当前位置:高等教育资讯网  >  中国高校课件下载中心  >  大学文库  >  浏览文档

复旦大学:《网络科学导论 Introduction to Network Science》教学参考文献_Modularity and community structure of networks

资源类别:文库,文档格式:PDF,文档页数:6,文件大小:604.18KB,团购合买
点击下载完整版文档(PDF)

Modularity and community structure in networks M.E. Newman* Department of Physics and Center for the Study of Complex Systems, University of Michigan, Ann Arbor, MI 48109 Edited by Brian Skyrms, University of California, Irvine, CA, and approved April 19, 2006(received for review February 26, 2006) Many networks of interest in the sciences, including social net- rks, computer networks, and metabolic and regulatory net works, are found to divide naturally into communities or modules. The problem of detecting and characterizing this community struc ture is one of the outstanding issues in the study of networked systems. One highly effective approach is the optimization of the quality function known as"modularity"over the possible divisions of a network. Here l show that the modularity can be expressed in terms of the eigenvectors of a characteristic matrix for the net- work, which I call the modularity matrix, and that this expressic leads to a spectral algorithm for community detection that returns results of demonstrably higher quality than competing methods in shorter running times. I illustrate the method with appl ns to several published network data sets clustering I partitioning I modules metabolic network I social network M any systems of scientific interest can be represented as networks, sets of nodes or vertices joined in pairs by lines or edges. Examples include the internet and the worldwide web, Fig. 1. The vertices i etworks fall natural to groups or commu- metabolic networks, food webs, neural networks, communica- nities, sets of vertices(shaded)within which there are many edges, with only tion and distribution networks, and social networks. The study of a smaller number of edges between vertices of different groups networked systems has a history stretching back several centu last decade cially in the mathematical sciences, partly as a goals of the two camps that make quite different technical result of the increasing availability of accurate large-scale data approaches desirable. A typical problem in graph partitioning is analyses of these data have revealed some une d structural h of features, such as high network transitivity(1), power-law degree cessor communication. In such an application the number of distributions(2), and the existence of repeated local motifs(3); processors is usually known in advance and at least an approx- see refs. 4-6 for review imate figure for the number of tasks that each processor ca One issue that has received a considerable amount of attention handle. Thus we know the number and size of the groups into the detection and characterization of community structure in the best division of the network regardless of whether a good groups of vertices, with only sparser connections between groups division even exists; there is little point in an algorithm or (Fig. 1). The ability to detect such groups could be of significant method that fails to divide the network in some cases. practical importance. For instance, groups within the worldwide Community structure detection, by contrast, is perhaps best web might correspond to sets of web pages on related topics( 9); thought of as a data analysis technique used to shed light on the or communities(10). Merely the finding that a network contains works, internet and web data, or biochemical networks. Com- tightly knit groups at all can convey useful information: if a munity structure methods normally assume that the network of metabolic network were divided into such groups, for instance, interest divides naturally into subgroups and the experimenter's it could provide evidence for a modular view of the network's job is to find those groups. The number and size of the groups dynamics, with different groups of nodes performing different are thus determined by the network itself and not by the functions with some degree of independence (11, 12) experimenter. Moreover, community structure methods may Past work on methods for discovering groups in networks explicitly admit the possibility that no good division of the divides into two principal lines of research, both with long network exists, an outcome that is itself considered to be of histories. The first, which goes by the name of graph partitioning, interest for the light it sheds on the topology of the network has been pursued particularly in computer science and related This article focuses on community structure detection in fields, with applications in parallel computing and integrated network data sets representing real-world systems of interest circuit design, among other areas(13, 14 ). The second, identified However, both the similarities and differences between commu- names sue Ich as block modeling, hierarchical clustering, or nity ure methods and graph partitioning will motivate mmunity structure detection, has been pursued by sociologists many developments that follo and more recently by physicists, biologists, and applied mathe maticians, with applications especially to social and biological networks (7, 15, 16 Conflict of statement: No conflic It is tempting to suggest that these two lines of research are This paper was submitted directly (Track m to the PNAS office ally addressing the same question, albeit by somewhat different *E-mail: mejn@umich.edu means. There are, however, important differences between the 02006 by The National Academy of Sciences of the USA pnas. org/cgi/doi/10.1073/pnas.060160210 PNAs|June6,2006|vol.103|no.23|8577-8582

Modularity and community structure in networks M. E. J. Newman* Department of Physics and Center for the Study of Complex Systems, University of Michigan, Ann Arbor, MI 48109 Edited by Brian Skyrms, University of California, Irvine, CA, and approved April 19, 2006 (received for review February 26, 2006) Many networks of interest in the sciences, including social net￾works, computer networks, and metabolic and regulatory net￾works, are found to divide naturally into communities or modules. The problem of detecting and characterizing this community struc￾ture is one of the outstanding issues in the study of networked systems. One highly effective approach is the optimization of the quality function known as ‘‘modularity’’ over the possible divisions of a network. Here I show that the modularity can be expressed in terms of the eigenvectors of a characteristic matrix for the net￾work, which I call the modularity matrix, and that this expression leads to a spectral algorithm for community detection that returns results of demonstrably higher quality than competing methods in shorter running times. I illustrate the method with applications to several published network data sets. clustering  partitioning  modules  metabolic