Limitations of VCG-Based Mechanisms Shahar Dobzinski



Noam Nisan



September 12, 2010

Abstract We consider computationally-efficient truthful mechanisms that use the VCG payment scheme, and study how well they can approximate the social welfare in auction settings. We present a novel technique for setting lower bounds on the approximation ratio of this type of mechanisms. Our technique is based on setting lower bounds on the communication complexity by analyzing combinatorial properties of the algorithms. Specifically, for combinatorial auctions among sub1

modular (and thus also subadditive) bidders we prove an Ω(m 6 ) lower bound, which is close to the 1

known upper bound of O(m 2 ), and qualitatively higher than the constant factor approximation possible from a purely computational point of view.

1

Introduction

1.1

Background

Algorithmic Mechanism design attempts to design protocols for distributed environments, such as the Internet, where the different participants each have their own selfish goals and are assumed to rationally attempt optimizing their own goals rather than just follow any prescribed protocol. The key target in this area is the design of truthful mechanisms – also called incentive-compatible or strategy-proof mechanisms – whose payment schemes motivate the participants to correctly report their private information. For a general introduction to the economic field of mechanism design see [24] and for an introduction to algorithmic mechanism design and further motivation see [27]. ∗ †

Department of Computer Science, Cornell University, [email protected] The School of Computer Science and Engineering, The Hebrew University of Jerusalem, [email protected].

1

Typical problems in this setting involve allocation of various resources and a paradigmatic abstraction is that of combinatorial auctions. In this problem m heterogenous “items” need to be allocated between n “bidders”. Each bidder i holds a valuation function vi that specifies for each subset of the items S ⊆ {1, . . . , m} the bidder’s value vi (S) from winning the “bundle” S. The valuation function vi is the private information of each bidder i. The challenge is to find a partition (S1 , . . . , Sn ) of the items that maximizes the social welfare Σi vi (Si ). A combination of issues makes this problem particularly difficult. First, there is a representational difficulty: a naive representation of each valuation function requires describing 2m values, one for each bundle, while we are interested in algorithms with running time polynomial in the natural parameters of the problem, n and m. Second, an exact solution is hard to compute and even to approximate even in cases where a succinct description of the valuation exists [31]. The final difficulty is to design algorithms to handle the strategic behavior of the bidders. The first problem, the representational difficulty, is handled in this paper by a communication complexity approach (see [19]): we assume that each valuation is represented by a black box. The mechanism repeatedly queries the bidders (the black boxes). The answer to a query to bidder i depends only on vi and the previous answers to the queries by the other bidders. We count only the number of bits transferred (and not, e.g., the computational difficulty of answering the queries). The lower bounds we are interested in apply for every type of query and are in fact communication complexity lower bounds. However, the upper bounds we are interested in (and are not discussed in this paper) assume some specific natural type of query (usually a “demand query” (see [6]). The rest of the paper deals with handling the other two difficulties. The key positive technique for achieving incentive compatibility is that of VCG mechanisms [33, 7, 16]. A powerful observation is that if the algorithmic outcome a always maximizes the social welfare, Σi vi (a) (or weighted versions of it), then the VCG payment rule results in a truthful mechanism. However, in most interesting computational scenarios, including combinatorial auctions, achieving exact optima is computationally intractable, and one must settle for heuristics or approximations. A key clash between the strategic and algorithmic considerations is that once only approximations or heuristics are chosen, the VCG payment rule no longer leads to incentive compatibility. We discuss this issue in more details later.

2

The heart of “algorithmic mechanism design” is in trying to overcome this clash: design mechanisms that are both computationally tractable (and thus only approximate the optimum) and strategically truthful. The key question is always to what extent do the strategic requirements degrade the quality of the solution beyond the degradation implied by the purely computational constraints. So far, the success achieved is only partial. For some domains, “single-parameter” domains, where, roughly speaking, the private information of each player consists of only one number, one can summarize the state of the art as having clear characterizations and many good upper bounds (approximation mechanisms) for various classes (e.g., [23, 25, 1, 8]). On the other hand, for more general “multi-parameter” problems where the private information consists of more than one number (like combinatorial auctions, where the private information of the bidders is a hard-to-represent valuation funcion), one can safely say that little progress has been made in obtaining upper or lower bounds. This paper wishes to address this issue, attempting to prove lower bounds – showing that the strategic constraints do impose an additional burden beyond the computational ones. See [27] for a discussion.

1.2

Algorithmic Uses of VCG

We are unable to prove general lower bounds, so we limit ourselves to a natural – and interesting in itself – class of mechanisms: VCG-based ones. Before defining this class we describe the VCG payment scheme. 1.2.1

The VCG Payment Scheme

Arguably the main positive result of mechanism design is the VCG payment scheme. Let us describe this payment scheme when applied to combinatorial auctions. First, find the optimal solution (O1 , ..., On ), and allocate accordingly. Then, pay each bidder the sum of the utilities of the rest of the bidders. That is, bidder i receives1 a payment of Σk6=i vk (Ok ). Consider the total utility of bidder i: vi (Oi ) + Σk6=i vk (Ok ) (the value he gains from the bundle he got plus the payment he receives). Hence, the total utility of each bidder is equal to the value of the allocation obtained by the algorithm. Observe that the allocation that maximizes the utility of the bidders is the optimal one. Bidding 1 To simplify the presentation, we describe a payment scheme in which the mechanism pays the bidders, while it is more natural to assume that the bidders pay the auctioneer for receiving items. The standard way to do that is to subtract a suitable constant from the payments. See [27] for more details. The subtraction has no algorithmic or game-theoretic effect whatsoever, so this paper ignores it, in order to keep the presentation simpler.

