits partitional complementarity, but does not know the exact partition. Proposition 7 can be easily extended to the case with multiple possible partitions {𝒦i}i=1Isuperscriptsubscriptsubscript𝒦𝑖𝑖1𝐼\{\mathcal{K}_{i}\}_{i=1}^{I}{ caligraphic_K start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT, where I𝐼Iitalic_I is bounded. In this case =iI𝒦isubscript𝑖𝐼subscript𝒦𝑖\mathcal{M}=\cup_{i\in I}\mathcal{K}_{i}caligraphic_M = ∪ start_POSTSUBSCRIPT italic_i ∈ italic_I end_POSTSUBSCRIPT caligraphic_K start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and kPoly(M)similar-to𝑘𝑃𝑜𝑙𝑦𝑀k\sim Poly(M)italic_k ∼ italic_P italic_o italic_l italic_y ( italic_M ). Such partitional complementarity preference structure arises when there is clear synergy between “nearby” bundles. Think about land auctions, for example. There are finitely many possible partitions that are determined by the major divisions of lands by rivers, highways, or railroads. If two distinct lands are segregated by those divisions, then there is substitutability among them. In such cases, our theory guarantees the performance of CASA with the partitional menu.

Homogeneous goods and quantity-CASA

Definition 7.

The goods are homogeneous if there exists u:[v¯,v¯]:𝑢¯𝑣¯𝑣u:\mathbb{N}\to[\underline{v},\mkern 1.5mu\overline{\mkern-1.5muv\mkern-1.5mu}% \mkern 1.5mu]italic_u : blackboard_N → [ under¯ start_ARG italic_v end_ARG , over¯ start_ARG italic_v end_ARG ] such that for any 𝐯Supp(𝔾)𝐯Supp𝔾\bm{v}\in\mathrm{Supp}(\mathbb{G})bold_italic_v ∈ roman_Supp ( blackboard_G ) and bS𝑏𝑆b\in Sitalic_b ∈ italic_S,

vb=u(|b|).subscript𝑣𝑏𝑢𝑏\displaystyle v_{b}=u(|b|).italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = italic_u ( | italic_b | ) .

With homogeneous goods, a representative bidder’s valuation for any bundle only depends on the size of the bundle. Note that the dependence of u𝑢uitalic_u on |b|𝑏|b|| italic_b | is arbitrary. We do not even require monotonicity. In this case, we redefine the notion of feasible allocations to :()={X|bX|b|M}:conditional-set𝑋subscript𝑏𝑋𝑏𝑀\mathcal{B}:(\mathcal{M})=\{X\subset\mathcal{M}|\sum_{b\in X}\,|b|\leq M\}caligraphic_B : ( caligraphic_M ) = { italic_X ⊂ caligraphic_M | ∑ start_POSTSUBSCRIPT italic_b ∈ italic_X end_POSTSUBSCRIPT | italic_b | ≤ italic_M }, i.e., an allocation is feasible as long as the total number of items being allocated is below M𝑀Mitalic_M.

Proposition 8.

If goods are homogeneous, then the menu =l{1,,M}{bl1,,blMl}subscript𝑙1𝑀superscriptsubscript𝑏𝑙1superscriptsubscript𝑏𝑙𝑀𝑙\mathcal{M}=\cup_{l\in\{1,\ldots,M\}}\,\{b_{l}^{1},\ldots,b_{l}^{\lfloor\frac{% M}{l}\rfloor}\}caligraphic_M = ∪ start_POSTSUBSCRIPT italic_l ∈ { 1 , … , italic_M } end_POSTSUBSCRIPT { italic_b start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , … , italic_b start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⌊ divide start_ARG italic_M end_ARG start_ARG italic_l end_ARG ⌋ end_POSTSUPERSCRIPT } is 𝔾𝔾\mathbb{G}blackboard_G-sufficient and kM2+M2𝑘superscript𝑀2𝑀2k\leq\frac{M^{2}+M}{2}italic_k ≤ divide start_ARG italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_M end_ARG start_ARG 2 end_ARG, where {blj}superscriptsubscript𝑏𝑙𝑗\{b_{l}^{j}\}{ italic_b start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT } are distinct bundles of size l𝑙litalic_l.