network  social network Many systems of scientific interest can be represented as networks, sets of nodes or vertices joined in pairs by lines or edges. Examples include the internet and the worldwide web, metabolic networks, food webs, neural networks, communica￾tion and distribution networks, and social networks. The study of networked systems has a history stretching back several centu￾ries, but it has experienced a particular surge of interest in the last decade, especially in the mathematical sciences, partly as a result of the increasing availability of accurate large-scale data describing the topology of networks in the real world. Statistical analyses of these data have revealed some unexpected structural features, such as high network transitivity (1), power-law degree distributions (2), and the existence of repeated local motifs (3); see refs. 4–6 for reviews. One issue that has received a considerable amount of attention is the detection and characterization of community structure in networks (7, 8), meaning the appearance of densely connected groups of vertices, with only sparser connections between groups (Fig. 1). The ability to detect such groups could be of significant practical importance. For instance, groups within the worldwide web might correspond to sets of web pages on related topics (9); groups within social networks might correspond to social units or communities (10). Merely the finding that a network contains tightly knit groups at all can convey useful information: if a metabolic network were divided into such groups, for instance, it could provide evidence for a modular view of the network’s dynamics, with different groups of nodes performing different functions with some degree of independence (11, 12). Past work on methods for discovering groups in networks divides into two principal lines of research, both with long histories. The first, which goes by the name of graph partitioning, has been pursued particularly in computer science and related fields, with applications in parallel computing and integrated circuit design, among other areas (13, 14). The second, identified by names such as block modeling, hierarchical clustering, or community structure detection, has been pursued by sociologists and more recently by physicists, biologists, and applied mathe￾maticians, with applications especially to social and biological networks (7, 15, 16). It is tempting to suggest that these two lines of research are really addressing the same question, albeit by somewhat different means. There are, however, important differences between the goals of the two camps that make quite different technical approaches desirable. A typical problem in graph partitioning is the division of a set of tasks between the processors of a parallel computer so as to minimize the necessary amount of interpro￾cessor communication. In such an application the number of processors is usually known in advance and at least an approx￾imate figure for the number of tasks that each processor can handle. Thus we know the number and size of the groups into which the network is to be split. Also, the goal is usually to find the best division of the network regardless of whether a good division even exists; there is little point in an algorithm or method that fails to divide the network in some cases. Community structure detection, by contrast, is perhaps best thought of as a data analysis technique used to shed light on the structure of large-scale network data sets, such as social net￾works, internet and web data, or biochemical networks. Com￾munity structure methods normally assume that the network of interest divides naturally into subgroups and the experimenter’s job is to find those groups. The number and size of the groups are thus determined by the network itself and not by the experimenter. Moreover, community structure methods may explicitly admit the possibility that no good division of the network exists, an outcome that is itself considered to be of interest for the light it sheds on the topology of the network. This article focuses on community structure detection in network data sets representing real-world systems of interest. However, both the similarities and differences between commu￾nity structure methods and graph partitioning will motivate many of the developments that follow. Conflict of interest statement: No conflicts declared. This paper was submitted directly (Track II) to the PNAS office. *E-mail: mejn@umich.edu. © 2006 by The National Academy of Sciences of the USA Fig. 1. The vertices in many networks fall naturally into groups or commu￾nities, sets of vertices (shaded) within which there are many edges, with only a smaller number of edges between vertices of different groups. www.pnas.orgcgidoi10.1073pnas.0601602103 PNAS  June 6, 2006  vol. 103  no. 23  8577– 8582 APPLIED MATHEMATICS

The Method of Optimal Modularity reformulation of the modularity in terms of the spectral prop- Suppose then that we are given, or discover, the structure of erties of the network of interest some network and that we want to determine whether there Suppose our network contains n vertices. For a particular exists any natural division of its vertices into nonoverlapping division of the network into two groups let s;=1 if vertex i groups or communities, where these communities may be of any belongs to group l and s=-I if it belongs to group 2. And let the number of edges between vertice es i and j be Ai, which will Let us approach this question in stages and focus initially on normally be 0 or 1, although larger values are possible in the problem of whether any good division of the network exists networks where multiple edges are allowed. (The quantities A, into just two communities. Perhaps the most obvious way to are the elements of the so-called adjacency matrix. )At the same tackle this problem is to look for divisions of the vertices into two time, the expected number of edges between vertices i and jif the groups. This"minimum cut"approach is the approach most degrees of the vertices and m=2>, k is the total number of edges often adopted in the graph-partitioning literature. Howeve discussed above, the community structure problem differs cru- Ay-kki /2m over all pairs of vertices i,j that fall in the same nities are not normally known in advance. If community sizes are erving that the quantity ;(sp;+ 1)is 1 ifi andj are in the unconstrained then we are, for instance, at liberty to select the and 0 otherwise, we can then express the modularity as trivial division of the network that puts all of the vertices in one of our two groups and none in the other, which guarantees we Qm∑(4-2)+D-m∑(4 will have zero intergroup edges. This division is, in a sense optimal, but clearly it does not tell us anything of any worth. We can,if we want, artificially forbid this solution, but then a division [1 that puts just one