Lecture notes for COMS 6998 - Social Network Economics
Fall 2013, Columbia University
by A. Chaintreau


Part A.2 - Manipulation

Motivation

Many social interaction with opinions do not result in a single state representing the opinion that is now believed by all nodes, and when it does, there is no guarantee that this equilibrium is what our previous models predict. There are various reasons for that, and the first one that we study is the presence of nodes who interact without respecting fully the update rules that we have defined so far. We focus on two particular anomaly:

Like in the previous naive learning we studied, the node treats information received from its neighbors as if they constitute entirely new information to be treated as in the first round, except that after doing that, the node does not update the initial value to the current estimate. Extreme forms of influence (totally assymetric for a node) or bias leads to stubborness: nodes that affect the outcome of the system without revising their estimates at anytime.

Update model with influence and bias**

We know will consider different cases depending in which transitions (averaging, influence, others) are either be triggered by a pair

or it can alternatively be triggered a node.

The case in which all nodes simultaneously perform a cost minimization will be called synchronous, the cases where a single node is chosen at random to make a transition alone is called asynchronous.

Some immediate remarks on this model:


Analysis of the model with nodal updates

We first consider results from two paper that study a Synchronous cost minimization where nodes display general bias and/or stubborness. In contrast with the iterated averaging we have studied, nodes in the limit do not necessarily come to an agreement. The first paper characterizes the evolution of the system and allows us to understand its limit in more explicit mathematical terms. The second paper characterizes the amount of disagreement, be it internal (towards initial opinion for biased nodes) or external (towards other nodes). It shows that the sum of this disagreement cost attained by the equilibrium of this system is never too far from what would be the minimum cost always present for all opinions.

Convergence and Limit expressed using Random Walks

J. Ghaderi and R. Srikant, “Opinion Dynamics in Social Networks: A Local Interaction Game with Stubborn Agents,” ACC ‘13: Proceedings of the American Control Conference, vol. cs.GT. 2013.

Summary of main contributions:

Model: The paper assumes that the network of neighbors is undirected (or equivalently, symmetric) and that it is connected (and, hence, strongly connected).

Nodes initially have an opinion xi(0) based on their private information. The nodes that keep this opinion are refered to as partially stubborn whereas the nodes that never update their value are totally stubborn, others are unbiased.

If we assume that all users aim at minimizing the following individually perceived cost

Costi=Ki(xixi(0))2+jN(i)(xixj)2,
then assuming neighbors and initial opinion is fixed, the node will follow the nodal update rule below:
xi1Ki+|N(i)|Kixi(0)+jN(i)xj

Analysis of convergence and limit:

Proof of the equilibrium attained (Lemma 3):

Studying the disagreement cost at equilibrium

D. Bindel, J. M. Kleinberg, and S. Oren, “How Bad is Forming Your Own Opinion?,” Foundations of Computer Science (FOCS), 2011 IEEE 52nd Annual Symposium on, pp. 57–66, 2011.

Summary of contributions: * Show that when costs correspond to a sum on undirected edges, the equilibrium always incurs a total disagreement cost that is never more than 9/8 of any other possible states. This result is tight on an elementary example. * Proves no general bound can be derived for the case of cost defined on directed edges (in which depending on the direction of disagreement, it is counted with a different weight), even if the degree remains small. However, directed graphs with unit weight that are Eulerian (each node has equal in and out degree) offers some appropriate bounds depending on their degree. * Shows that various graph modifications techniques to reduce disagreement in directed graphs leads to large computational complexity, and provide a few cases of single modifications that can be computed.

Model:

Let us consider a simple example first:

The main contributions of this paper is to study that gap.

The proof of the main prices of anarchy results for general undirected graphs is as follows.

i,j>iwi,j(xixj)2=i,j>iwi,jxi(xixj)i,j>iwi,jxj(xixj)=i,j>iwi,jxi(xixj)+i,j>iwi,jxj(xjxi)=i,jiwi,jxi(xixj)=ixijwi,jxi+jixi(wi,j)xj=xTLx
where L is a matrix Li,j=wi,j for ij and Li,i=jwi,j (sometimes called the Laplacian matrix, generally defined by DegreeAdjacency that has multiple properties related to the graph, in particular it is positive semidefinite matrix with only non-negative eigenvalues).

Hence, the total cost of the network can be written as Cost(x)=2xTLx+||xx(0)||2. We can rewrite

Cost(x)=2xTLx+(xx(0))T(xx(0))=xT(2L+I)x2x(0)Tx+x(0)Tx(0).
this function is a quadratic form, and since 2L+I is positive definitive (i.e., its eigenvalues are all positive), the only minimizer of this cost satisfies (2L+I)x=x(0) or, equivalently, x=(2L+I)1x(O).

Note that any Nash equilibrium x~ satisfies (I+L)x~=x(0). Indeed, we can write

((I+L)x~)i=(1+jiwi,j)x~ijix~j
which is by definition of the best response seen above, equal to xi(0). This equality implies for the same reasons as above that x~=(I+L)1x(0).