In this case, CASA simply auctions Ml𝑀𝑙\lfloor\frac{M}{l}\rfloor⌊ divide start_ARG italic_M end_ARG start_ARG italic_l end_ARG ⌋ copies of each quantity level lM𝑙𝑀l\leq Mitalic_l ≤ italic_M via individual ascending auctions. Like the discussion in partitional complementarity, there may be finitely many types of homogeneous goods. As long as the number of types I𝐼Iitalic_I is bounded, the menu consists of all combinations of Ml𝑀𝑙\lfloor\frac{M}{l}\rfloor⌊ divide start_ARG italic_M end_ARG start_ARG italic_l end_ARG ⌋ copies of each type is sufficient and of size Poly(M)𝑃𝑜𝑙𝑦𝑀Poly(M)italic_P italic_o italic_l italic_y ( italic_M ). Such preference structure is typical in examples like the spectrum auctions. Different frequencies are almost physically homogeneous, except that “middle” frequencies might be of different value from “boundary” frequencies.

5 Concluding remarks

In this paper, we design an auction format of CASA that guarantees an approximately optimal ex-post revenue. To achieve this, we only need to assume minimal rationality on the part of the bidders. In addition, we show that CASA is robust to distributional and strategic uncertainties under certain approximations. In practice, however, these approximation gaps may become non-negligible, rendering the deployment of CASA challenging.

  • Thin markets: The revenue performance of CASA as well as its strategic robustness crucially hinges on the rank k𝑘kitalic_k (menu size) being small relative to the number of bidders. In the online advertising examples we introduce, the complete menu is small enough that a handful of bidders may be sufficient to make CASA an appealing design. However, other interesting auctions may suffer the large menu problem (e.g. the land auctions) or the thin market problem (e.g. the route auctions of rideshare apps) or both (e.g., the spectrum auctions), rendering the guarantee underpowered.

    In the latter cases, the menu sufficiency-approximation efficiency tradeoff becomes eminent. Our theory suggests the importance of preference estimation in those settings. Finding a simple sufficient menu keeps the revenue guarantee appealing and CASA directly applicable. Even in settings with a large number of items and a small number of bidders where our theory has little bite, menu design may still be a cost-effective way to promote competition and improve the revenue performance of existing auctions.

  • Proxy bidding: While CASA simplifies the bidding process by clarifying “whether to quit,” the complexity of determining “which bundles to bid on” and “how much to bid” remains unresolved. A truthful and full proxy-bidding version of CASA is not yet known to us. This makes the deployment of CASA challenging in environments that require fast resolutions of auctions. Nevertheless, we propose that advancements in AI could mitigate this by introducing ”copilot” features that assist bidders in decision-making. By integrating AI as the bidding proxy, bidders would only need to specify values for desired bundles, with the AI advising on bid placement. This could evolve into a hybrid model where bidders either rely fully on platform-provided AI, develop their own bidding algorithms, or use a combination of both strategies.