3

untruthfully can only result in changing the allocation to a suboptimal one, hence decreasing the utility of bidder i. Thus bidding truthfully is the best action for each bidder. See [27] for a more formal discussion, including a discussion of the straightforward extension to weighted VCG. 1.2.2

Maximal-in-Range Mechanisms

The obvious drawback of using the VCG mechanism is that it requires us to find the optimal solution. In many setting finding the optimal solution is not computationally feasible, and this is true in particular in the settings considered in this paper. In general, obtaining an approximate solution using an approximation algorithm and using the VCG payment scheme (paying each bidder the sum of the utilities of the rest of the bidders) does not result in a truthful mechanism – this will be discussed shortly. However, the following class of algorithms does result in a truthful mechanism when the VCG payment scheme is used: Definition: An algorithm (that produces an output a ∈ A for each input v1 , . . . , vn , where A is the set of possible outcomes) is called “maximal-in-range” (henceforth MIR) if it completely optimizes the weighted social welfare over some subrange R ⊆ A. I.e., for some R ⊆ A, some w1 , . . . , wn > 0, and a constant ca for every outcome a, we have that for all v1 , . . . , vn , a ∈ arg maxa∈R Σi wi vi (a) + ca . I.e., MIR algorithms use the following natural and simple strategy to find an approximately optimal solution: optimally search within a pre-specified sub-range of feasible solutions – a subrange over which optimal search is algorithmically feasible. For example, for combinatorial auctions with subadditive valuations the algorithm of [11] chooses the best allocation from the set that consists of all allocations where each bidder is allocated at most one item and all allocations where one bidder receives all items. In a more algorithmic context, an example for such a strategy is approximating the optimal Steiner tree by taking the best spanning tree [32]. Another example is Arora’s approximation for the traveling salesman problem [2]2 . The main result of [28] states that this is essentially it – if a truthful mechanism uses the VCG payment scheme (is VCG-Based ) it is essentially maximal in range: 2

For most algorithmic uses, especially for approximation purposes, the natural choice is to set all the wi ’s to 1 and all the ca ’s to 0.

4

Theorem [28]: The allocation algorithm of any truthful VCG-based mechanism for combinatorial auctions is equivalent to a maximal-in-range algorithm3 . “Equivalent” here means that the social utilities are identical for all inputs, i.e. if S and T are the outputs of the two allocation algorithms for input v1 , . . . , vn then Σi wi vi (Si ) + cS = Σi wi vi (Ti ) + cT . In particular, the outputs must coincide generically – except perhaps in case of ties. In [28] this theorem was presented for two specific types of mechanism design problems, but the result is more general. See the conference version of this paper [9] for an extension. In this paper, we prove bounds on the power of MIR mechanisms4 . Let us explicitly justify why are lower bounds for this restricted class interesting: 1. Their power: Many truthful mechanisms put forward do fall into this category. This in√ cludes a slightly non-trivial O(m/ log m)-approximation for general combinatorial auctions √ [17], an O( m) approximation for combinatorial auctions with subadditive bidders [11], a 2-approximation for multi-unit auctions, which is improved to a PTAS for certain bidding languages [10], welfare maximization in congestion games [5], and several auctions for “geometric figures” on the plane [3]. 2. Lack of alternatives: Not only are VCG mechanisms the only general method known for constructing truthful mechanisms in multi-parameter settings, there is just a single additional example of polynomial-time non-VCG truthful mechanism in any multi-parameter domain that provides a good approximation ratio [4]. Moreover, Roberts’ classic theorem [30] states that in completely unrestricted domains the only truthful mechanisms are MIR mechanisms. In [20], it was suggested that Roberts’ theorem could be extended to many other domains including combinatorial auctions and multi-unit auctions. In [20], they were only able to prove this for special cases or under additional assumptions, and left the general question open. If this line of attack reaches its conclusion, then our lower bounds would apply in general. We should note that if randomization is allowed then the second point no longer holds with such force, as several randomized truthful mechanisms are known [12, 21, 14]. It is not known, however, 3 4

Nisan and Ronen [28] do not consider weighted version of VCG but their proof carries on also to this case. Our results carry on also to VCG-based mechanisms, as we will point out.

5

whether any of these can be de-randomized (even if P = BP P ). The resolution of this question may rely on the successful characterization of deterministic truthful mechanisms5 .

1.3

Our Main Result

The main interest of this paper is in the following combinatorial problem: what is the approximation factor that can be achieved by MIR mechanisms for combinatorial auctions? Our lower bound applies to the subclass of submodular valuations (for each S ⊆ T ⊆ M and j ∈ / T , we have that vi (T ∪ {j}) − vi (T ) ≤ vi (S ∪ {j}) − vi (S)) and thus also to its superset class of subadditive valuations (vi (S ∪T ) ≤ vi (S)+vi (T ) for all S, T ) – two classes of valuations which have been extensively studied [22, 11, 14, 13, 15, 18]. In particular, constant upper bounds are known for this class [15, 22]. The √ best deterministic mechanism for this case is the MIR O( m)-mechanism of [11]. We note that the technique we present is quite general, and we believe it will turn even more useful by extending our results to other domains as well. Theorem: Every MIR mechanism for approximating the welfare in combinatorial auctions with submodular bidders that uses a sub-exponential number of queries to the bidders achieves an approximation factor of Ω(min(n, m1/6 )). The proof proceeds by combinatorially analyzing maximum in range allocation algorithms. The analysis shows that if the range is “large” then optimizing over it requires exponential communication, while if it is “small” then it can not achieve a good approximation ratio. It turns out that “large” and “small” in this sense cannot just be interpreted in terms of the size of the range. Instead we define two ”complexity measures” of a set of partitions (which is what the range is). One of them, termed the intersection number, is shown to bound from below the communication complexity of optimization over the range. The other, termed the cover number, is shown to bound from above the approximation ratio achieved by allocations in the range. Our main combinatorial lemma, which may be of independent interest, shows that these two complexity measures are related to each other. We stress that although communication complexity methods were already used in [26, 29] in the context of combinatorial auctions, our methods are completely different and this difference is 5

Nevertheless, we will see that the lower bound does hold for some randomized algorithms, namely when the algorithm is a probability distribution over MIR algorithms.

