\subsection{D-Optimality Criterion} Ideally, motivated by the $D$-optimality criterion, we would like to design a mechanism that maximizes \eqref{dcrit} within a good approximation ratio. As \eqref{dcrit} may take arbitrarily small negative values, to define a meaningful approximation we consider the (equivalent) maximization of $V(S) = f(\det\T{X_S}X_S )$, for some strictly increasing, on-to function $f:\reals_+\to\reals_+$. However, the following lower bound implies that such an optimization goal cannot be attained under the costraints of truthfulness, budget feasibility, and individional rationallity. \begin{lemma} For any $M>1$, there is no $M$-approximate, truthful, budget feasible, individionally rational mechanism for budget feasible experiment design with value fuction $V(S) = \det{\T{X_S}X_S}$. \end{lemma} \begin{proof} \input{proof_of_lower_bound1} \end{proof} This negative result motivates us into looking at the following modified objective: \begin{align}V(S) = \log\det(I_d+\T{X_S}X_S), \label{modified}\end{align} where $I_d\in \reals^{d\times d}$ is the identity matrix. One possible interpretation of \eqref{modified} is that, prior to the auction, the experimenter has free access to $d$ experiments whose features form an ortho-normal basis in $\reals^d$. However, \eqref{modified} can also be motivated in the context of \emph{Bayesian experimental design} \cite{chaloner1995bayesian}. In short, the objective \eqref{modified} arises naturally when the experimenter retrieves the model $\beta$ through \emph{ridge regression}, rather than the linear regression \eqref{leastsquares} over the observed data; we explore this connection in Section~\ref{sec:bed}. From a practical standpoint, \eqref{modified} is a good approximation of \eqref{dcrit} when the number of experiments is large. Crucially, \eqref{modified} is submodular and satifies $V(\emptyset) = 0$, allowing us to use the extensive machinery for the optimization of submodular functions, as well as recent results in the context of budget feasible auctions. \subsection{Truthful, Constant Approximation Mechanism} In this section we present a mechanism for \EDP. Previous works on maximizing submodular functions \cite{nemhauser, sviridenko-submodular} and designing auction mechanisms for submodular utility functions \cite{singer-mechanisms, chen, singer-influence} rely on a greedy heuristic. In this heuristic, points are added to the solution set according to the following greedy selection rule: assume that you have already selected a set $S$ of point, then the next point to be selected is: \begin{displaymath} i = \argmax_{j\in\mathcal{N}\setminus S}\frac{V(S\cup\{i\}) - V(S)}{c_i} \end{displaymath} This is the generalization of the \emph{value-per-cost} ratio used in greedy heuristic for knapsack problems. However note that for general submodular functions, the value of a point depends on the set of points which have already been selected. Unfortunately, even for the non-strategic case, the greedy heuristic gives an unbounded approximation ratio. It has been noted by Khuller et al. \cite{khuller} that for the maximum coverage problem, taking the maximum between the greedy solution and the point of maximum value gives a $\frac{2e}{e-1}$ approximation ratio. In the general case, let us recall lemma 3.1 from \cite{singer-influence} which follows from \cite{chen}: \begin{lemma}[Singer \cite{singer-influence}]\label{lemma:greedy-bound} Let $S_G$ be the set computed by the greedy heuristic and $i^*$ the point of maximum value: \begin{displaymath} i^* = \argmax_{i\in\mathcal{N}} V(i) \end{displaymath} then the following inequality holds: \begin{displaymath} OPT(V,\mathcal{N},B) \leq \frac{e}{e-1}\big( 3 V(S_G) + 2 V(i^*)\big) \end{displaymath} \end{lemma} Hence, taking the maximum between the greedy set and the point of maximum value yields an approximation ratio of $\frac{5e}{e-1}$. However, Singer \cite{singer-influence} notes that this approach breaks incentive compatibility and therefore cannot be directly applied to the strategic case. Two approaches have been studied to deal with the strategic case and rely on comparing the point of maximum value to a quantity which can be proven to be not too far from the greedy solution and maintains incentive compatibility. \begin{itemize} \item In \cite{chen}, the authors suggest using $OPT(V,\mathcal{N}\setminus\{i^*\}, B)$ where $i^*$ is the point of maximum value. While this yields an approximation ratio of 8.34, in the general case, the optimal value cannot be computed in polynomial time. \item For the set coverage problem, Singer \cite{singer-influence} uses a relaxation of the value function which can be proven to have a constant approximation ratio to the value function. \end{itemize} Here, we will use the following relaxation of the value function \eqref{vs}. Let us define the function $L_\mathcal{N}$: \begin{displaymath} \forall\lambda\in[0,1]^{|\mathcal{N}|}\,\quad L_{\mathcal{N}}(\lambda) \defeq \log\det\left(I_d + \sum_{i\in\mathcal{N}} \lambda_i x_i \T{x_i}\right) \end{displaymath} Our mechanism for \EDP{} is presented in Algorithm~\ref{mechanism}. \begin{algorithm} \caption{Mechanism for \EDP{}}\label{mechanism} \begin{algorithmic}[1] \State $i^* \gets \argmax_{j\in\mathcal{N}}V(j)$ \State $x^* \gets \argmax_{x\in[0,1]^{|\mathcal{N}|}} \{L_{\mathcal{N}\setminus\{i^*\}}(x) \mid c(x)\leq B\}$ \Statex \If{$L(x^*) < CV(i^*)$} \State \textbf{return} $\{i^*\}$ \Else \State $i \gets \argmax_{1\leq j\leq n}\frac{V(j)}{c_j}$ \State $S_G \gets \emptyset$ \While{$c_i\leq \frac{B}{2}\frac{V(S_G\cup\{i\})-V(S_G)}{V(S_G\cup\{i\})}$} \State $S_G \gets S_G\cup\{i\}$ \State $i \gets \argmax_{j\in\mathcal{N}\setminus S_G} \frac{V(S_G\cup\{j\})-V(S_G)}{c_j}$ \EndWhile \State \textbf{return} $S_G$ \EndIf \end{algorithmic} \end{algorithm} \emph{Remarks} \begin{enumerate} \item the function $L_\mathcal{N}$ is concave (see lemma~\ref{lemma:concave}) thus the maximization step on line 2 of the mechanism can be computed in polynomial time, which proves that the mechanism overall has a polynomial complexity. \item the stopping rule in the while loop is more sophisticated than just checking against the budget constraint ($c(S) \leq B$). This is to ensure budget feasibility (see lemma~\ref{lemma:budget-feasibility}). \end{enumerate} We can now state the main result of this section: \begin{theorem}\label{thm:main} The mechanism described in Algorithm~\ref{mechanism} is truthful, individually rational and budget feasible. Furthermore, choosing: \begin{displaymath} C = C^* = \frac{12e-1 + \sqrt{160e^2-48e + 9}}{2(e-1)} \end{displaymath} we get an approximation ratio of: \begin{displaymath} 1 + C^* = \frac{14e-3 + \sqrt{160e^2-48e + 9}}{2(e-1)}\simeq 18.68 \end{displaymath} \end{theorem} \begin{proof} \emph{Truthfulness.