References

  • Aggarwal and Hartline, (2006) Aggarwal, G. and Hartline, J. D. (2006). Knapsack auctions. In SODA, volume 6, pages 1083–1092.
  • Ausubel et al., (2006) Ausubel, L. M., Cramton, P., and Milgrom, P. (2006). The clock-proxy auction: A practical combinatorial auction design. Handbook of spectrum auction design, pages 120–140.
  • Ausubel and Milgrom, (2002) Ausubel, L. M. and Milgrom, P. R. (2002). Ascending auctions with package bidding. The BE Journal of Theoretical Economics, 1(1):20011001.
  • Bergemann and Morris, (2005) Bergemann, D. and Morris, S. (2005). Robust mechanism design. Econometrica, 73:1771–1813.
  • Börgers, (1991) Börgers, T. (1991). Undominated strategies and coordination in normalform games. Social Choice and Welfare, 8:65–78.
  • Brooks and Du, (2021) Brooks, B. and Du, S. (2021). Optimal auction design with common values: An informationally-robust approach. Econometrica, 89(3):1313–1360.
  • Brooks and Du, (2023) Brooks, B. and Du, S. (2023). On the structure of informationally robust optimal mechanisms. Available at SSRN 3663721.
  • Bulow and Klemperer, (1996) Bulow, J. and Klemperer, P. (1996). Auctions versus negotiations. American Economic Review, 86(1):180–194.
  • Carroll, (2014) Carroll, G. (2014). A complexity result for undominated-strategy implementation. Technical report, Working Paper.
  • Carroll, (2017) Carroll, G. (2017). Robustness and separation in multidimensional screening. Econometrica, 85(2):453–488.
  • Carroll, (2019) Carroll, G. (2019). Design for weakly structured environments. The Future of Economic Design: The Continuing Development of a Field as Envisioned by Its Researchers, pages 27–33.
  • Chen and Li, (2018) Chen, Y.-C. and Li, J. (2018). Revisiting the foundations of dominant-strategy mechanisms. Journal of Economic Theory, 178:294–317.
  • Chung and Ely, (2007) Chung, K.-S. and Ely, J. C. (2007). Foundations of dominant-strategy mechanisms. Review of Economic Studies, 74(2):447–476.
  • Cramton and Schwartz, (2000) Cramton, P. and Schwartz, J. A. (2000). Collusive bidding: Lessons from the fcc spectrum auctions. Journal of regulatory Economics, 17(3):229–252.
  • Cramton et al., (2006) Cramton, P. C., Shoham, Y., Steinberg, R., and Smith, V. L. (2006). Combinatorial auctions, volume 1. MIT press Cambridge.
  • Du, (2018) Du, S. (2018). Robust mechanisms under common valuation. Econometrica, 86(5):1569–1588.
  • Edelman et al., (2007) Edelman, B., Ostrovsky, M., and Schwarz, M. (2007). Internet advertising and the generalized second-price auction: Selling billions of dollars worth of keywords. American economic review, 97(1):242–259.
  • Goldberg and Hartline, (2001) Goldberg, A. V. and Hartline, J. D. (2001). Competitive auctions for multiple digital goods. In European Symposium on Algorithms, pages 416–427. Springer.
  • Grimm et al., (2003) Grimm, V., Riedel, F., and Wolfstetter, E. (2003). Low price equilibrium in multi-unit auctions: the gsm spectrum auction in germany. International journal of industrial organization, 21(10):1557–1569.
  • Hartline, (2013) Hartline, J. D. (2013). Mechanism design and approximation. Book draft. October, 122(1).
  • He and Li, (2022) He, W. and Li, J. (2022). Correlation-robust auction design. Journal of Economic Theory, 200:105403.
  • He et al., (2022) He, W., Li, J., and Zhong, W. (2022). Order statistics of large samples: theory and an application to robust auction design. Technical report, Technical report, Mimeo.
  • Jackson, (1992) Jackson, M. O. (1992). Implementation in undominated strategies: A look at bounded mechanisms. The Review of Economic Studies, 59(4):757–775.
  • (24) Jehiel, P. and Moldovanu, B. (2001a). Efficient design with interdependent valuations. Econometrica, 69(5):1237–1259.
  • (25) Jehiel, P. and Moldovanu, B. (2001b). The european umts/imt-2000 licence auctions. Imt-2000 Licence Auctions (May 2001).
  • Klemperer, (2002) Klemperer, P. (2002). What really matters in auction design. Journal of economic perspectives, 16(1):169–189.
  • Levin and Skrzypacz, (2016) Levin, J. and Skrzypacz, A. (2016). Properties of the combinatorial clock auction. American Economic Review, 106(9):2528–2551.
  • Li and Dworczak, (2021) Li, J. and Dworczak, P. (2021). Are simple mechanisms optimal when agents are unsophisticated? In Proceedings of the 22nd ACM Conference on Economics and Computation, pages 685–686.
  • Li, (2017) Li, S. (2017). Obviously strategy-proof mechanisms. American Economic Review, 107(11):3257–3287.
  • Milgrom, (2000) Milgrom, P. (2000). Putting auction theory to work: The simultaneous ascending auction. Journal of political economy, 108(2):245–272.
  • Myerson, (1981) Myerson, R. (1981). Optimal auction design. Mathematics of Operations Research, 6(1):58–71.
  • Rothkopf et al., (1998) Rothkopf, M. H., Pekeč, A., and Harstad, R. M. (1998). Computationally manageable combinational auctions. Management science, 44(8):1131–1147.
  • Roughgarden, (2015) Roughgarden, T. (2015). Approximately optimal mechanism design: Motivation, examples, and lessons learned. ACM SIGecom Exchanges, 13(2):4–20.
  • Suzdaltsev, (2022) Suzdaltsev, A. (2022). Distributionally robust pricing in independent private value auctions. Journal of Economic Theory, 206:105555.
  • Yamashita, (2015) Yamashita, T. (2015). Implementation in weakly undominated strategies, with applications to auctions and bilateral trade. Review of Economic Studies, 82(3):1223–1246.
  • Yamashita and Zhu, (2022) Yamashita, T. and Zhu, S. (2022). On the foundations of ex post incentive-compatible mechanisms. American Economic Journal: Microeconomics, 14(4):494–514.
  • Zhang, (2022) Zhang, W. (2022). Correlation-robust optimal auctions.