Note that one can immediately use these two properties to show that the price of anarchy is upper bounded by 2, but a better bound exists, that we now detail.

Given the form of x and x~, we have:

Cost(x~)=x~T2Lx~+||x~x(0)||2=x~T2Lx~+(x~x(0))T(x~x(0))=x(0)T((I+L)12L(I+L)1+((I+L)1I)2)x(0)=x(0)TMx(0)
where we call the (complicated) middle Matrix M.

Similarly

Cost(x)=x(0)T((I+2L)12L(I+2L)1+((I+2L)1I)2)x(0)=x(0)TNx(0)
where N is the middle matrix.

As a consequence, since if we diagonalize L in another base of vector that are orthogonal, the same base diagonalize M and N and the eigenvalues of M and N corresponds exactly to the function of the value in their respective form, hence

λMi=11+λi2λi11+λi+(11+λi1)2=2λi+λ2i(1+λi)2=λi(λi+2)(1+λi)2.
Similarly, we have
λNi=11+2λi2λi11+2λi+(11+2λi1)2=2λi+4λ2i(1+2λi)2=2λi1+2λi
Hence λMi/λNi=(λi+2)(1+2λi)2(1+λi)2)=λ2i+5/2λi+1λ2i+2λi+1, this last function is seen to have a maximum equal to 9/8 when λi=1 which concludes the proof.

Analysis of the model with Pairwise Updates

The pairwise version of the above evolution is necessarily asynchronous and hence leads to larger effects of randomness.

The main results in that pairwise model is to prove the convergence of the system to a limit that can be characterized by the evolution of an associated Markov Chain.

Two particular cases are studied depending on the topology of the networks

Case 1: no man on an island

For this particular case, one can show that the system always converge to a state in which all nodes agree. However, the value of this agreement is a random variable that depends on the system itself. The following paper establishes various bounds to study the effect of “influential” or “forceful” link.

D. Acemoglu, A. Ozdaglar, and A. ParandehGheibi, “Spread of (mis) information in social networks,” Games and Economic Behavior, vol. 70, no. 2, pp. 194–227, 2010.

When some agents disproportionally influence others (while still participating to the update of belief), how this leads to a consensus whose distribution may be influence by these agents. Under which conditions can this distortion be avoided.

The main contributions of the paper are:

The core of the proof relies on a perturbation result from Markov Chain theory that involves the so-called fundamental matrix of the Markov Chain with transition matrix C (defined as Y=k0(CkC), where C has all rows equal to π). Since the evolution of the Markov Chain is perturbed by D (with sums along rows equal to 0), we can deduce ||E[s]1ne||2||DY||2$ which yields the result.

Case 2: All nodes under influence of stubborn

The analysis of this case generally focuses on the simpler case where all nodes with an outgoing influential or forceful edges are stubborn. This case is already complex to analyze as the system enters a persistent disagreement and fluctuation states where it converges only in distribution. This essentially means that the system becomes time-invariant, a property of its ergodicity. However, just like for the nodal update, one can again show that the expected opinion of a node is related to the behavior of a random walk.

D. Acemoglu, G. Como, F. Fagnani, and A. Ozdaglar, “Opinion fluctuations and disagreement in social networks,” Mathematics of Operations Research, vol. 38, no. 1, pp. 1–27, 2013.

Continuous opinion dynamics with a fixed set of stubborn agents. The main difference with the work above is that we do not assume any more that the active links formed a strongly connected graph. In particular, some agents may be influencing other without being influenced at all. These are stubborn and will not change their positions once they have adopted it initially.

Model

Main contributions of the paper:

E. Yildiz, D. Acemoglu, A. Ozdaglar, A. Saberi, and A. Scaglione, “Binary Opinion Dynamics with Stubborn Agents,” ACM Transactions on Economics and Computation, 2013. (previously appeared as “Discrete opinion dynamics with stubborn agents,” SSRN 1744113, 2011).

Considers a special case of the model above: there are only two opinions {0,1}, and all influential transitions follow a complete trust (i.e., θi,j=1 whenever this transition occurs).

The results of this paper are

Going further: More on study of disagreement

Most of the on going works deal with the nodal update model that is conceptually simple to relate to the notion of Best Response and Nash Equilibrium.

K. Bhawalkar, S. Gollapudi, and K. Munagala, “Coevolutionary opinion formation games,” presented at the STOC ‘13: Proceedings of the 45th annual ACM symposium on Symposium on theory of computing, 2013.

Proposes more price of anarchy results for a similar model and considers a case where edges may be created/deleted by current agreement/disagreement.

F. Chierichetti, J. M. Kleinberg, and S. Oren, “On discrete preferences and coordination,” presented at the EC ‘13: Proceedings of the fourteenth ACM conference on Electronic commerce, 2013.

When opinion or choices are discrete, and the distance function comes from a preference ranking, how does a network of nodes lead to various efficiency.

Written with StackEdit.