} The algorithm only describes the allocation rule. However, it suffices to prove that the mechanism is monotone, then Myerson's theorem (see theorem~\ref{thm:myerson}) ensures us that by paying each allocated user his threshold payment yields a truthful mechanism. The proof of the monotonicity has already been done in \cite{singer-influence} and is given here in lemma~\ref{lemma:monotone} below for the sake of completeness. \emph{Budget feasibility.} Thanks to the analysis of the threshold payment in \cite{chen}, the budget feasibility follows easily. The proof is given in lemma~\ref{lemma:budget-feasibility} below. \emph{Approximation ratio.} The proof of the approximation ratio follows the same path as in \cite{chen} and is done in lemma~\ref{lemma:approx}. Our main contribution is to prove that the relaxation $L_\mathcal{N}$ has a constant approximation ratio to the optimal solution (lemma~\ref{lemma:relaxation}). This follows from the \emph{pipage rounding} method described in \cite{pipage} where $L_\mathcal{N}$ is first compared to the \emph{multilinear relaxation} of the value function (lemma~\ref{lemma:relaxation-ratio}) for which it is possible to round solutions (making fractional components integral) while maintaining feasibility (lemma~\ref{lemma:rounding}). \end{proof} \begin{lemma}\label{lemma:monotone} The mechanism is monotone. \end{lemma} \begin{proof} We assume by contradiction that there exists a user $i$ that has been selected by the mechanism and that would not be selected had he reported a cost $c_i'\leq c_i$ (all the other costs staying the same). If $i\neq i^*$ and $i$ has been selected, then we are in the case where $L(x^*) \geq C V(i^*)$ and $i$ was included in the result set by the greedy part of the mechanism. By reporting a cost $c_i'\leq c_i$, using the submodularity of $V$, we see that $i$ will satisfy the greedy selection rule: \begin{displaymath} i = \argmax_{j\in\mathcal{N}\setminus S} \frac{V(S\cup\{j\}) - V(S)}{c_j} \end{displaymath} in an earlier iteration of the greedy heuristic. Let us denote by $S_i$ (resp. $S_i'$) the set to which $i$ is added when reporting cost $c_i$ (resp. $c_i'$). We have $S_i'\subseteq S_i$. Moreover: \begin{align*} c_i' & \leq c_i \leq \frac{B}{2}\frac{V(S_i\cup\{i\})-V(S_i)}{V(S_i\cup\{i\})}\\ & \leq \frac{B}{2}\frac{V(S_i'\cup\{i\})-V(S_i')}{V(S_i'\cup\{i\})} \end{align*} Hence $i$ will still be included in the result set. If $i = i^*$, $i$ is included iff $L(x^*) \leq C V(i^*)$. Reporting $c_i'$ instead of $c_i$ does not change the value $V(i^*)$ nor $L(x^*)$ (which is computed over $\mathcal{N}\setminus\{i^*\}$). Thus $i$ is still included by reporting a different cost. \end{proof} \begin{lemma}\label{lemma:budget-feasibility} The mechanism is budget feasible. \end{lemma} \begin{proof} Let us denote by $S_G$ the set selected by the greedy heuristic in the mechanism of Algorithm~\ref{mechanism}. Let $i\in S_G$, let us also denote by $S_i$ the solution set that was selected by the greedy heuristic before $i$ was added. Recall from \cite{chen} that the following holds for any submodular function: if the point $i$ was selected by the greedy heuristic, then: \begin{equation}\label{eq:budget} c_i \leq \frac{V(S_i\cup\{i\}) - V(S)}{V(S_G)} B \end{equation} Assume now that our mechanism selects point $i^*$. In this case, his payment his $B$ and the mechanism is budget-feasible. Otherwise, the mechanism selects the set $S_G$. In this case, \eqref{eq:budget} shows that the threshold payment of user $i$ is bounded by: \begin{displaymath} \frac{V(S_i\cup\{i\}) - V(S_i)}{V(S_G)} B \end{displaymath} Hence, the total payment is bounded by: \begin{displaymath} \sum_{i\in S_M} \frac{V(S_i\cup\{i\}) - V(S_i)}{V(S_G)} B \leq B \end{displaymath} \end{proof} \begin{lemma}\label{lemma:approx} Let $S^*$ be the set allocated by the mechanism. Let us write: \begin{displaymath} C_{\textrm{max}} = \max\left(1+C,\frac{3e}{e-1}\left( 1 + \frac{8e}{C (e-1) -10e +2}\right)\right) \end{displaymath} Then: \begin{displaymath} OPT(V, \mathcal{N}, B) \leq C_\text{max}\cdot V(S^*) \end{displaymath} \end{lemma} \begin{proof} If the condition on line 3 of the algorithm holds, then: \begin{displaymath} V(i^*) \geq \frac{1}{C}L(x^*) \geq \frac{1}{C}OPT(V,\mathcal{N}\setminus\{i\}, B) \end{displaymath} But: \begin{displaymath} OPT(V,\mathcal{N},B) \leq OPT(V,\mathcal{N}\setminus\{i\}, B) + V(i^*) \end{displaymath} Hence: \begin{displaymath} V(i^*) \geq \frac{1}{C+1} OPT(V,\mathcal{N}, B) \end{displaymath} If the condition of the algorithm does not hold, by applying lemmas \ref{lemma:relaxation} and \ref{lemma:greedy-bound}: \begin{align*} V(i^*) & \leq \frac{1}{C}L(x^*) \leq \frac{1}{C} \big(4 OPT(V,\mathcal{N}, B) + 2 V(i^*)\big)\\ & \leq \frac{1}{C}\left(\frac{4e}{e-1}\big(3 V(S_G) + 2 V(i^*)\big) + V(i^*)\right) \end{align*} Thus: \begin{align*} V(i^*) \leq \frac{12e}{C(e-1)- 10e + 2} V(S_G) \end{align*} Finally, using again lemma~\ref{lemma:greedy-bound}, we get: \begin{displaymath} OPT(V, \mathcal{N}, B) \leq \frac{3e}{e-1}\left( 1 + \frac{8e}{C (e-1) -10e +2}\right) V(S_G)\qed \end{displaymath} \end{proof} The optimal value for $C$ is: \begin{displaymath} C^* = \argmin_C C_{\textrm{max}} \end{displaymath} This equation has two solutions. Only one of those is such that: \begin{displaymath} C(e-1) -10e +2 \geq 0 \end{displaymath} which is needed in the proof of the previous lemma. Computing this solution, gives the result of the theorem. \subsection{An approximation ratio for $L_\mathcal{N}$}\label{sec:relaxation} Our main contribution is lemma~\ref{lemma:relaxation} which gives an approximation ratio for $L_\mathcal{N}$ and is key in the proof of theorem~\ref{thm:main}. We say that $R_\mathcal{N}:[0,1]^{|\mathcal{N}|}\rightarrow\reals$ is a relaxation of the value function $V$ over $\mathcal{N}$ if it coincides with $V$ at binary points. Formally, for any $S\subseteq\mathcal{N}$, let $\mathbf{1}_S$ denote the indicator vector of $S$. $R_\mathcal{N}$ is a relaxation of $V$ over $\mathcal{N}$ iff: \begin{displaymath} \forall S\subseteq\mathcal{N},\; R_\mathcal{N}(\mathbf{1}_S) = V(S) \end{displaymath} We can extend the optimisation problem defined above to a relaxation by extending the cost function: \begin{displaymath} \forall \lambda\in[0,1]^{|\mathcal{N}|},\; c(\lambda) = \sum_{i\in\mathcal{N}}\lambda_ic_i \end{displaymath} The optimisation problem becomes: \begin{displaymath} OPT(R_\mathcal{N}, B) = \max_{\lambda\in[0,1]^{|\mathcal{N}|}}\left\{R_\mathcal{N}(\lambda)\,|\, c(\lambda)\leq B\right\} \end{displaymath} The relaxations we will consider here rely on defining a probability distribution over subsets of $\mathcal{N}$. Let $\lambda\in[0,1]^{|\mathcal{N}|}$, let us define: \begin{displaymath} P_\mathcal{N}^\lambda(S) = \prod_{i\in S}\lambda_i \prod_{i\in\mathcal{N}\setminus S}(1-\lambda_i) \end{displaymath} $P_\mathcal{N}^\lambda(S)$ is the probability of picking the set $S$ if we select a subset of $\mathcal{N}$ at random