A Omitted Proofs

A.1 Proof of Proposition 3

  • Proof.

Pick an arbitrary bundle b𝑏b\in\mathcal{M}italic_b ∈ caligraphic_M. Let vb=𝟏b=bU[0,1]subscript𝑣superscript𝑏subscript1superscript𝑏𝑏𝑈01v_{b^{\prime}}=\bm{1}_{b^{\prime}=b}\cdot U[0,1]italic_v start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = bold_1 start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_b end_POSTSUBSCRIPT ⋅ italic_U [ 0 , 1 ]; that is, b𝑏bitalic_b is the only valuable bundle and its value is uniformly distributed on [0,1]01[0,1][ 0 , 1 ]. Let G𝐺Gitalic_G denote such a distribution and 𝔾={G}𝔾𝐺\mathbb{G}=\{G\}blackboard_G = { italic_G }. Then, V(F)12subscript𝑉𝐹12V_{\mathcal{M}}(F)\geq\frac{1}{2}italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) ≥ divide start_ARG 1 end_ARG start_ARG 2 end_ARG for any F𝔽𝐹𝔽F\in\mathbb{F}italic_F ∈ blackboard_F. Define Fsuperscript𝐹F^{*}italic_F start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT as follows: uniformly randomly pick k1𝑘1k-1italic_k - 1 bidders and their values for b𝑏bitalic_b are identical and distributed according to U[1k1N,1]𝑈1𝑘1𝑁1U[1-\frac{k-1}{N},1]italic_U [ 1 - divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG , 1 ]. For the remaining bidders, their values for b𝑏bitalic_b are identical and distributed according to U[0,1k1N]𝑈01𝑘1𝑁U[0,1-\frac{k-1}{N}]italic_U [ 0 , 1 - divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG ]. It is straightforward to verify that F𝔽superscript𝐹𝔽F^{*}\in\mathbb{F}italic_F start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∈ blackboard_F and

𝔼F[Rk(𝒗)]=𝔼U[0,1k1N][x]=12k12NinfF𝔽V(F)O(kN).subscript𝔼superscript𝐹delimited-[]subscriptsuperscript𝑅𝑘𝒗subscript𝔼𝑈01𝑘1𝑁delimited-[]𝑥12𝑘12𝑁subscriptinfimum𝐹𝔽subscript𝑉𝐹𝑂𝑘𝑁\displaystyle\mathbb{E}_{F^{*}}[R^{k}_{\mathcal{M}}(\bm{v})]=\mathbb{E}_{U[0,1% -\frac{k-1}{N}]}[x]=\frac{1}{2}-\frac{k-1}{2N}\leq\inf_{F\in\mathbb{F}}\,V_{% \mathcal{M}}(F)-\textstyle O\left(\frac{k}{N}\right).blackboard_E start_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT [ italic_R start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) ] = blackboard_E start_POSTSUBSCRIPT italic_U [ 0 , 1 - divide start_ARG italic_k - 1 end_ARG start_ARG italic_N end_ARG ] end_POSTSUBSCRIPT [ italic_x ] = divide start_ARG 1 end_ARG start_ARG 2 end_ARG - divide start_ARG italic_k - 1 end_ARG start_ARG 2 italic_N end_ARG ≤ roman_inf start_POSTSUBSCRIPT italic_F ∈ blackboard_F end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_F ) - italic_O ( divide start_ARG italic_k end_ARG start_ARG italic_N end_ARG ) .