6

inherent: [26, 29] did not consider incentives issues at all. We also show how the lower bound can be extended to the case where the approximation algorithm is guaranteed to provide a good approximation ratio only on “most” inputs. We use this result to show that a randomized MIR algorithm – a probability distribution over polynomial time deterministic MIR algorithms – cannot obtain a good approximation ratio; If such a randomized algorithm would have existed, it is not hard to see that there is a deterministic algorithm (in its support) that provide a good approximation ratio on “most” inputs.

1.4

Open Problems

Open problems are left at various levels of generality. At the most specific level, the problem is to close the gap between the m1/6 lower bound and the m1/2 upper bound. (We can only improve the lower bound to m1/5−² .) At the next level, the question is how well can MIR mechanisms approximate the social welfare in combinatorial auctions with general valuations? We only have a slightly better lower bound than what we have for the submodular case, but the upper bound of [17] is nearly linear. Of course, the real questions are always how well can arbitrary computationally efficient truthful mechanism do – not just MIR ones – and obtaining any such lower bound would be of great interest. This would likely require some advances in the “LMN-program” [20] of characterizing truthful mechanisms in multi-parameter domains.

2

Preliminaries

In a combinatorial auction we have a set M , |M | = m, of heterogeneous items and a set of N bidders, |N | = n. Each bidder i has a valuation function vi : 2M → R. We assume that each valuation vi is normalized (i.e., vi (∅) = 0) and monotone (for each S ⊆ T , vi (S) ≤ vi (T )). An allocation is an n-tuple S = (S1 , . . . , Sn ), where for each i, Si ⊆ M , and for each i 6= i0 , Si ∩ Si0 = ∅. The welfare of an allocation S is defined to be Σi vi (Si ). Fixing w1 , . . . , wn and cS (one for each possible allocation S) the weighted welfare of S is Σi wi vi (Si ) + cS . A valuation v is said to be submodular if it exhibits decreasing marginal utilities. I.e., for each S ⊆ T ⊆ M and j ∈ / T , we have that vi (T ∪ {j}) − vi (T ) ≤ v i (S ∪ {j}) − vi (S). We will also use a very simple subset of submodular valuations called additive valuations. A valuation v is said to be 7

additive if for each S ⊆ M , we have that v(S) = Σj∈S v({j}). We will call an additive valuation v uniformly additive if for each item j we have that v({j}) = {0, 1}. Let A be an algorithm for combinatorial auctions. Let A(I) denote the welfare of the allocation that A produces given the instance I. The approximation ratio of A is the minimum fraction of welfare it achieves over all instances: inf I∈I

A(I) OP T (I) ,

where I is the set of all instances, and OP T (I)

is the welfare of the best allocation. Similarly, the approximation ratio of a randomized algorithm is A(I) inf I∈I E[ OP T (I) ], where the expectation is taken over the internal random coin flips of the algorithm.

We will also consider algorithms that guarantee a good approximation ratio on a large fraction of the instances (not necessarily on all of them). Definition 2.1 A deterministic approximation algorithm is β-good on some set of instances U if the guaranteed approximation is provided only to (at least) β fraction of the instances in U . Trivially, an α-approximation algorithm (in the usual sense of the term) is 1-good on the set of all instances. In this paper we restrict ourselves to a special kind of algorithms: VCG based. These algorithms are essentially the set of algorithms that are implementable using the VCG payment scheme, and are a slight generalization of maximal in range algorithms. Definition 2.2 (Essentially [28]) A VCG-based algorithm always chooses an allocation T such that wi vi (Ti ) + cT = maxS=(S1 ,...,Sn )∈R wi vi (Si ) + cS , for some range R. The wi ’s and the cS ’s are predetermined non-negative constants and in particular does not depend on the valuations of bidders. It is not hard to see that using the VCG payment scheme (or a weighted version of it) makes VCG-based algorithms truthful. We will also consider randomized VCG-based algorithms: in this paper we interpret this term as a probability distribution over deterministic VCG-based algorithms. Finally, all logarithms in the paper are in base 2.

8

3

Combinatorial Auctions with Submodular Bidders

3.1

The Main Result

In this section we analyze the power of MIR algorithms in the context of combinatorial auctions √ with submodular bidders. For this setting, an O( m)-approximation MIR algorithm is known [11]. We will show that this is (almost) the best approximation one can get using MIR algorithms. The theorem is stated only for MIR algorithms but we will point out how it can be extended to algorithms that are equivalent to MIR algorithms, and thus to all VCG-based mechanisms. Theorem 3.1 Every MIR algorithm for approximating the welfare in combinatorial auctions with 1

submodular bidders that uses O(em 15 ) bits of communication achieves an approximation factor of Ω(min(n, m1/6 )). The result also holds for randomized MIR mechanisms. We first prove the result for MIR algorithms where the allocation constants, the c0S s, are identical (without loss of generality, all are zero). Then, in Subsection 3.5, we show how this result is extended to general MIR algorithms, i.e., ones where the cS ’s are not necessarily zero. We define two complexity measures for the range R of an MIR algorithm A: the cover number, and the intersection number. The cover number roughly corresponds to the size of the range R. We will show, using the probabilistic method, that if the cover number is “small” then there exists an instance such that A fails to provide a good approximation. Therefore, the range R must be “large”. In this case we will show that the intersection number of A must be exponential. We will see that the intersection number serves as a lower bound on the communication complexity of A, and so we get that any MIR approximation algorithm that provides a good approximation ratio must have exponential communication complexity. We actually prove the result for deterministic

1 n -good

MIR

algorithms. In the next section we use this to see how the bound can be extended to randomized MIR mechanisms (1-good) as well. The proof of the theorem starts with Subsection 3.2, where the cover number is formally defined and its relation to the approximation ratio is shown. In Subsection 3.3 we define and discuss the second measure: the intersection number. The proof concludes in Subsection 3.4 by showing the relationship between the measures.