by deciding independently for each point to include it in the set with probability $\lambda_i$ (and to exclude it with probability $1-\lambda_i$). For readability, let us define $A(S) = I_d + \T{X_S}X_S$, so that the value function is $V(S) = \log\det A(S)$. We have already defined the function $L_\mathcal{N}$ which is a relaxation of the value function. Note that it is possible to express it in terms of the probabilities $P_\mathcal{N}$: \begin{align*} L_{\mathcal{N}}(\lambda) & = \log\det \mathbb{E}_{S\sim P_\mathcal{N}^\lambda}\big[A(S)\big]\\ & = \log\det\left(\sum_{S\subseteq N} P_\mathcal{N}^\lambda(S)A(S)\right)\\ & = \log\det\left(I_d + \sum_{i\in\mathcal{N}} \lambda_ix_i\T{x_i}\right)\\ & \defeq \log\det \tilde{A}(\lambda) \end{align*} We will also use the \emph{multi-linear extension}: \begin{align*} F_\mathcal{N}(\lambda) & = \mathbb{E}_{S\sim P_\mathcal{N}^\lambda}\big[\log\det A(S)\big]\\ & = \sum_{S\subseteq\mathcal{N}} P_\mathcal{N}^\lambda(S) V(S)\\ & = \sum_{S\subseteq\mathcal{N}} P_\mathcal{N}^\lambda(S) \log\det A(S)\\ \end{align*} \begin{lemma}\label{lemma:concave} The \emph{concave relaxation} $L_\mathcal{N}$ is concave\footnote{Hence this relaxation is well-named!}. \end{lemma} \begin{proof} This follows from the concavity of the $\log\det$ function over symmetric positive semi-definite matrices. More precisely, if $A$ and $B$ are two symmetric positive semi-definite matrices, then: \begin{multline*} \forall\alpha\in [0, 1],\; \log\det\big(\alpha A + (1-\alpha) B\big)\\ \geq \alpha\log\det A + (1-\alpha)\log\det B \end{multline*} \end{proof} It has been observed already (see for example \cite{dughmi}) that the multi-linear extension presents the cross-convexity property: it is convex along any direction $e_i-e_j$ where $e_i$ and $e_j$ are two elements of the standard basis. This property allows to trade between two fractional components of a point without diminishing the value of the relaxation. The following lemma follows is inspired by a similar lemma from \cite{dughmi} but also ensures that the points remain feasible during the trade. \begin{lemma}[Rounding]\label{lemma:rounding} For any feasible $\lambda\in[0,1]^{|\mathcal{N}|}$, there exists a feasible $\bar{\lambda}\in[0,1]^{|\mathcal{N}|}$ such that at most one of its component is fractional, that is, lies in $(0,1)$ and: \begin{displaymath} F_{\mathcal{N}}(\lambda)\leq F_{\mathcal{N}}(\bar{\lambda}) \end{displaymath} \end{lemma} \begin{proof} We give a rounding procedure which given a feasible $\lambda$ with at least two fractional components, returns some $\lambda'$ with one less fractional component, feasible such that: \begin{displaymath} F_\mathcal{N}(\lambda) \leq F_\mathcal{N}(\lambda') \end{displaymath} Applying this procedure recursively yields the lemma's result. Let us consider such a feasible $\lambda$. Let $i$ and $j$ be two fractional components of $\lambda$ and let us define the following function: \begin{displaymath} F_\lambda(\varepsilon) = F(\lambda_\varepsilon) \quad\textrm{where} \quad \lambda_\varepsilon = \lambda + \varepsilon\left(e_i-\frac{c_i}{c_j}e_j\right) \end{displaymath} It is easy to see that if $\lambda$ is feasible, then: \begin{multline}\label{eq:convex-interval} \forall\varepsilon\in\Big[\max\Big(-\lambda_i,(\lambda_j-1)\frac{c_j}{c_i}\Big), \min\Big(1-\lambda_i, \lambda_j \frac{c_j}{c_i}\Big)\Big],\;\\ \lambda_\varepsilon\;\;\textrm{is feasible} \end{multline} Furthermore, the