vertex in one group and the rest in the other where the second equality follows from the observation that will often be optimal, and so forth The problem is that simply counting edges is not a good way to ki= Ey Ai. The leading factor of 1/4m is merely conv al: it is included for compatibility with the previous uantify the intuitive concept of community structure. A good definition of modularity(17) division of a network into communities is not merely one in which Eq. 1 can conveniently be written in matrix form there are few edges between communities: it is one in which there are fewer than expected edges between communities. If the number of edges between two groups is only what one would expect on the Q [2] asis of random chance then few thoughtful observers would claim this constitutes evidence of meaningful community structure. On where s is the column vector whose elements are the S; and we the other hand, if the number of edges between groups is signifi- have defined a real symmetric matrix B with elements cantly less than we expect by chance, or equivalent if the number within groups is significantly more, then it is reasonable to conclude kki 3 that something interesting is going on. This idea, that true community structure in a network corre sponds to a statistically surprising arrangement of edges, can be which we call the modularity matrix. Much of our attention in quantified by using the measure known as modularity (17). The this article will be devoted to the properties of this matrix. For dges falling within groups minus the expected number in an columns sum to zero, so that it always has an eigenvector equivalent network with edges placed at random. (A precise (1, 1, I,.. )with eigenvalue zero. This observation is reminiscent mathematical formulation is given below.) The modularity can be either positive or negative, with basis for one of the best-known methods of graph partitioning sitive values indicating the possible presence of community spectral partitioning(21, 22), and has the same property. And structure. Thus, one can search for community structure pre indeed, the methods presented here have many similarities to cisely by looking for the divisions of a network that have positive, as wey partitioning, although there are some crucial differences IV The evidence so far suggests that this approach, of looking for Given Eq. 2, we proceed by writing s as a linear combination of the normalized eigenvectors u; of B so that s= 2i-1 a;u; with divisions with high modularity, is a very effective way to tackle ai=us. Then we find the problem. For instance, Guimera and Amaral(12) and later Danon et al. ( 8)optimized modularity over possible partitions of computer-generated test networks by using simulated annealing 1 In direct comparisons using standard measures, Danon et al Q=n2aB2a=切m2(吗s)3B.4 found that this method outperformed all other methods for community detection of which they were aware, in most cases by where B, is the eigenvalue of B corresponding to eigenvector u, an impressive margin. On the basis of such results we consider Let us assume that the eigenvalues are labeled in maximization of the modularity to be perhaps the definitive order, B1==.2, We want to maximize the modularity current method of community detection, being at the same time by choosing an appropriate division of the network, or equiva based on sensible statistical principles and highly effective in lently by choosing the value of the index vector s. This means practice. Unfortunately, optimization by simulated annealing is choosing s so as to concentrate as much weight as possible in not a workable approach for the large network problems facing terms of the sum in Eq. 4 involving the larg most todays scientists, because it demands too much computational eigenvalues. If there were no other constraints on our choice of effort. A number of alternative heuristic methods have been s(apart from normalization), this would be an easy task: we estigated, such as greedy algorithms(18)and extremal opti- would simply chose s proportional to the eigenvector un. This mization(19). Here we take a different approach based on a places all of the weight in the term involving the largest 8578iwww.pnas.org/cgi/doi/10.1073/pnas.060160210

The Method of Optimal Modularity Suppose then that we are given, or discover, the structure of some network and that we want to determine whether there exists any natural division of its vertices into nonoverlapping groups or communities, where these communities may be of any size. Let us approach this question in stages and focus initially on the problem of whether any good division of the network exists into just two communities. Perhaps the most obvious way to tackle this problem is to look for divisions of the vertices into two groups so as to minimize the number of edges running between the groups. This ‘‘minimum cut’’ approach is the approach most often adopted in the graph-partitioning literature. However, as discussed above, the community structure problem differs cru￾cially from graph partitioning in that the sizes of the commu￾nities are not normally known in advance. If community sizes are unconstrained then we are, for instance, at liberty to select the trivial division of the network that puts all of the vertices in one of our two groups and none in the other, which guarantees we will have zero intergroup edges. This division is, in a sense, optimal, but clearly it does not tell us anything of any worth. We can, if we want, artificially forbid this solution, but then a division that puts just one vertex in one group and the rest in the other will often be optimal, and so forth. The problem is that simply counting edges is not a good way to quantify the intuitive concept of community structure. A good division of a network into communities is not merely one in which there are few edges between communities; it is one in which there are fewer than expected edges between communities. If the number of edges between two groups is only what one would expect on the basis of random chance, then few thoughtful observers would claim this constitutes evidence of meaningful community structure. On the other hand, if the number of edges between groups is signifi￾cantly less than we expect by chance, or equivalent if the number within groups is significantly more, then it is reasonable to conclude that something interesting is going on. This idea, that true community structure in a network corre￾sponds to a statistically surprising arrangement of edges, can be quantified by using the measure known as modularity (17). The modularity is, up to a multiplicative constant, the number of edges falling within groups minus the expected number in an equivalent network with edges placed at random. (A precise mathematical formulation is given below.) The modularity can be either positive or negative, with positive