9

3.2

Complexity Measure I: The Cover Number

Intuitively we wish to rely on the size of the range. Yet, naive counting will fail to provide good results, since a single allocation in the range may contain many “degenerate allocations”. For example, if the range contains an allocation that assigns all items to some bidder i, it also contains all allocations such that i is assigned any subset of the items, and the rest of the bidders get nothing. These exponentially many allocations are degenerate in the sense that we can assume that they are not in the range of the algorithm without changing the guaranteed approximation ratio of the A. We therefore use an alternative measure for describing the “size” of the range. Definition 3.2 A set C of allocations is said to be a cover set of another set of allocations R if for each S ∈ R there exists some C ∈ C such that for all i, Si ⊆ Ci . The cover number of a set of allocations R is defined to be the size of the minimum cardinality cover set of R. The cover number is denoted by cover(R). Definition 3.3 An instance I is called perfectly uniform-additive if the valuation of each bidder i is uniformly additive, and each item j is demanded by exactly one bidder, i.e., for each item j, vi ({j}) = 1 for exactly one bidder i, and for every k 6= i we have that vk ({j}) = 0. In the next lemma we prove that if cover(R) is small, then there exists some instance in which A provides only Ω(n)-approximation. Lemma 3.4 Let A be a

1.01 n -approximation

algorithm that is x-good on the set of all perfectly m

uniform-additive instances. Then, cover(R) ≥ x · e 300n . Proof:

We select an instance from the set of all perfectly uniform-additive instances uniformly

at random. Since the valuations are additive, we only need to specify the value of vi ({j}) for each bidder i and item j. This is done in the following way: for each item j ∈ M choose exactly one bidder, where each bidder is selected with probability of exactly n1 . Let i be the selected bidder. We set the value of vi ({j}) to be 1. For each i0 6= i we set the value of vi0 ({j}) to be 0. Notice that the value of the optimal solution in each such instance is exactly m. Let U be the set of instances that are perfectly uniform-additive for which A provides an approximation ratio of

1.01 n .

The following version of the Chernoff bounds will be useful: 10

Lemma 3.5 (Chernoff bound) Let X1 , . . . , Xm be independent random variables that take values in {0, 1}, such that for all i, Pr[Xi = 1] = p for some p. Then for every 0 ≤ δ ≤ 2e − 1 it holds that: Pr[Σi Xi > (1 + δ)pm] ≤ e−

pmδ 2 3

Let C be the minimum cardinality cover set of R with |C| = cover(R). Fix some C ∈ C. The probability that vi ({j}) = 1, and that j ∈ Ci is exactly Chernoff bound, Pr[Σi vi (Ci ) >

1+δ n m]

≤e

2 − δ3nm

1 n,

for any bidder i and item j. By the

. Suppose for contradiction that cover(R) < xe

δ2 m 3n

.

Choose a random perfectly uniform-additive instance. By the union bound, with probability greater than 1 − x no allocation in C provides a welfare greater than

1+δ n m.

Obviously, no allocation in R

can provide a welfare greater than ( 1+δ n )m for these instances. A contradiction. The lemma follows by choosing δ = .01.

3.3

Complexity Measure II: The Intersection Number

The second complexity measure we consider is the intersection number. We will show that the intersection number of the range of an MIR algorithm is a lower bound on its communication complexity. Before defining the intersection number, we need a structural definition of a set of allocations. Definition 3.6 We say that a set of allocations R is regular if there exist constants s1 , . . . , sn such that for all S ∈ R and for all 1 ≤ i ≤ n it holds that |Si | = si . We are now ready to define the complexity measure itself. Definition 3.7 A set of allocations D is called an (i, j)-intersection set if for each D, D0 ∈ D, D 6= D0 , it holds that Di ∩ Dj0 6= ∅. Define the intersection number of a set of allocations R, denoted by intersect(R), to be the maximum cardinality regular (i, j)-intersecting set which is a subset of R, where the maximum is taken over all pairs of bidders i and j. Notice that in the definition of the intersection number we require that the intersection set will be regular.

11

The next lemma shows that we can use the intersection number as a lower bound on the communication complexity of the algorithm. Lemma 3.8 Let A be an M IR algorithm for combinatorial auctions with submodular bidders with range R. Let intersect(R) = d. Then, the communication complexity of A is at least d. Proof:

We reduce from the disjointness problem (see [19])6 . In this problem Alice holds a d-bit

string a1 , . . . , ad , and Bob holds a d-bit string b1 , . . . , bd . The goal is to decide whether there exists some index k such that ak = bk = 1. Solving the disjointness problem requires d communication bits. Let D = {D1 , . . . , Dd } be an (i, j)-intersection set of R. D is regular, so for each bidder t there exists a constant st such that |Dt | = st , for all D ∈ D. Construct a combinatorial auction with m items in the following way: Alice will play the role of bidder i, and Bob will play the role of the rest of the bidders, in particular bidder j. We now define the valuations of the bidders. Let the valuation of bidder i played by Alice be (w1 , . . . , wn are the constants in the definition of an M IR algorithm):    |S|/wi , |S| ≤ si − 1;    vi (S) = si /wi , ∃k s.t. Dik ⊆ S and ak = 1;      (si − 2−(|S|−si +1) )/wi , otherwise. The valuation vj is defined in an analogous way. The valuations of the rest of the bidders have zero value for every bundle. The reader is encouraged to verify that all valuations are indeed submodular. Observe that if there exists some index k such that the k’th input bit of both players is 1, then the optimal welfare is si + sj . Otherwise, the optimal welfare is strictly less than si + sj . To see this notice that if bidder i gains a value of si from the bundle S1 he was assigned by A, then there must be an index k such that Dik ⊆ S1 and ak = 1. In order of bidder j to gain a value of sj /wj he must 0

have an index k 0 such that Djk ⊆ S2 . However, D is an (i, j)-intersection set and so it must hold 0

that Dik ∩ Djk 6= ∅, and thus S1 ∩ S2 6= ∅. Clearly, the optimal weighted welfare in this case is less than si + sj . 6