function $F_\lambda$ is convex, indeed: \begin{align*} F_\lambda(\varepsilon) & = \mathbb{E}_{S'\sim P_{\mathcal{N}\setminus\{i,j\}}^\lambda(S')}\Big[ (\lambda_i+\varepsilon)\Big(\lambda_j-\varepsilon\frac{c_i}{c_j}\Big)V(S'\cup\{i,j\})\\ & + (\lambda_i+\varepsilon)\Big(1-\lambda_j+\varepsilon\frac{c_i}{c_j}\Big)V(S'\cup\{i\})\\ & + (1-\lambda_i-\varepsilon)\Big(\lambda_j-\varepsilon\frac{c_i}{c_j}\Big)V(S'\cup\{j\})\\ & + (1-\lambda_i-\varepsilon)\Big(1-\lambda_j+\varepsilon\frac{c_i}{c_j}\Big)V(S')\Big]\\ \end{align*} Thus, $F_\lambda$ is a degree 2 polynomial whose dominant coefficient is: \begin{multline*} \frac{c_i}{c_j}\mathbb{E}_{S'\sim P_{\mathcal{N}\setminus\{i,j\}}^\lambda(S')}\Big[ V(S'\cup\{i\})+V(S'\cup\{i\})\\ -V(S'\cup\{i,j\})-V(S')\Big] \end{multline*} which is positive by submodularity of $V$. Hence, the maximum of $F_\lambda$ over the interval given in \eqref{eq:convex-interval} is attained at one of its limit, at which either the $i$-th or $j$-th component of $\lambda_\varepsilon$ becomes integral. \end{proof} \begin{lemma}\label{lemma:relaxation-ratio} The following inequality holds: \begin{displaymath} \forall\lambda\in[0,1]^{|\mathcal{N}|},\; \frac{1}{2} \,L_\mathcal{N}(\lambda)\leq F_\mathcal{N}(\lambda)\leq L_{\mathcal{N}}(\lambda) \end{displaymath} \end{lemma} \begin{proof} We will prove that $\frac{1}{2}$ is a lower bound of the ratio $\partial_i F_\mathcal{N}(\lambda)/\partial_i L_\mathcal{N}(\lambda)$. This will be enough to conclude, by observing that: \begin{displaymath} \frac{F_\mathcal{N}(\lambda)}{L_\mathcal{N}(\lambda)} \sim_{\lambda\rightarrow 0} \frac{\sum_{i\in \mathcal{N}}\lambda_i\partial_i F_\mathcal{N}(0)} {\sum_{i\in\mathcal{N}}\lambda_i\partial_i L_\mathcal{N}(0)} \end{displaymath} and that an interior critical point of the ratio $F_\mathcal{N}(\lambda)/L_\mathcal{N}(\lambda)$ is characterized by: \begin{displaymath} \frac{F_\mathcal{N}(\lambda)}{L_\mathcal{N}(\lambda)} = \frac{\partial_i F_\mathcal{N}(\lambda)}{\partial_i L_\mathcal{N}(\lambda)} \end{displaymath} Let us start by computing the derivatives of $F_\mathcal{N}$ and $L_\mathcal{N}$ with respect to the $i$-th component. For $F$, it suffices to look at the derivative of $P_\mathcal{N}^\lambda(S)$: \begin{displaymath} \partial_i P_\mathcal{N}^\lambda(S) = \left\{ \begin{aligned} & P_{\mathcal{N}\setminus\{i\}}^\lambda(S\setminus\{i\})\;\textrm{if}\; i\in S \\ & - P_{\mathcal{N}\setminus\{i\}}^\lambda(S)\;\textrm{if}\; i\in \mathcal{N}\setminus S \\ \end{aligned}\right. \end{displaymath} Hence: \begin{multline*} \partial_i F_\mathcal{N} = \sum_{\substack{S\subseteq\mathcal{N}\\ i\in S}} P_{\mathcal{N}\setminus\{i\}}^\lambda(S\setminus\{i\})V(S)\\ - \sum_{\substack{S\subseteq\mathcal{N}\\ i\in \mathcal{N}\setminus S}} P_{\mathcal{N}\setminus\{i\}}^\lambda(S)V(S)\\ \end{multline*} Now, using that every $S$ such that $i\in S$ can be uniquely written as $S'\cup\{i\}$, we can write: \begin{multline*} \partial_i F_\mathcal{N} = \sum_{\substack{S\subseteq\mathcal{N}\\ i\in\mathcal{N}\setminus S}} P_{\mathcal{N}\setminus\{i\}}^\lambda(S)V(S\cup\{i\})\\ - \sum_{\substack{S\subseteq\mathcal{N}\\ i\in \mathcal{N}\setminus S}} P_{\mathcal{N}\setminus\{i\}}^\lambda(S)V(S)\\ \end{multline*} Finally, by using the expression for the marginal contribution of $i$ to $S$: \begin{displaymath} \partial_i F_\mathcal{N}(\lambda) = \sum_{\substack{S\subseteq\mathcal{N}\\ i\in\mathcal{N}\setminus