A.2 Proof of Proposition 5

  • Proof.

We slightly abuse notation and represent an allocation by a vector of sets 𝒃=(b1,b2,,bN)𝒃subscript𝑏1subscript𝑏2subscript𝑏𝑁\bm{b}=(b_{1},b_{2},\ldots,b_{N})bold_italic_b = ( italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_b start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ), where bnbn=subscript𝑏𝑛subscript𝑏superscript𝑛b_{n}\cap b_{n^{\prime}}=\emptysetitalic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∩ italic_b start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = ∅ and bnsubscript𝑏𝑛b_{n}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the bundle allocated to bidder n𝑛nitalic_n. Let Nsubscript𝑁\mathcal{B}_{N}caligraphic_B start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT denote the set of all feasible allocations with N𝑁Nitalic_N bidders. Let 𝒃(𝒗)superscript𝒃𝒗\bm{b}^{*}(\bm{v})bold_italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_v ) denote the efficient allocation.

We establish a lower bound of the revenue-guarantee of the VCG mechanism by constructing, for each n𝑛nitalic_n, an allocation 𝒃nN1superscript𝒃𝑛subscript𝑁1\bm{b}^{n}\in\mathcal{B}_{N-1}bold_italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∈ caligraphic_B start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT of the objects to the bidders other than bidder n𝑛nitalic_n. Clearly, for any such profile 𝒃nsuperscript𝒃𝑛\bm{b}^{n}bold_italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT,

RVCG(𝒗)=subscript𝑅𝑉𝐶𝐺𝒗absent\displaystyle R_{VCG}(\bm{v})=italic_R start_POSTSUBSCRIPT italic_V italic_C italic_G end_POSTSUBSCRIPT ( bold_italic_v ) = n=1N(sup𝒃N1nnvbnnnnvbn(𝒗)n)superscriptsubscript𝑛1𝑁subscriptsupremum𝒃subscript𝑁1subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscript𝑏superscript𝑛subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscriptsuperscript𝑏superscript𝑛𝒗\displaystyle\sum_{n=1}^{N}\left(\sup_{\bm{b}\in\mathcal{B}_{N-1}}\sum_{n^{% \prime}\neq n}v^{n^{\prime}}_{b_{n^{\prime}}}-\sum_{n^{\prime}\neq n}v^{n^{% \prime}}_{b^{*}_{n^{\prime}}(\bm{v})}\right)∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( roman_sup start_POSTSUBSCRIPT bold_italic_b ∈ caligraphic_B start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) end_POSTSUBSCRIPT )
\displaystyle\geq n=1N(nnvbnnnnnvbn(𝒗)n).superscriptsubscript𝑛1𝑁subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscriptsuperscript𝑏𝑛superscript𝑛subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscriptsuperscript𝑏superscript𝑛𝒗\displaystyle\sum_{n=1}^{N}\left(\sum_{n^{\prime}\neq n}v^{n^{\prime}}_{b^{n}_% {n^{\prime}}}-\sum_{n^{\prime}\neq n}v^{n^{\prime}}_{b^{*}_{n^{\prime}}(\bm{v}% )}\right).∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) end_POSTSUBSCRIPT ) . (2)

For each n𝑛nitalic_n, we construct an allocation 𝒃nN1superscript𝒃𝑛subscript𝑁1\bm{b}^{n}\in\mathcal{B}_{N-1}bold_italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∈ caligraphic_B start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT via the following algorithm:

Algorithm. Bundle bnn=subscriptsuperscript𝑏𝑛superscript𝑛b^{n}_{n^{\prime}}=\emptysetitalic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = ∅ for all nsuperscript𝑛n^{\prime}italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. Set O=bn(𝒗)𝑂subscriptsuperscript𝑏𝑛𝒗O=b^{*}_{n}(\bm{v})italic_O = italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_v ).