We note that the disjointness lower bound holds for randomized algorithms as well. However, we do not use this fact, even though our main theorem does apply to randomized algorithms. In fact, we could have used the simple fooling set technique (described in [19]) to obtain the same results (we reduce from the disjointness problem to simplify the presentation).

12

By construction, if the optimal weighted welfare is si + sj then it can be achieved by an allocation in R. A is a maximal in range algorithm, and so the weighted welfare of the allocation returned by A in this case must be si + sj . Thus, we will be able to decide if there is a some index k such that ak = bk = 1. This proves that the communication complexity of A is at least as that of the disjointness problem: d. Notice that our lower bound applies even for computing the value of the optimal allocation in R, and thus applies not only to maximal in range algorithms but also to algorithms that are equivalent to maximal in range algorithms (i.e., VCG-based algorithms).

3.4

The Relationship between the Measures

It is easy to see that cover(R) ≥ intersect(R). This subsection shows that the gap between the two is not too large. Specifically, if intersect(R) is small, then cover(R) is small too. Lemma 3.9 Let R be a set of allocations with intersect(R) ≤ d. Then 3

2

cover(R) < (8d)m 5 n · m4m 5 n 1

2

1

m

As a corollary7 , let n = m 6 . If cover(R) > e 300n then intersect(R) ≥ em 15 . Thus, proving the lemma, together with Lemmas 3.4 (with x =

1 n)

and 3.8, derives Theorem 3.1, at least for the case

where all cS ’s are zero. The case where all the cS ’s are not zero is handled in the next subsection. Proof:

(of Lemma 3.9) The lemma will follow from the following claim.

e be a regular set of allocations. If intersect(R) e ≤d Claim 3.10 Fix some w, 1 ≤ w ≤ m. Let R e where then there is a subset E of R

|E| e |R|

2

2

≥ (8d)−mn/w 4−n , and cover(E) ≤ wn mwn nn .

The lemma is proved by partitioning R to up to mn classes of regular allocations, R1 , . . . , Rmn , one for each possible choice of constants s1 , . . . , sn from Definition 3.6. Each si is between 1 and m, so there are at most mn classes. For each class Rs we will set an upper bound on cover(Rs ) separately: 1

7 ²

em

m

The result is actually stronger: fix a constant ² > 0, and let n < m 5 −² . If cover(R) > e 300n then intersect(R) ≥ . The statement of the theorem improves accordingly.

13

Let E1s be the set obtained from the claim. Look at Rs \ E1s , and obtain from the claim another set E2s ⊆ Rs \ E1s with small cover, and so on. After (8d)

mn w

2

· 4n · log |Rs | steps Rs is completely

covered. Now cover(Rs ) can be bounded from above by Σk cover(Eks ). By bounding from above the 2

size of each class |Rs | by |R| ≤ nm , we have that (by choosing w = m 5 ): n

n

m s cover(R) ≤ Σm a=1 cover(Rs ) ≤ Σa=1 Σk |Ek |

≤ mn · (8d)

mn w

2

2

· 4n · m log n · wn mwn nn

3

2

≤ (8d)m 5 n · m4m 5 n

2

Before proving Claim 3.10 itself, and thus Lemma 3.9, we will need some notation. Definition 3.11 Let T1 , . . . , Tn ⊆ M (not necessarily disjoint). We say that a set of allocations R is (T1 , . . . , Tn )-structured if for all S ∈ R it holds that Si ⊆ Ti . Definition 3.12 We say that an allocation S is w-(i, j)-aligned in structure (T1 , . . . , Tn ), if |Si ∩ Tj | ≤ w. We omit w when it will be clear from the context. e which is “sufficiently” The idea in proving Claim 3.10 will be to find a large subset E ⊆ R, aligned. Next we show that such subset has a small cover number. Claim 3.13 Let E be a T = (T1 , . . . , Tn )-structured set of allocations. If for each pair of bidders i and j either all allocations in E are w-(i, j)-aligned in T or all allocations in E are w-(j, i)-aligned 2w

in T , then cover(E) ≤ wn mn Proof:

nn .

(of Claim 3.13) For each bidder i define a set of bidders Ii , where bidder j is in Ii if all

allocations in E are (i, j)-aligned in T . Observe that for each i and j, either j ∈ Ii or i ∈ Ij . Let Bi = Ti \ (∪j∈Ii Tj ). The construction guarantees that (B1 , . . . , Bn ) “almost” covers E in the sense that for bidder i and S ∈ E, |Si \ Bi | ≤ nw. Also notice that by construction for each two different bidders i and j, Bi ∩ Bj = ∅. Define the cover C as follows: C = {P |P is an allocation of the form ((B1 ∪ Q1 ) \ ∪j6=1 Qj , . . . , (Bn ∪ Qn ) \ ∪j6=n Qj ), and for each i, |Qi | ≤ nw}

14

Observe that since each |Qi | ≤ nw we have that |C| ≤ (Σnw r=1

¡m¢ n nw n n n2 w nn . r ) ≤ (wn(m) ) = w m

Also notice that C is a cover set of E. To see this, fix an allocation S ∈ E. For each i, let Qi = Si \ Bi . Observe that each |Qi | ≤ nw, and that the (Qi ∪ Bi )’s define an allocation that is in C and covers S.