values indicating the possible presence of community structure. Thus, one can search for community structure pre￾cisely by looking for the divisions of a network that have positive, and preferably large, values of the modularity (18). The evidence so far suggests that this approach, of looking for divisions with high modularity, is a very effective way to tackle the problem. For instance, Guimera` and Amaral (12) and later Danon et al. (8) optimized modularity over possible partitions of computer-generated test networks by using simulated annealing. In direct comparisons using standard measures, Danon et al. found that this method outperformed all other methods for community detection of which they were aware, in most cases by an impressive margin. On the basis of such results we consider maximization of the modularity to be perhaps the definitive current method of community detection, being at the same time based on sensible statistical principles and highly effective in practice. Unfortunately, optimization by simulated annealing is not a workable approach for the large network problems facing today’s scientists, because it demands too much computational effort. A number of alternative heuristic methods have been investigated, such as greedy algorithms (18) and extremal opti￾mization (19). Here we take a different approach based on a reformulation of the modularity in terms of the spectral prop￾erties of the network of interest. Suppose our network contains n vertices. For a particular division of the network into two groups let si  1 if vertex i belongs to group 1 and si  1 if it belongs to group 2. And let the number of edges between vertices i and j be Aij, which will normally be 0 or 1, although larger values are possible in networks where multiple edges are allowed. (The quantities Aij are the elements of the so-called adjacency matrix.) At the same time, the expected number of edges between vertices i and j if edges are placed at random is kikj2m, where ki and kj are the degrees of the vertices and m  1 2 i ki is the total number of edges in the network. Thus the modularity Q is given by the sum of Aij kikj2m over all pairs of vertices i,j that fall in the same group. Observing that the quantity 1 2 (sisj  1) is 1 if i and j are in the same group and 0 otherwise, we can then express the modularity as Q  1 4m ij Aij kikj 2m sisj 1  1 4m ij Aij kikj 2msisj, [1] where the second equality follows from the observation that 2m  i ki  ij Aij. The leading factor of 14m is merely conventional: it is included for compatibility with the previous definition of modularity (17). Eq. 1 can conveniently be written in matrix form as Q  1 4m sTBs, [2] where s is the column vector whose elements are the si and we have defined a real symmetric matrix B with elements Bij  Aij kikj 2m , [3] which we call the modularity matrix. Much of our attention in this article will be devoted to the properties of this matrix. For the moment, note that the elements of each of its rows and columns sum to zero, so that it always has an eigenvector (1,1,1, . . .) with eigenvalue zero. This observation is reminiscent of the matrix known as the graph Laplacian (20), which is the basis for one of the best-known methods of graph partitioning, spectral partitioning (21, 22), and has the same property. And indeed, the methods presented here have many similarities to spectral partitioning, although there are some crucial differences as well, as we will see. Given Eq. 2, we proceed by writing s as a linear combination of the normalized eigenvectors ui of B so that s  i1 n aiui with ai  ui T s. Then we find Q  1 4m i aiui T B j ajuj  1 4m i1 n ui T s 2 i, [4] where i is the eigenvalue of B corresponding to eigenvector ui. Let us assume that the eigenvalues are labeled in decreasing order, 1  2    n. We want to maximize the modularity by choosing an appropriate division of the network, or equiva￾lently by choosing the value of the index vector s. This means choosing s so as to concentrate as much weight as possible in terms of the sum in Eq. 4 involving the largest (most positive) eigenvalues. If there were no other constraints on our choice of s (apart from normalization), this would be an easy task: we would simply chose s proportional to the eigenvector u1. This places all of the weight in the term involving the largest 8578  www.pnas.orgcgidoi10.1073pnas.0601602103 Newman

eigenvalue Bl, the other terms being automatically zero, because the eigenvectors are orthogonal Unfortunately, there is another constraint on the problem im- posed by the restriction of the elements of s to the values +l, which neans s cannot normally be chosen parallel to un. Let us do our best. however, and make it as close to parallel as possible, which is equivalent to maximizing the dot product uis. It is straightforward to see that the maximum is achieved by setting Si =+1 if the rresponding element of un is positive and si=-l otherwise. In other words, all vertices whose corresponding elements are positive go in one group and all of the rest in the other. This then gives us r the algorithm for dividing the network: we compute the leading eigenvector of the modularity matrix and divide the vertices into two groups according to the signs of the elements in this vector. We immediately notice some satisfying features of this lethod First, as has been made clear, it works even though the sizes of the communities are not specified. Unlike conventional Fig. 2. Application of the eigenvector ethod to the karate club 23. Shapes of vertices group edges, there is no need to constrain the group sizes or sponding individuals in the two known factio membership of the corre. artificially forbid the trivial solution with all vertices in a single dotted line indicates the split found by the algorithm, which matches the group. There is an eigenvector(1, 1, l,. )corresponding to such membership The shades of the vertices indicate the strength of their of the corresponding elements of the a trivial solution, but its eigenvalue is zero. All other eigenvectors eigenvector are orthogonal to this one and hence must possess both positive and negative elements. Thus, as long as there is any positive same gr The vertices in Fig. 2 are shaded according to the It is, however, possible for there to be no positive eigenvalues of elements in the leading eigenvector of the modularity the modularity matrix. In this case the leading eigenvector is the these values seem also to accord well with known soc vector(1, 1, 1,.. corresponding to all vertices in a single group within the club. In particular, the three vertices with gether. But this is precisely the correct result: the algorithm is in weights, either positive or negative(black and white vertices in this case telling us that there is no division of the network that Fig. 2), correspond to the known ringleaders of the two factions results in positive