  • (1).

    For each nnsuperscript𝑛𝑛n^{\prime}\neq nitalic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n:

    If bn(𝒗)subscriptsuperscript𝑏superscript𝑛𝒗b^{*}_{n^{\prime}}(\bm{v})\neq\emptysetitalic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) ≠ ∅, set bnn=bn(𝒗)subscriptsuperscript𝑏𝑛superscript𝑛subscriptsuperscript𝑏superscript𝑛𝒗b^{n}_{n^{\prime}}=b^{*}_{n^{\prime}}(\bm{v})italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ).

    Let N¯={n:bnn=,nn}¯𝑁conditional-setsuperscript𝑛formulae-sequencesubscriptsuperscript𝑏𝑛superscript𝑛superscript𝑛𝑛\bar{N}=\{n^{\prime}:b^{n}_{n^{\prime}}=\emptyset,n^{\prime}\neq n\}over¯ start_ARG italic_N end_ARG = { italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT : italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = ∅ , italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n }.

  • (2).

    If O𝑂O\neq\emptysetitalic_O ≠ ∅, then pick oO𝑜𝑂o\in Oitalic_o ∈ italic_O.

    Set bnn={o}subscriptsuperscript𝑏𝑛superscript𝑛𝑜b^{n}_{n^{\prime}}=\{o\}italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = { italic_o } for some nargmaxn′′N¯vn′′({o})superscript𝑛subscriptsuperscript𝑛′′¯𝑁subscript𝑣superscript𝑛′′𝑜n^{\prime}\in\arg\max_{n^{\prime\prime}\in\bar{N}}v_{n^{\prime\prime}}(\{o\})italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ roman_arg roman_max start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ∈ over¯ start_ARG italic_N end_ARG end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( { italic_o } ).

    Update OO{o}𝑂𝑂𝑜O\leftarrow O\setminus\{o\}italic_O ← italic_O ∖ { italic_o } and N¯N¯{n}¯𝑁¯𝑁superscript𝑛\bar{N}\leftarrow\bar{N}\setminus\{n^{\prime}\}over¯ start_ARG italic_N end_ARG ← over¯ start_ARG italic_N end_ARG ∖ { italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT }.

  • (3).

    Repeat (2) until O=𝑂O=\emptysetitalic_O = ∅.

  • (4).

    Return allocation 𝒃n=(b1n,b2n,,bn1n,bn+1n,,bNn)superscript𝒃𝑛subscriptsuperscript𝑏𝑛1subscriptsuperscript𝑏𝑛2subscriptsuperscript𝑏𝑛𝑛1subscriptsuperscript𝑏𝑛𝑛1subscriptsuperscript𝑏𝑛𝑁\bm{b}^{n}=(b^{n}_{1},b^{n}_{2},\ldots,b^{n}_{n-1},b^{n}_{n+1},\ldots,b^{n}_{N})bold_italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = ( italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT , italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT , … , italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ).

In words, if an object is allocated to a bidder other than bidder n𝑛nitalic_n under 𝒃(𝒗)superscript𝒃𝒗\bm{b}^{*}(\bm{v})bold_italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_v ), then the object is still allocated to that bidder. We then iteratively pick an object o𝑜oitalic_o that is allocated to bidder n𝑛nitalic_n under 𝒃(𝒗)superscript𝒃𝒗\bm{b}^{*}(\bm{v})bold_italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_v ), and allocate the object to the bidder nsuperscript𝑛n^{\prime}italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT whose value for the object v{o}nsubscriptsuperscript𝑣superscript𝑛𝑜v^{n^{\prime}}_{\{o\}}italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT is the highest among all the bidders who are not allocated any object yet. For each obn𝑜subscriptsuperscript𝑏𝑛o\in b^{*}_{n}italic_o ∈ italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, define nosubscript𝑛𝑜n_{o}italic_n start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT to be the index nsuperscript𝑛n^{\prime}italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT such that bnn={o}subscriptsuperscript𝑏𝑛superscript𝑛𝑜b^{n}_{n^{\prime}}=\{o\}italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = { italic_o }.