Now we are ready to prove the main claim, and thus finish the proof of Lemma 3.9. Proof:

(of Claim 3.10) The construction of E will be divided into several steps. During the

e : R0 , R1 , . . . and structures T 0 , T 1 , . . ., such construction we maintain a sequence of subsets of R e and T 0 = (M, . . . , M ). that each Rt is T t -structured. We start by setting R0 = R In each step we look at a pair of bidders i and j such that neither all allocations in Rt are (i, j)aligned in T t nor all allocations in E are (j, i)-aligned in T t . If there is no such pair then let E = Rt and the construction is over. Otherwise, look at all allocations in Rt that are either (i, j)-aligned or (j, i)-aligned in T t . If there are at least |Rt |/2 such allocations then we set Rt+1 to be the largest set of the two: all allocations in Rt that are (i, j)-aligned, or all allocations in Rt that are (j, i)-aligned. Set the structure T t+1 to be T t . Notice that |Rt+1 | ≥ |Rt |/4, and that Rt+1 is T t+1 -structured. We call this step an alignment step, and proceed to the next step. Otherwise, let R0t be the set of allocations in Rt that are neither (i, j)-aligned nor (j, i)-aligned. Notice that |R0t ≥

Rt 2 |.

Take a maximal (i, j)-intersection set D ⊆ R0t – of size at most d. Now for

every allocation S ∈ R0t \ D there exists some D ∈ D such that Di ∩ Sj = ∅ or Dj ∩ Si = ∅. Otherwise we have that S ∈ D, contradicting the fact that D is a maximal intersection set. Thus, for some D ∈ D we have that for at least (|R0t | − d)/(2d) allocations in R0t either Di ∩ Sj = ∅ or Dj ∩ Si = ∅. Let us assume that for at least (|R0t | − d)/(2d) allocations in R0t the first option occurs. Define Rt+1 to be this set of (|R0t | − d)/(2d) ≥ |Rt |/(8d) allocations. Let Tjt+1 = Tjt \ Di . Also let Tkt+1 = Tkt , for each k 6= i. Now notice that since D is a set of allocations that are not (i, j)-aligned in Tt , we have that |Di ∩ Tjt | > w. We therefore have that |Tjt+1 | < |Tjt | − w. (The other case is handled similarly, but this time by shrinking Tit+1 rather than Tjt+1 .) By construction we have that Rt+1 is T t+1 -structured. Term this step a shrinkage step, and continue to the next step. Denote by l the number of steps the process went on. At most

nm w

steps are shrinkage steps, since ¡ ¢ in each shrinkage step Σi |Tit | loses an additive factor of at least w. In addition, there are at most n2 alignment steps, one for each pair of bidders. Therefore |E| = |Rl | ≥

15

e |R| 2 . (8d)mn/w 4(n )

Also note that

in the end of the process for each pair of bidders i and j either all allocations in E are (i, j)-aligned in T l or all allocations in E are (j, i)-aligned in T l (observe that an allocation that became properly aligned after an alignment step will remain so during the rest of the process.) By Claim 3.13 we have that cover(E) ≤ wn mn

2w

, and thus Claim 3.10 is proved.

This concludes the proof of Lemma 3.9.

3.5

Handling Non-Identical cS ’s

Now for the case where the cS ’s are not identical. Let G = {S|S ∈ arg maxS cS }. We claim that we can regard the M IR algorithm A as an M IR algorithm A0 with range G and identical cS ’s, as far as our lower bound is concerned. The idea is that if the cS ’s are sufficiently large comparing to the welfare of the best allocation, so the algorithm always outputs one of the allocations in G – the one with the highest weighted welfare among all allocations in G. Therefore, we can treat A as an M IR algorithm with range G, and the same proof goes through. A bit more formally, this is done as follows. Let t1 be the largest cS (so if T ∈ G then cT = t1 ), and let t2 be the second largest cS . Let t = t1 − t2 . Now divide all valuations that appear in Lemmas 3.4 and 3.8 the same constant c, so that for every instance that is considered in the proof of these lemmas we have that t > Σi vi (M ) > OP T , where OPT is the optimal weighted welfare in this instance (notice that in the proof we consider only a finite number of instances, so this scaling is possible). Observe that the MIR algorithm A chooses only allocations from G. Furthermore, observe that the proofs of the lemmas goes through, as all our arguments are clearly invariant to scaling. Hence the theorem holds for all VCG-based algorithms.

4

Reducing Randomized Algorithms to Deterministic “Good” Ones

The result of the next section clearly holds for deterministic MIR mechanisms by considering 1-good mechanisms. We now show that it applies also to randomized MIR mechanisms. The following proposition shows that: Proposition 4.1 Let U be some set of instances, and let A be a randomized algorithm that for every instance I provides an expected welfare of

OP T (I) , α

for some α > 1. Then, in the support of A there

16

is a (deterministic) Proof:

1 α -good

(on U , for every U ) α-approximation algorithm.

Fix some instance I ∈ U . The expected welfare A provides for this instance is

OP T (I) . α 1 α:

it

OP T (I) α

it

A lower bound on the probability that A provides an approximation ratio better than α is is achieved if for each toss of the random coins where A provides a welfare better than

actually finds the optimal solution, and provides a welfare of 0 for the rest of the tosses. Clearly, we must have now that there exists a deterministic algorithm B in the support of A that provides an approximation ratio of α for

1 α

of the instances in U .

As a corollary, to show that a randomized algorithm cannot provide an expected approximation ratio of are

1 α,

1 α -good

it is enough to prove a lower bound on deterministic

1 α -approximation

algorithms that

(on some set U ). Our main result, described in the previous section, does exactly that.