modularity, as can immediately be seen from Eq 4, because all terms in the sum will be zero or negative. The Dividing Networks into More than Two Communities nodularity of the undivided network is zero, which is the best that In the preceding section a simple matrix-based method for algorithm has the ability not only to divide networks effectively, but Many networks, however, contain more than two communities, also to refuse to divide them when no good division exists. The so we would like to extend the method to find good divisions of network is indivisible if the modularity matrix has no positive networks into larger numbers ot parts. The standard approach to eigenvalues. This idea will play a crucial role in later developments The algorithm as described makes use only of the signs of the two: we use the algorithm of the previous section first to divide elements of the leading eigenvector, but the magnitudes conver the network into two parts, then divide those parts, and so fort formation, too. Vertices corresponding to elements of large doing this it is crucial to note that it is not correct, after first agnitude make large contributions to the modularity. eg. 4 dividing a network in two, to simply delete the edges falling and conversely for small ones. Alternatively, if we take the between the two parts and then apply the algorithm again to each optimal division of a network into two groups and move a vertex subgraph. This is because the degrees appearing in the defini- gives an indication of how much the magent for that vertex tion, Eq 1, of the modularity will change if edges are deleted, and from one group to the other, the vector eler ill decrease: any subsequent maximization of modularity would thus maxi- vertices corresponding to elements of large magnitude cannot be mize the wrong quantity. Instead, the correct approach is to write moved without incurring a large modularity penalty, whereas the additional contribution 4@ to the modularity upon furthe those corresponding to smaller elements can be moved at dividing a group g of size ng in two as relatively little cost. Thus, the elements of the leading eigenvec tor measure how firmly each vertex belongs to its assigned community, those with large vector elements being strong ∑B4s+1)-∑Bn central members of their communities. whereas those with j∈g maller elements are more ambivalent As an example of the operation of this algorithm, Fig. 2 show 1|「∑B-∑2 e result of its application to a famous network from the social ij∈g f,∈g science literature, which has become something of a standard test for community detection algorithms. The network is the "karate club" network of Zachary(23), which shows the pattern of friendships between the members of a karate club at an American university in the 1970s. This example is of particular terest because, shortly after the observation and construction of the network, the club in question split in two as a result of an 4m"Bss, ernal dispute. applying our eigenvector-based algorithm to network, we find the division indicated by the dotted line in where &y is the Kronecker 8-symbol, we have made use of si Fig 2, which coincides exactly with the known division of the club 1, and B)is the ng X ng matrix with elements indexed by the in real life, indicated by the shapes of the vertices. labels i, of vertices within group g and having values Newman PNAs|June6,2006|vo.103|no.23|8579

eigenvalue 1, the other terms being automatically zero, because the eigenvectors are orthogonal. Unfortunately, there is another constraint on the problem im￾posed by the restriction of the elements of s to the values 1, which means s cannot normally be chosen parallel to u1. Let us do our best, however, and make it as close to parallel as possible, which is equivalent to maximizing the dot product u1 T s. It is straightforward to see that the maximum is achieved by setting si  1 if the corresponding element of u1 is positive and si  1 otherwise. In other words, all vertices whose corresponding elements are positive go in one group and all of the rest in the other. This then gives us the algorithm for dividing the network: we compute the leading eigenvector of the modularity matrix and divide the vertices into two groups according to the signs of the elements in this vector. We immediately notice some satisfying features of this method. First, as has been made clear, it works even though the sizes of the communities are not specified. Unlike conventional partitioning methods that minimize the number of between￾group edges, there is no need to constrain the group sizes or artificially forbid the trivial solution with all vertices in a single group. There is an eigenvector (1,1,1, . . .) corresponding to such a trivial solution, but its eigenvalue is zero. All other eigenvectors are orthogonal to this one and hence must possess both positive and negative elements. Thus, as long as there is any positive eigenvalue this method will not put all vertices in the same group. It is, however, possible for there to be no positive eigenvalues of the modularity matrix. In this case the leading eigenvector is the vector (1,1,1, . . .) corresponding to all vertices in a single group together. But this is precisely the correct result: the algorithm is in this case telling us that there is no division of the network that results in positive modularity, as can immediately be seen from Eq. 4, because all terms in the sum will be zero or negative. The modularity of the undivided network is zero, which is the best that can be achieved. This is an important feature of the algorithm. The algorithm has the ability not only to divide networks effectively, but also to refuse to divide them when no good division exists. The networks in this latter case will be called indivisible. That is, a network is indivisible if the modularity matrix has no positive eigenvalues. This idea will play a crucial role in later developments. The algorithm as described makes use only of the signs of the elements of the leading eigenvector, but the magnitudes convey information, too. Vertices corresponding to elements of large magnitude make large contributions to the modularity, Eq. 4, and conversely for small ones. Alternatively, if we take the optimal division of a network into two groups and move a vertex from one group to the other, the vector element for that vertex gives an indication of how much the modularity will decrease: vertices corresponding to elements of large magnitude cannot be moved without incurring a large modularity penalty, whereas those corresponding to smaller elements can be moved at relatively little cost. Thus, the elements of the leading eigenvec￾tor measure how firmly each vertex belongs to its assigned community, those with large vector elements being strong central members of their communities, whereas those with smaller elements are more ambivalent. As an