summaryrefslogtreecommitdiffstats
path: root/appendix.tex
blob: fc32f10c42fdead9d6ccc9066ffded7a4cc5f1bd (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
\section{Properties of the Value Function $V$} \label{app:properties}
For the sake of concreteness, we prove below the positivity, monotonicity, and submodularity of $V(S) = \log\det(I_d+X_S^TX_S)$ from basic principles. We note however that these properties hold more generally for the information gain under a wider class of models than the linear model with Gaussian noise and prior that we study here: we discuss this in more detail in Appendix~\ref{sec:ext}.

For two symmetric matrices $A$ and $B$, we write $A\succ B$ ($A\succeq B$) if
$A-B$ is positive definite (positive semi-definite).  This order allows us to
define the notion of a \emph{decreasing} as well as  \emph{convex} matrix
function, similarly to their real counterparts. With this definition, matrix
inversion is decreasing and convex over symmetric positive definite
matrices (see Example 3.48 p. 110 in \cite{boyd2004convex}).

Recall that the determinant of a matrix equals the product of its eigenvalues. The positivity of $V(S)$ follows from the fact that $X_S^TX_S$ is positive semi-definite and, as such $I_d+X_S^TX_S\succeq I_d$, so all its eigenvalues are larger than or equal to one, and they are all one if $S=\emptyset$. The marginal contribution of item $i\in\mathcal{N}$ to set
    $S\subseteq \mathcal{N}$ can be written as 
\begin{align}
V(S\cup \{i\}) - V(S)& =   \frac{1}{2}\log\det(I_d 
    + \T{X_S}X_S + x_i\T{x_i})
    - \frac{1}{2}\log\det(I_d + \T{X_S}X_S)\nonumber\\
    &  = \frac{1}{2}\log\det(I_d + x_i\T{x_i}(I_d +
\T{X_S}X_S)^{-1})\nonumber\\
&= \frac{1}{2}\log(1 + \T{x_i}A(S)^{-1}x_i)\label{eq:marginal_contrib}
\end{align}
where $A(S) \defeq I_d+ \T{X_S}X_S$, and the last equality follows from the
Sylvester's determinant identity~\cite{sylvester}. Monotonicity therefore follows from the fact that $A(S)^{-1}$ is positive semidefinite. Finally, since the inverse is decreasing over positive definite matrices, we have
\begin{gather}
        \forall S\subseteq\mathcal{N},\quad A(S)^{-1} \succeq A(S\cup\{i\})^{-1}. \label{eq:inverse}
\end{gather}
and submodularity also follows, as a function is submodular if and only if the marginal contributions are non-increasing in $S$. \qed


%\begin{proof}
%\end{proof}

%We now prove that $F$ admits the following exchange property: let $\lambda$ be a feasible element of $[0,1]^n$, it is possible to trade one fractional component of $\lambda$ for another until one of them becomes integral, obtaining a new element $\tilde{\lambda}$ which is both feasible and for which $F(\tilde{\lambda})\geq F(\lambda)$. Here, by feasibility of a point $\lambda$, we mean that it satisfies the budget constraint $\sum_{i=1}^n \lambda_i c_i \leq B$.  This rounding property is referred to in the literature as \emph{cross-convexity} (see, \emph{e.g.}, \cite{dughmi}), or $\varepsilon$-convexity by \citeN{pipage}.

%\begin{lemma}[Rounding]\label{lemma:rounding}
%    For any feasible $\lambda\in[0,1]^{n}$, there exists a feasible
%    $\bar{\lambda}\in[0,1]^{n}$ such that at most one of its components is
%    fractional %, that is, lies in $(0,1)$ and:
%     and $F_{\mathcal{N}}(\lambda)\leq F_{\mathcal{N}}(\bar{\lambda})$.
%\end{lemma}
%\begin{proof}
%\end{proof}

%The $\log\det$ function is concave and self-concordant (see
%\cite{boyd2004convex}), in this case, the analysis of the barrier method in
%in \cite{boyd2004convex} (Section 11.5.5) can be summarized in the following
%lemma:

%\begin{lemma}\label{lemma:barrier}
%For any $\varepsilon>0$, the barrier method computes an $\varepsilon$-accurate
%approximation of $L^*_c$ in time $O(poly(n,d,\log\log\varepsilon^{-1})$.
%\end{lemma}
%Note, that the feasible set in Problem~\eqref{eq:primal} increases (for the
%inclusion) when the cost decreases. 
%non-increasing.