Acknowledgements We thank Liad Blumrosen and Nikhil Devanur for pointing out that the algorithms of [3] and [2] are maximal in range. We also thank Liad Blumrosen for comments on an earlier draft of this paper. This research was supported by a grant from the Israeli Academy of Sciences. The first author is supported by the Adams Fellowship Program of the Israel Academy of Sciences and Humanities.

References [1] A. Archer, C. Papadimitriou, K. Talwar, and E. Tardos. An approximate truthful mechanism for combinatorial auctions with single parameter agent. In SODA’03. [2] Sanjeev Arora. Polynomial time approximation schemes for Euclidean traveling salesman and other geometric problems. Journal of the ACM, 45(5):753–782, 1998. [3] Moshe Babaioff and Liad Blumrosen. Computationally-feasible auctions for convex bundles. In 7th. International Workshop on Approximation Algorithms for Combinatorial Optimization Problems (APPROX). LNCS Vol. 3122., pages 27–38, 2004. [4] Yair Bartal, Rica Gonen, and Noam Nisan. Incentive compatible multi unit combinatorial auctions. In TARK 03.

17

[5] Liad Blumrosen and Shahar Dobzinski. Welfare maximization in congestion games. In IEEE Journal on Selected Areas in Communications. Preliminary version in EC’06. [6] Liad Blumrosen and Noam Nisan. 2007. Combinatorial Auctions (a survey). In “Algorithmic Game Theory”, N. Nisan, T. Roughgarden, E. Tardos and V. Vazirani, editors. [7] E. H. Clarke. Multipart pricing of public goods. Public Choice, pages 17–33, 1971. [8] Peerapong Dhangwatnotai, Shahar Dobzinski, Shaddin Dughmi, and Tim Roughgarden. Truthful approximation schemes for single-parameter agents. In FOCS’08. [9] Shahar Dobzinski and Noam Nisan. Limitations of vcg-based mechanisms. Preliminary version in STOC’07. [10] Shahar Dobzinski and Noam Nisan. Mechanisms for multi-unit auctions. In EC’07. [11] Shahar Dobzinski, Noam Nisan, and Michael Schapira. Approximation algorithms for combinatorial auctions with complement-free bidders. In STOC’05. [12] Shahar Dobzinski, Noam Nisan, and Michael Schapira. Truthful randomized mechanisms for combinatorial auctions. In STOC’06. [13] Shahar Dobzinski and Michael Schapira. An improved approximation algorithm for combinatorial auctions with submodular bidders. In SODA’06. [14] Uriel Feige. On maximizing welfare where the utility functions are subadditive. In STOC’06. [15] Uriel Feige and Jan Vondrak. Approximation algorithms for allocation problems: Improving the factor of 1-1/e. In FOCS’06. [16] T. Groves. Incentives in teams. Econometrica, pages 617–631, 1973. [17] Ron Holzman, Noa Kfir-Dahav, Dov Monderer, and Moshe Tennenholtz. Bundling equilibrium in combinatrial auctions. Games and Economic Behavior, 47:104–123, 2004. [18] Subhash Khot, Richard J. Lipton, Evangelos Markakis, and Aranyak Mehta. Inapproximability results for combinatorial auctions with submodular utility functions. In WINE’05, 2005.

18

[19] Eyal Kushilevitz and Noam Nisan. Communication Complexity. Cambridge University Press, 1997. [20] Ron Lavi, Ahuva Mu’alem, and Noam Nisan. Towards a characterization of truthful combinatorial auctions. In FOCS’03. [21] Ron Lavi and Chaitanya Swamy. Truthful and near-optimal mechanism design via linear programming. In FOCS 2005. [22] Benny Lehmann, Daniel Lehmann, and Noam Nisan. Combinatorial auctions with decreasing marginal utilities. In EC’01. [23] Daniel Lehmann, Liadan Ita O’Callaghan, and Yoav Shoham. Truth revelation in approximately efficient combinatorial auctions. In JACM 49(5), pages 577–602, Sept. 2002. [24] A. Mas-Collel, W. Whinston, and J. Green. Microeconomic Theory. Oxford university press, 1995. [25] Ahuva Mu’alem and Noam Nisan. Truthful approximation mechanisms for restricted combinatorial auctions. In AAAI-02, 2002. [26] Noam Nisan. The communication complexity of approximate set packing and covering. In ICALP 2002. [27] Noam Nisan. 2007. Introduction to Mechanism Design (for Computer Scientists). In “Algorithmic Game Theory”, N. Nisan, T. Roughgarden, E. Tardos and V. Vazirani, editors. [28] Noam Nisan and Amir Ronen. Computationally feasible vcg-based mechanisms. In EC’00. [29] Noam Nisan and Ilya Segal. The communication requirements of efficient allocations and supporting prices, 2006. In the Journal of Economic Theory. [30] Kevin Roberts. The characterization of implementable choise rules. In Jean-Jacques Laffont, editor, Aggregation and Revelation of Preferences. Papers presented at the first European Summer Workshop of the Economic Society, pages 321–349. North-Holland, 1979.

19

[31] Tuomas Sandholm. Algorithm for optimal winner determination in combinatorial auctions. In Artificial Intelligence, volume 135, pages 1–54, 2002. [32] Vijay V. Vazirani. Approximation algorithms. Springer-Verlag New York, Inc., New York, NY, USA, 2001. [33] W. Vickrey. Counterspeculation, auctions and competitive sealed tenders. Journal of Finance, pages 8–37, 1961.

20

Limitations of VCG-Based Mechanisms

Sep 12, 2010 - when the VCG payment scheme is used: Definition: An algorithm (that produces an output a ∈ A for each input v1,...,vn, where A is the set of possible outcomes) is called “maximal-in-range” (henceforth MIR) if it completely optimizes the weighted social welfare over some subrange R⊆A. I.e., for some R⊆A ...