S}} P_{\mathcal{N}\setminus\{i\}}^\lambda(S) \log\Big(1 + \T{x_i}A(S)^{-1}x_i\Big) \end{displaymath} The computation of the derivative of $L_\mathcal{N}$ uses standard matrix calculus and gives: \begin{displaymath} \partial_i L_\mathcal{N}(\lambda) = \T{x_i}\tilde{A}(\lambda)^{-1}x_i \end{displaymath} Using the following inequalities: \begin{gather*} \forall S\subseteq\mathcal{N}\setminus\{i\},\quad P_{\mathcal{N}\setminus\{i\}}^\lambda(S)\geq P_{\mathcal{N}\setminus\{i\}}^\lambda(S\cup\{i\})\\ \forall S\subseteq\mathcal{N},\quad P_{\mathcal{N}\setminus\{i\}}^\lambda(S) \geq P_\mathcal{N}^\lambda(S)\\ \forall S\subseteq\mathcal{N},\quad A(S)^{-1} \geq A(S\cup\{i\})^{-1}\\ \end{gather*} we get: \begin{align*} \partial_i F_\mathcal{N}(\lambda) & \geq \sum_{\substack{S\subseteq\mathcal{N}\\ i\in\mathcal{N}\setminus S}} P_\mathcal{N}^\lambda(S) \log\Big(1 + \T{x_i}A(S)^{-1}x_i\Big)\\ & \geq \frac{1}{2} \sum_{\substack{S\subseteq\mathcal{N}\\ i\in\mathcal{N}\setminus S}} P_\mathcal{N}^\lambda(S) \log\Big(1 + \T{x_i}A(S)^{-1}x_i\Big)\\ &\hspace{-3.5em}+\frac{1}{2} \sum_{\substack{S\subseteq\mathcal{N}\\ i\in\mathcal{N}\setminus S}} P_\mathcal{N}^\lambda(S\cup\{i\}) \log\Big(1 + \T{x_i}A(S\cup\{i\})^{-1}x_i\Big)\\ &\geq \frac{1}{2} \sum_{S\subseteq\mathcal{N}} P_\mathcal{N}^\lambda(S) \log\Big(1 + \T{x_i}A(S)^{-1}x_i\Big)\\ \end{align*} Using that $A(S)\geq I_d$ we get that: \begin{displaymath} \T{x_i}A(S)^{-1}x_i \leq 1 \end{displaymath} Moreover: \begin{displaymath} \forall x\leq 1,\; \log(1+x)\geq x \end{displaymath} Hence: \begin{displaymath} \partial_i F_\mathcal{N}(\lambda) \geq \frac{1}{2} \T{x_i}\bigg(\sum_{S\subseteq\mathcal{N}}P_\mathcal{N}^\lambda(S)A(S)^{-1}\bigg)x_i \end{displaymath} Finally, using that the inverse is a matrix convex function over symmetric positive definite matrices: \begin{align*} \partial_i F_\mathcal{N}(\lambda) &\geq \frac{1}{2} \T{x_i}\bigg(\sum_{S\subseteq\mathcal{N}}P_\mathcal{N}^\lambda(S)A(S)\bigg)^{-1}x_i\\ & \geq \frac{1}{2} \partial_i L_\mathcal{N}(\lambda)\qed \end{align*} \end{proof} \begin{lemma}\label{lemma:relaxation} We have: \begin{displaymath} OPT(L_\mathcal{N}, B) \leq 4 OPT(V,\mathcal{N},B) + 2\max_{i\in\mathcal{N}}V(i) \end{displaymath} \end{lemma} \begin{proof} Let us consider a feasible point $\lambda^*\in[0,1]^{|\mathcal{N}|}$ such that $L_\mathcal{N}(\lambda^*) = OPT(L_\mathcal{N}, B)$. By applying lemma~\ref{lemma:relaxation-ratio} and lemma~\ref{lemma:rounding} we get a feasible point $\bar{\lambda}$ with at most one fractional component such that: \begin{equation}\label{eq:e1} L_\mathcal{N}(\lambda^*) \leq 2 F_\mathcal{N}(\bar{\lambda}) \end{equation} Let $\lambda_i$ denote the fractional component of $\bar{\lambda}$ and $S$ denote the set whose indicator vector is $\bar{\lambda} - \lambda_i e_i$. Using the fact that $F_\mathcal{N}$ is linear with respect to the $i$-th component and is a relaxation of the value function, we get: \begin{displaymath} F_\mathcal{N}(\bar{\lambda}) = V(S) +\lambda_i V(S\cup\{i\}) \end{displaymath} Using the submodularity of $V$: \begin{displaymath} F_\mathcal{N}(\bar{\lambda}) \leq 2 V(S) + V(i) \end{displaymath} Note that since $\bar{\lambda}$ is feasible, $S$ is also feasible and $V(S)\leq OPT(V,\mathcal{N}, B)$. Hence: \begin{equation}\label{eq:e2} F_\mathcal{N}(\bar{\lambda}) \leq 2 OPT(V,\mathcal{N}, B) + \max_{i\in\mathcal{N}} V(i) \end{equation} Putting \eqref{eq:e1} and \eqref{eq:e2} together gives the results. \end{proof}