example of the operation of this algorithm, Fig. 2 shows the result of its application to a famous network from the social science literature, which has become something of a standard test for community detection algorithms. The network is the ‘‘karate club’’ network of Zachary (23), which shows the pattern of friendships between the members of a karate club at an American university in the 1970s. This example is of particular interest because, shortly after the observation and construction of the network, the club in question split in two as a result of an internal dispute. Applying our eigenvector-based algorithm to the network, we find the division indicated by the dotted line in Fig. 2, which coincides exactly with the known division of the club in real life, indicated by the shapes of the vertices. The vertices in Fig. 2 are shaded according to the values of the elements in the leading eigenvector of the modularity matrix, and these values seem also to accord well with known social structure within the club. In particular, the three vertices with the heaviest weights, either positive or negative (black and white vertices in Fig. 2), correspond to the known ringleaders of the two factions. Dividing Networks into More than Two Communities In the preceding section a simple matrix-based method for finding a good division of a network into two parts is described. Many networks, however, contain more than two communities, so we would like to extend the method to find good divisions of networks into larger numbers of parts. The standard approach to this problem, and the one adopted here, is repeated division into two: we use the algorithm of the previous section first to divide the network into two parts, then divide those parts, and so forth. In doing this it is crucial to note that it is not correct, after first dividing a network in two, to simply delete the edges falling between the two parts and then apply the algorithm again to each subgraph. This is because the degrees appearing in the defini￾tion, Eq. 1, of the modularity will change if edges are deleted, and any subsequent maximization of modularity would thus maxi￾mize the wrong quantity. Instead, the correct approach is to write the additional contribution Q to the modularity upon further dividing a group g of size ng in two as Q  1 2m  1 2 i, jg Bijsisj 1 i, jg Bij  1 4m  i, jg Bijsisj i, jg Bij  1 4m i, jg Bij ij kg Biksisj  1 4m sTBg s, [5] where ij is the Kronecker -symbol, we have made use of si 2  1, and B(g) is the ng ng matrix with elements indexed by the labels i,j of vertices within group g and having values Fig. 2. Application of the eigenvector-based method to the karate club network of ref. 23. Shapes of vertices indicate the membership of the corre￾sponding individuals in the two known factions of the network, and the dotted line indicates the split found by the algorithm, which matches the factions exactly. The shades of the vertices indicate the strength of their membership, as measured by the value of the corresponding elements of the eigenvector. Newman PNAS  June 6, 2006  vol. 103  no. 23  8579 APPLIED MATHEMATICS

B)=Bi-8g2 Bik. found by the algorithm described here and three other sions [61 Table 1. Comparison of modularities for the network din k∈g previously published methods as described in the text, for six Because Eq 5 has the same form as Eq 2 we can now apply the networks of varying sizes spectral approach to this generalized modularity matrix, just as Modularity Q before, to maximize A@. Note that the rows and columns of B g) still sum to zero and that AQ is correctly zero if group g is Network Size n This article undivided. Note also that for a complete network Eq 6 reduces Karate 0.419 to the previous definition of the modularity matrix, Eq. 3, Jazz musicians 0.439 0.445 0.442 because k Bi is zero in that case. In repeatedly subdividing the network, an important question E-mail 0.574 we need to address is at what point to halt the subdivision Key signing 10,680081607330.846 ess. a nice feature of this method is that it provides a clear Physicists 0.6680.679 0.723 answer to this question: if there exists no division of a subgraph that will increase the modularity of the network, or equivalent order, the karate club network of Zachary(23), the that gives a positive value for A@, then there is nothing to (27), a metabolic network for the nematode Caenorhabditis elegans(28), a gained by dividing the subgraph and it should be left alone; it is network of e-mail contacts at a university(29), a trust network of mutual indivisible in the sense of the previous section. This ha signing of cryptography keys (30), when there are no positive eigenvalues to the matrix B), and working in condensed matter physics(31). No modularity figure is given for thus the leading eigenvalue provides a simple check for the the last network with the GN algorithm because the slow o(n)operation of termination of the subdivision process: if the leading eigenvalue the algorithm prevents its application to such large systems. GN, Glrvan and is zero, which is the smallest value it can take, then the subgraph Newman (10): CNM, Clauset et al. (26): DA, Duch and Arenas (19) is indivisible Note, however, that although the absence of positive eigen- algorithm(25), and indeed the Kernighan-Lin algorithm pro- alues is a sufficient condition for indivisibility, it is not necessary one. In particular, if there are only small positive ided the inspiration for the method. Despite its simplicity, we find that this method works mod negative B, may outweigh those for positive. It is straightforward ately well. It is not competitive with the best previous methods nodularity contribution sibility, however; we simply calculate the but it gives respectable modularity values in the trial applications d split directly and we have made, However, the method really comes into its own confirm that it is greater than zero Thus the algorithm is as follows. We construct the modularity duced earlier. It a common approach in standard graph eigenvalue and the corresponding eigenvector. We divide the graph Laplacian o s to use spectral partitioning based on the matrix, Eq 3, for the network and find its leading(most positive) partitioning probler an initial broad division of a network network into two parts according to the signs of the elements of Kernighan-Lin algorithm. For community structure problems sing the generalized modularity matrix, Eq 6. If at any stage we spectral approach based on the leading eigenvector of the to the total modularity, we leave the corresponding subgraph that the communities should take and this general form can then undivided.When the entire network has been decomposed into be fine-tuned by the vertex moving method to reach the best indivisible subgraphs in this way, the algorithm ends One immediate corollary of this approach is that all "com- subdivide