224KB Sizes 0 Downloads 299 Views

Recommend Documents

LIMITATIONS OF GRADIENT METHODS IN ...
iBP stands for BP using incremental learning, while. tBP using the trophic .... Foundation , Prentice-hall, New J ersey, second edi- tion 1999. [4] J .B. Pollack, ...

Mechanisms of
the community, will acquire language with great speed and facility; an ability that .... have relied on demanding methods to obtain the highly informative data base ..... are more adequate than vowels to subserve the storage of a large number of.

Democracy Against Neoliberalism - Paradoxes, Limitations ...
Page 1 of 22. Critical Sociology. 1–22. © The Author(s) 2014. Reprints and permissions: sagepub.co.uk/journalsPermissions.nav. DOI: 10.1177/0896920513507789. crs.sagepub.com. Democracy against Neoliberalism: Paradoxes, Limitations,. Transcendence.

lpm-limitations-rsl.pdf
... relative inconsistency. The four. resulting logics we call LPm,LP=,LP⊆, and LP⊇. © 2013 Association for Symbolic Logic. 1 doi:10.1017/S1755020300000000.

Democracy Against Neoliberalism - Paradoxes, Limitations ...
Whoops! There was a problem loading more pages. Whoops! There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Democracy Against Neoliberalism - Paradoxes, Limitat

Neural mechanisms of synergy formation *
activity, whose equilibrium configurations .... rnusculo-skeletal body schema that imple- ... S-units model the different skeletal body segments, considered as.

The Limitations of Stranger-Interviewers in Rural Kenya
Dec 31, 2007 - social relationship with respondents. ..... on the role of social networks in contraceptive .... viewer selection list (Gwassi 1995:19), although.

M1 Limitations of software and hardware used to manipulate graphic ...
M1 Limitations of software and hardware used to manipulate graphic images.pdf. M1 Limitations of software and hardware used to manipulate graphic images.

The Limitations of Stranger-Interviewers in Rural Kenya
build on one of two ideas about micro level social interaction: .... comfort was our growing awareness of two fac- tors. First ... scripts available at the project Web site) and countless .... survey if no one from their own immediate or extended ...

Relevance and Limitations of Crowding, Fractal, and ...
replication, recombination, and repair, but the molecular mechanisms of these transactions ... The best-studied chromosome component is the nucleosome .... of hard steric repulsion between macromolecules and by soft nonspecific interactions ...... up

impacts of property tax limitations on education spending
and Saas, 2006; Temple, 1996; Fisher and Gade,. 1991). One type of property tax ..... The tax rates are county- wide real property tax rates that apply to all non-.

impacts of property tax limitations on education spending
California restricts the growth of an individual's assessed. IMPACTS OF PROPERTY TAX LIMITATIONS .... changes in school spending, which are constrained.

Understanding the Mechanisms of Economic ...
if they are hard to explain in other ways, we attach additional credence to the mechanism ... who, in the second paper, took the predictions to international data. .... drive the life-cycle correlation, and in both cases, households of all ages appea

Mechanisms of Proinflammatory Cytokine- Induced ...
Because our data show that MEKK3 organizes signal- dependent ... In an in vitro system, MEKK3 phosphorylates the polypeptide mapping to the regulatory loop of ... The migration of molecular weight markers (in kilodaltons) is also indicated.

Molecular mechanisms of experience-dependent ...
allowing the recovery of neuronal response properties, such as .... in open-eye drive are not detected until after the weakening of ... to synaptic and network plasticity in visual cortex. (figure 2 ..... These data point to an important feed-forward

Modelling copepod development: current limitations ...
times may afford a selective advantage to individuals that can reduce or .... distribution parameters require sophisticated software packages ...... Current off northern Scotland. .... icus in the Gulf of St Lawrence and Scotian Shelf: study with a.

Performance limitations in sensitivity reduction for ... - ScienceDirect.com
ity optimization to a nonlinear time-varying setting. Keywords: Non-minimum phase systems; nonlinear systems; sensitivity reduction; performance limitations; disturbance re- jection. considered in [7,10,16]. For example it has been shown for plants w

Relevance and Limitations of Crowding, Fractal, and ...
recovery after photobleaching (FRAP) or any variant of this technique. (Bancaud .... 1978 (Jackson, 1978), can now provide genome wide contact maps of chro-.

M1 Limitations of software and hardware used to manipulate graphic ...
store). • Not user friendly (Like Photoshop starting off on. aperture as a new ... M1 Limitations of software and hardware used to manipulate graphic images.pdf.

NHTSA Traffic Safety Facts - Limitations of Drug Testing in Fatals ...
NHTSA Traffic Safety Facts - Limitations of Drug Testing in Fatals November 2014 (2).pdf. NHTSA Traffic Safety Facts - Limitations of Drug Testing in Fatals ...

The potential and limitations of using carbon dioxide - Royal Society
A number of companies are already exploring these areas. It is likely that research currently underway will ... BEIS (Department for Business Energy & Industrial Strategy). 2015 Data ... products; and from approximately 1800Mt9 to 2000Mt10 of ...

Potential and Limitations of Pro-Poor Macroeconomics
and penalizes small and medium-sized enterprises (SMEs). ..... factor behind this trend is the pro-cyclical nature of interest rates and spreads that reduces ... infrastructural support, a tax free status and public guarantees for loans, or relying o

Neurocognitive mechanisms of cognitive control - Semantic Scholar
Convergent evidence highlights the differential contributions of various regions of the prefrontal cortex in the service of cognitive control, but ..... Presentation of a stop-signal yields a rapid increase in activity in FEF gaze-holding neurons and