It follows from Equation 2 that

RVCG(𝒗)subscript𝑅𝑉𝐶𝐺𝒗absent\displaystyle R_{VCG}(\bm{v})\geqitalic_R start_POSTSUBSCRIPT italic_V italic_C italic_G end_POSTSUBSCRIPT ( bold_italic_v ) ≥ n=1N(nnvbnnnnnvbn(𝒗)n)superscriptsubscript𝑛1𝑁subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscriptsuperscript𝑏𝑛superscript𝑛subscriptsuperscript𝑛𝑛subscriptsuperscript𝑣superscript𝑛subscriptsuperscript𝑏superscript𝑛𝒗\displaystyle\sum_{n=1}^{N}\left(\sum_{n^{\prime}\neq n}v^{n^{\prime}}_{b^{n}_% {n^{\prime}}}-\sum_{n^{\prime}\neq n}v^{n^{\prime}}_{b^{*}_{n^{\prime}}(\bm{v}% )}\right)∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ≠ italic_n end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) end_POSTSUBSCRIPT )
=\displaystyle== n=1Nobnv{o}nosuperscriptsubscript𝑛1𝑁subscript𝑜subscriptsuperscript𝑏𝑛subscriptsuperscript𝑣subscript𝑛𝑜𝑜\displaystyle\sum_{n=1}^{N}\sum_{o\in b^{*}_{n}}v^{n_{o}}_{\{o\}}∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_o ∈ italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT
\displaystyle\geq o=1Mv{o}(M+1)superscriptsubscript𝑜1𝑀subscriptsuperscript𝑣𝑀1𝑜\displaystyle\sum_{o=1}^{M}v^{(M+1)}_{\{o\}}∑ start_POSTSUBSCRIPT italic_o = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_v start_POSTSUPERSCRIPT ( italic_M + 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT (3)
=\displaystyle== RM+1(𝒗).subscriptsuperscript𝑅𝑀1𝒗\displaystyle{R}^{M+1}_{\mathcal{M}}(\bm{v}).italic_R start_POSTSUPERSCRIPT italic_M + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( bold_italic_v ) .

The first equality holds since (a) when bn(𝒗)subscriptsuperscript𝑏superscript𝑛𝒗b^{*}_{n^{\prime}}(\bm{v})\neq\emptysetitalic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) ≠ ∅, bnn=bn(𝒗)subscriptsuperscript𝑏𝑛superscript𝑛subscriptsuperscript𝑏superscript𝑛𝒗b^{n}_{n^{\prime}}=b^{*}_{n^{\prime}}(\bm{v})italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ), and (b) when bn(𝒗)=subscriptsuperscript𝑏superscript𝑛𝒗b^{*}_{n^{\prime}}(\bm{v})=\emptysetitalic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_v ) = ∅, bnnsubscriptsuperscript𝑏𝑛superscript𝑛b^{n}_{n^{\prime}}italic_b start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is either {o}𝑜\{o\}{ italic_o } for some oO𝑜𝑂o\in Oitalic_o ∈ italic_O, or \emptyset otherwise. The second inequality follows from the construction of 𝒃isuperscript𝒃𝑖\bm{b}^{i}bold_italic_b start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT: when an object obi𝑜subscriptsuperscript𝑏𝑖o\in b^{*}_{i}italic_o ∈ italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is being allocated, it is allocated to the bidder nsuperscript𝑛n^{\prime}italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT whose value for the object v{o}nsubscriptsuperscript𝑣superscript𝑛𝑜v^{n^{\prime}}_{\{o\}}italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT is the highest among all the bidders who are not allocated any object yet. Since each iteration assigns at least one good to one bidder and there are at most M𝑀Mitalic_M goods, we have v{o}nosubscriptsuperscript𝑣subscriptsuperscript𝑛𝑜𝑜v^{n^{\prime}_{o}}_{\{o\}}italic_v start_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT must be at least the (M+1)thsuperscript𝑀1𝑡(M+1)^{th}( italic_M + 1 ) start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT highest value among all v{o}nsubscriptsuperscript𝑣𝑛𝑜v^{n}_{\{o\}}italic_v start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT { italic_o } end_POSTSUBSCRIPT. ∎