the network until every remaining subgraph is indi- munities"in the network are, by definition, indivisible sub- visible, and no further improvement in the modularity is possible graphs. A number of authors have in the past proposed formal (We note in passing that in principle the fine-tuning method definitions of what a community is(9, 16, 24). The present could also be used to refine results from other modularity method provides an alternative first-principles definition of a maximization algorithms, such as the extremal optimization ble sub rithm of ref. 19) Further Techniques for Modularity Maximization Typically, the fine-tuning stages of the algorithm add only a few percent to the final value of the modularity. For the karate In this section I describe briefly another method for dividing club network of Fig. 2, for instance, the spectral method on its networks in two by modularity optimization, which is entirely own finds a division of the network with modularity 0=0.39 ifferent from the spectral method. Although not of special which improves to Q=0.419 upon fine-tuning. Nonetheless, an interest on its own. this second method is. as will be show provement of this magnitude is enough, as we will see, to make shortly, very effective when combined with the spectral method the difference between a method that is merely good and one Suppose we are given some initial division of our vertices into that is exceptional. two groups. We then find among the vertices the one that, when moved to the other group, will give the biggest increase in the Example Applications modularity of the complete network, or the smallest decrease if In practice, the algorithm developed here gives excellent results no increase is possible. We make such moves repeatedly, with the For a quantitative comparison between this algorithm and others constraint that each vertex is moved only once. When all n we follow Duch and Arenas(19) and compare values of the vertices have been moved, we search the set of intermediate modularity for a variety of networks drawn from the literature states occupied by the network during the operation of the Results are shown in Table 1 for six different networks, the exact algorithm to find the state that has the greatest modularity. same six used by Duch and Arenas. We compare modularity Starting again from this state, we repeat the entire process figures against three previously published algorithms: the be- teratively until no further improvement in the modularity tweenness-based algorithm of Girvan and Newman(10), which results. Those familiar with the literature on graph partitioning is widely used and has been incorporated into some of the more may find this algorithm reminiscent of the Kernighan-Lin popular network analysis programs; the fast algorithm of Clauset 8580 pnas org/cgi/doi/10.1073/pnas. 0601602103

Bij g  Bij ij kg Bik. [6] Because Eq. 5 has the same form as Eq. 2 we can now apply the spectral approach to this generalized modularity matrix, just as before, to maximize Q. Note that the rows and columns of B(g) still sum to zero and that Q is correctly zero if group g is undivided. Note also that for a complete network Eq. 6 reduces to the previous definition of the modularity matrix, Eq. 3, because k Bik is zero in that case. In repeatedly subdividing the network, an important question we need to address is at what point to halt the subdivision process. A nice feature of this method is that it provides a clear answer to this question: if there exists no division of a subgraph that will increase the modularity of the network, or equivalently that gives a positive value for Q, then there is nothing to be gained by dividing the subgraph and it should be left alone; it is indivisible in the sense of the previous section. This happens when there are no positive eigenvalues to the matrix B(g) , and thus the leading eigenvalue provides a simple check for the termination of the subdivision process: if the leading eigenvalue is zero, which is the smallest value it can take, then the subgraph is indivisible. Note, however, that although the absence of positive eigen￾values is a sufficient condition for indivisibility, it is not a necessary one. In particular, if there are only small positive eigenvalues and large negative ones, the terms in Eq. 4 for negative i may outweigh those for positive. It is straightforward to guard against this possibility, however; we simply calculate the modularity contribution Q for each proposed split directly and confirm that it is greater than zero. Thus the algorithm is as follows. We construct the modularity matrix, Eq. 3, for the network and find its leading (most positive) eigenvalue and the corresponding eigenvector. We divide the network into two parts according to the signs of the elements of this vector, and then repeat the process for each of the parts, using the generalized modularity matrix, Eq. 6. If at any stage we find that a proposed split makes a zero or negative contribution to the total modularity, we leave the corresponding subgraph undivided. When the entire network has been decomposed into indivisible subgraphs in this way, the algorithm ends. One immediate corollary of this approach is that all ‘‘com￾munities’’ in the network are, by definition, indivisible sub￾graphs. A number of authors have in the past proposed formal definitions of what a community is (9, 16, 24). The present method provides an alternative first-principles definition of a community as an indivisible subgraph. Further Techniques for Modularity Maximization In this section I describe briefly another method for dividing networks in two by modularity optimization, which is entirely different from the spectral method. Although not of special interest on its own, this second method is, as will be shown shortly, very effective when combined with the spectral method. Suppose we are given some initial division of our vertices into two groups. We then find among the vertices the one that, when moved to the other group, will give the biggest increase in the modularity of the complete network, or the smallest decrease if no increase is possible. We make such moves repeatedly, with the constraint that each vertex is moved only once. When all n vertices have been moved, we search the set of intermediate states occupied by the network during the operation of the algorithm to find the state that has the greatest modularity. Starting again from this state, we repeat the entire process iteratively until no further improvement in the modularity results. Those familiar with the literature on graph partitioning may find this algorithm reminiscent of the Kernighan–Lin algorithm (25), and indeed the Kernighan–Lin algorithm pro￾vided the inspiration for the method. Despite its simplicity, we find that this method works moder￾ately well. It is not competitive with the best previous methods, but it