diff options
| author | Thibaut Horel <thibaut.horel@gmail.com> | 2012-11-03 14:42:42 +0100 |
|---|---|---|
| committer | Thibaut Horel <thibaut.horel@gmail.com> | 2012-11-03 14:42:42 +0100 |
| commit | f64ff3ca389ec33c240bf6d872033a4bc6e4d9c6 (patch) | |
| tree | 140986849d8c0d25ae5de9f3c8429cee8fccb125 | |
| parent | 58c72935ed45da7f7dea0e28009219e5c5ba919c (diff) | |
| download | recommendation-f64ff3ca389ec33c240bf6d872033a4bc6e4d9c6.tar.gz | |
Small fixes
| -rw-r--r-- | intro.tex | 3 | ||||
| -rw-r--r-- | main.tex | 41 |
2 files changed, 15 insertions, 29 deletions
@@ -63,3 +63,6 @@ $O(\log n)$-competitive algorithm. In the bidding model $O(1)$-competitive algorithm. \stratis{What is known about the maximization of logdet in the non-strategic case? Is it NP hard? Approximation ratios?} +\thibaut{Knapsack reduces to our problem in dimension 1, hence maximizing log +det is NP-hard. The approximation ratio is at least (1-1/e) by applying the +general approximation algorithm from Sviridenko \cite{sviridenko-submodular}.} @@ -62,10 +62,10 @@ the allocation function. The way this issue has been addressed thus far \cite{chen, singer-influence}, is to introduce a third quantity: if $V(i^*)$ is larger than this quantity, the mechanism allocates to $i^*$, otherwise it allocates to the greedy set $S_G$. -To keep a bounded approximation ratio, this quantity must be provably close to -$V(S_G)$ while maintaining the monotonicity of the allocation algorithm. -Furthermore, it must be computable in polynomial time to keep an overall -polynomial complexity for the allocation algorithm. If the underlying +This quantity must be provably close to $V(S_G)$, to keep a bounded +approximation ratio, while maintaining the monotonicity of the allocation +algorithm. Furthermore, it must be computable in polynomial time to keep an +overall polynomial complexity for the allocation algorithm. If the underlying non-strategic optimization program \eqref{eq:non-strategic} can be solved in polynomial time, Chen et al. \cite{chen} prove that allocating to $i^*$ when $V(i^*) \geq C\cdot OPT(V,\mathcal{N}\setminus\{i^*\}$ (for some constant $C$) @@ -73,26 +73,25 @@ and to $S_G$ otherwise yields a 8.34 approximation mechanism. For specific problems, Chen et al. \cite{chen} for knapsack on one hand, Singer \cite{singer-influence} for maximum coverage on the other hand, instead compare $V(i^*)$ to $OPT(R_{\mathcal{N}\setminus\{i\}}, B)$, where $R_\mathcal{N}$ -denotes a fractional relaxation of the function $V$ over the set -$\mathcal{N}$. +denotes a fractional relaxation of the function $V$ over the set $\mathcal{N}$. We say that $R_\mathcal{N}$ is a fractional relaxation of $V$ over the set $\mathcal{N}$, if (a) $R_\mathcal{N}$ is a function defined on the hypercube $[0, 1]^{|\mathcal{N}|}$ and (b) for all $S\subseteq\mathcal{N}$, if $\mathbf{1}_S$ denotes the indicator vector of $S$ we have $R_\mathcal{N}(\mathbf{1}_S) = V(S)$. A relaxation which is commonly used, due -to its nice properties in the context of maximizing submodular functions, is the -\emph{multi-linear} extension: +to its well-behaved properties in the context of maximizing submodular +functions, is the \emph{multi-linear} extension: \begin{equation}\label{eq:multilinear} F_\mathcal{N}^V(\lambda) = \mathbb{E}_{S\sim P_\mathcal{N}^\lambda}\left[V(S)\right] = \sum_{S\subseteq\mathcal{N}} P_\mathcal{N}^\lambda(S) V(S) \end{equation} -where $P_\mathcal{N}^\lambda(S)$ is the probability of choosing the set $S$ if +$P_\mathcal{N}^\lambda(S)$ is the probability of choosing the set $S$ if we decide for each element in $\mathcal{N}$ to pick it with probability $\lambda_i$ or to reject it with probability $1-\lambda_i$: \begin{displaymath} - P_\mathcal{N}^\lambda = \prod_{i\in S} \lambda_i + P_\mathcal{N}^\lambda(S) = \prod_{i\in S} \lambda_i \prod_{i\in\mathcal{N}\setminus S} 1 - \lambda_i \end{displaymath} For knapsack, Chen et al. \cite{chen} directly use the multi-linear extension as @@ -113,8 +112,8 @@ function. Note that in our case, the multi-linear extension can be written: Our relaxation follows naturally by swapping the expectation and the $\log\det$ in the above formula: \begin{align}\label{eq:concave} - L_\mathcal{N}(\lambda) & \defeq \log\det\mathbb{E}_{S\sim - P_\mathcal{N}^\lambda} A(S) \\ + L_\mathcal{N}(\lambda) & \defeq \log\det\left(\mathbb{E}_{S\sim + P_\mathcal{N}^\lambda}[A(S)]\right)\notag \\ &= \log\det\left(I_d + \sum_{i\in\mathcal{N}} \lambda_i x_i\T{x_i}\right) \end{align} @@ -158,7 +157,7 @@ We can now state the main result of this section: \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 + 1 + C^* = \frac{14e-3 + \sqrt{160e^2-48e + 9}}{2(e-1)}\simeq 19.68 \end{displaymath} \end{theorem} @@ -304,22 +303,6 @@ 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}. -\begin{lemma}\label{lemma:concave} - The \emph{concave relaxation} $L_\mathcal{N}$ is concave\footnote{Hence - this relaxation is well-named! -\stratis{avoid humor}}. -\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 |