%Furthermore, \eqref{eq:primal} being a convex optimization problem, using
%standard convex optimization algorithms (Lemma~\ref{lemma:barrier} gives
%a formal statement for our specific problem) we can compute
%a $\varepsilon$-accurate approximation of its optimal value as defined below.

%\begin{definition}
%$a$ is an $\varepsilon$-accurate approximation of $b$ iff $|a-b|\leq \varepsilon$.
%\end{definition}

%Note however that an $\varepsilon$-accurate approximation of a non-increasing
%function is not in general non-increasing itself. The goal of this section is
%to approximate $L^*_c$ while preserving monotonicity. The estimator we
%construct has a weaker form of monotonicity that we call
%\emph{$\delta$-monotonicity}.

%\begin{definition}
%Let $f$ be a function from $\reals^n$ to $\reals$, we say that $f$ is
%\emph{$\delta$-increasing along the $i$-th coordinate} iff:
%\begin{equation}\label{eq:dd}
%    \forall x\in\reals^n,\;
%    \forall \mu\geq\delta,\;
%    f(x+\mu e_i)\geq f(x)
%\end{equation}
%where $e_i$ is the $i$-th canonical basis vector of $\reals^n$. By extension,
%$f$ is $\delta$-increasing iff it is $\delta$-increasing along all coordinates.

%We define \emph{$\delta$-decreasing} functions by reversing the inequality in
%\eqref{eq:dd}.
%\end{definition}

%We consider a perturbation of \eqref{eq:primal} by introducing:
%\begin{equation}\tag{$P_{c, \alpha}$}\label{eq:perturbed-primal}
%    L^*_{c,\alpha} \defeq \max_{\lambda\in[\alpha, 1]^{n}}
%    \left\{L(\lambda) \Big| \sum_{i=1}^{n} \lambda_i c_i
%    \leq B\right\}
%\end{equation}
%Note that we have $L^*_c = L^*_c(0)$.  We will also assume that
%$\alpha<\frac{1}{nB}$ so that \eqref{eq:perturbed-primal} has at least one
%feasible point: $(\frac{1}{nB},\ldots,\frac{1}{nB})$. By computing
%an approximation of $L^*_{c,\alpha}$ as in Algorithm~\ref{alg:monotone}, we
%obtain a $\delta$-decreasing approximation of $L^*_c$.

\section{Budget Feasible Reverse Auction Mechanisms}\label{app:budgetfeasible}
We review in this appendix the formal definition of a budget feasible reverse auction mechanisms, as introduced by \citeN{singer-mechanisms}. We depart from the definitions in \cite{singer-mechanisms} only in considering $\delta$-truthful, rather than truthful, mechanisms.

Given a budget $B$ and a value function $V:2^{\mathcal{N}}\to\reals_+$, a \emph{mechanism} $\mathcal{M} = (S,p)$ comprises (a) an \emph{allocation function} 
$S:\reals_+^n \to 2^\mathcal{N}$ and (b) a \emph{payment function}
$p:\reals_+^n\to \reals_+^n$. 
 Let $s_i(c) = \id_{i\in S(c)}$ be the binary indicator of  $i\in S(c)$. We seek mechanisms that have the following properties \cite{singer-mechanisms}:
\begin{itemize}
\item \emph{Normalization.} Unallocated experiments receive zero payments: $s_i(c)=0$ implies $p_i(c)=0.\label{normal}$
\item\emph{Individual Rationality.} Payments for allocated experiments exceed
costs: $p_i(c)\geq c_i\cdot s_i(c).\label{ir}$
\item \emph{No Positive Transfers.} Payments are non-negative: $p_i(c)\geq 0\label{pt}$.
\item \emph{$\delta$-Truthfulness.} Reporting one's true cost is
an \emph{almost-dominant} \cite{schummer2004almost} strategy. Formally, let $c_{-i}$
 be a vector of costs of all agents except $i$. Then, $p_i(c_i,c_{-i})
 - s_i(c_i,c_{-i})\cdot c_i \geq p_i(c_i',c_{-i}) - s_i(c_i',c_{-i})\cdot c_i,
 $ for every $i \in \mathcal{N}$ and every two cost vectors $(c_i,c_{-i})$ and $(c_i',c_{-i})$ such that $|c_i-c_i'|>\delta.$ The mechanism is \emph{truthful} if $\delta=0$.
 \item \emph{Budget Feasibility.} The sum of the payments should not exceed the
    budget constraint, \emph{i.e.} $\sum_{i\in\mathcal{N}} p_i(c) \leq
    B.\label{budget}$


   \item \emph{$(\alpha,\beta)$-approximation.} The value of the allocated set  should not
    be too far from the optimum value of the full information case, as given by  \eqref{eq:non-strategic}. 
Formally, there must exist some $\alpha\geq 1$ and $\beta>0$
    such that $OPT \leq \alpha V(S(p))+\beta$, where     $OPT = \max_{S\subseteq\mathcal{N}} \left\{V(S) \;\mid \;
    \sum_{i\in S}c_i\leq B\right\}$.

 %   The approximation ratio captures the \emph{price of truthfulness}, \emph{i.e.},  the relative value loss incurred by adding the truthfulness constraint. % to the value function maximization.
    %We stress here that the approximation ratio is defined with respect \emph{to the maximal value in the full information case}.
    \item \emph{Computational Efficiency.} The allocation and payment function
    should be computable in time polynomial in various parameters. 
    %time in the number of agents $n$. %\thibaut{Should we say something about the black-box model for $V$?  Needed to say something in general, but not in our case where the value function can be computed in polynomial time}.
\end{itemize}

%
%Instead, \citeN{singer-mechanisms} and \citeN{chen}
%suggest to approximate $OPT_{-i^*}$ by a quantity $OPT'_{-i^*}$ satisfying the
%following properties:
%\begin{itemize}
%    \item $OPT'_{-i^*}$ must not depend on agent $i^*$'s reported cost and must
%    be decreasing with respect to the costs. This is to ensure the monotonicity
%    of the allocation function.
%    \item $OPT'_{-i^*}$ must be close to $OPT_{-i^*}$ to maintain a bounded
%    approximation ratio.
%    \item $OPT'_{-i^*}$ must be computable in polynomial time.
%\end{itemize}
%
%One of the main technical contributions of \citeN{chen} and
%\citeN{singer-influence} is to come up with appropriate such quantity by
%considering relaxations of \textsc{Knapsack} and \textsc{Coverage},
%respectively.
%
%\subsection{Our Approach}
%
%Using the results from Section~\ref{sec:approximation}, we come up with
%a suitable approximation $OPT_{-i^*}'$ of $OPT_{-i^*}$. Our approximation being
%$\delta$-decreasing, the resulting mechanism will be $\delta$-truthful as
%defined below.
%
%\begin{definition}
%Let $\mathcal{M}= (S, p)$ a mechanism, and let $s_i$ denote the indicator
%function of $i\in S(c)$, $s_i(c) \defeq \id_{i\in S(c)}$. We say that $\mathcal{M}$ is
%$\delta$-truthful iff:
%\begin{displaymath}
%\forall c\in\reals_+^n,\;
%\forall\mu\;\text{with}\; |\mu|\geq\delta,\;
%\forall i\in\{1,\ldots,n\},\;
%p(c)-s_i(c)\cdot c_i \geq p(c+\mu e_i) - s_i(c+\mu e_i)\cdot c_i
%\end{displaymath}
%where $e_i$ is the $i$-th canonical basis vector of $\reals^n$.
%\end{definition}
%
%Intuitively, $\delta$-truthfulness implies that agents have no incentive to lie
%by more than $\delta$ about their reported costs. In practical applications,
%the bids being discretized, if $\delta$ is smaller than the discretization
%step, the mechanism will be truthful in effect.

%$\delta$-truthfulness will follow from $\delta$-monotonicity by the following
%variant of Myerson's theorem.