gives respectable modularity values in the trial applications we have made. However, the method really comes into its own when it is used in combination with the spectral method intro￾duced earlier. It is a common approach in standard graph partitioning problems to use spectral partitioning based on the graph Laplacian to give an initial broad division of a network into two parts, and then refine that division by using the Kernighan–Lin algorithm. For community structure problems we find that the equivalent joint strategy works very well. The spectral approach based on the leading eigenvector of the modularity matrix gives an excellent guide to the general form that the communities should take and this general form can then be fine-tuned by the vertex moving method to reach the best possible modularity value. The whole procedure is repeated to subdivide the network until every remaining subgraph is indi￾visible, and no further improvement in the modularity is possible. (We note in passing that in principle the fine-tuning method could also be used to refine results from other modularity maximization algorithms, such as the extremal optimization algorithm of ref. 19.) Typically, the fine-tuning stages of the algorithm add only a few percent to the final value of the modularity. For the karate club network of Fig. 2, for instance, the spectral method on its own finds a division of the network with modularity Q  0.393, which improves to Q  0.419 upon fine-tuning. Nonetheless, an improvement of this magnitude is enough, as we will see, to make the difference between a method that is merely good and one that is exceptional. Example Applications In practice, the algorithm developed here gives excellent results. For a quantitative comparison between this algorithm and others we follow Duch and Arenas (19) and compare values of the modularity for a variety of networks drawn from the literature. Results are shown in Table 1 for six different networks, the exact same six used by Duch and Arenas. We compare modularity figures against three previously published algorithms: the be￾tweenness-based algorithm of Girvan and Newman (10), which is widely used and has been incorporated into some of the more popular network analysis programs; the fast algorithm of Clauset Table 1. Comparison of modularities for the network divisions found by the algorithm described here and three other previously published methods as described in the text, for six networks of varying sizes Network Size n Modularity Q GN CNM DA This article Karate 34 0.401 0.381 0.419 0.419 Jazz musicians 198 0.405 0.439 0.445 0.442 Metabolic 453 0.403 0.402 0.434 0.435 E-mail 1,133 0.532 0.494 0.574 0.572 Key signing 10,680 0.816 0.733 0.846 0.855 Physicists 27,519 — 0.668 0.679 0.723 The networks are, in order, the karate club network of Zachary (23), the network of collaborations between early jazz musicians of Gleiser and Danon (27), a metabolic network for the nematode Caenorhabditis elegans (28), a network of e-mail contacts at a university (29), a trust network of mutual signing of cryptography keys (30), and a coauthorship network of scientists working in condensed matter physics (31). No modularity figure is given for the last network with the GN algorithm because the slow O(n3) operation of the algorithm prevents its application to such large systems. GN, Glrvan and Newman (10); CNM, Clauset et al. (26); DA, Duch and Arenas (19). 8580  www.pnas.orgcgidoi10.1073pnas.0601602103 Newman

that align closely with political views, a finding that encourages us to believe that the algorithm is capable of extracting mean- ingful results from raw network data. It is particularly interesting to note that the centrist books belong to their own communities and are not, in most cases, merely lumped in with the liberals or conservatives; this finding may indicate that political moderates form their own purchasing community. For the second example, we consider a network of political commentary web sites, also called"weblogs"or"blogs,com- piled from online directories by Adamic and Glance, who also · assigned a political alignment, conservative or liberal, to each blog based on content. The 1, 225 vertices in the network studied here correspond to the 1, 225 blogs in the largest component of Adamic and Glances network, and undirected edges connect oks and edges join books frequently purchased by the same readers. vertices if either of the corresponding blogs contained a hyper ashed lines divide the four communities found by the algorithm, and shapes link to the other on its front page. On feeding this network present the political alignment of the books(circles are liberal, squares are through the algorithm we discover that the network divides cleanly into conservative and liberal communities and, remark ably, the optimal modularity of @=0.426 is found for a division into just two communities. One cor nity has 638 vertices of et al.(26), which optimizes modularity by using a greedy algo- which 620 (97%)represent conservative blogs. The other has 587 rithm;and the extremal optimization algorithm of Duch and vertices of which 548(93%)represent liberal blogs. The algo- method, by standard measures(8), if one discounts methods ithm found no division of either of these two groups that gives practical for large networks, such as exhaustive enumeration ny positive contribution to the modularity; as near as the algorithm is able to tell, these groups are"indivisible"in the of all partitions or simulated annealing. Table 1 reveals some sense defined here. This behavior is unique in my expenen 42 interesting patterns. The algorithm clearly outperforms the methods of Girvan and Newman and Clauset et al. for all of the among networks of this size and is perhaps a testament not only networks in the task of optimizing the modularity. The extremal o the widely noted polarization of the current political land- optimization method, on the other hand, is more competitive. two fact in the United States but also to the strong cohesion of the For the smaller networks, up to 1, 000 vertices, there is essen- ially no difference in performance between the method of this Implementation article and extremal optimization the modularity values for the This algorithm is fast as well as accurate. The most time divisions found by the two algorithms differ by no more than a consuming part of the algorithm is the evaluation of the leadin few parts in 1,000 for any given network. For the larger networks, eigenvector of the modularity matrix. The fastest method for however, the spectral algorithm does better than extremal o timization, and furthermore the gap widens as network size inding this eigenvector is the simple power method, the re- Increases, to a maximum modularity difference of 6% for the peated multiplication of the matrix into a trial vector. Although rgest network studied. For the very large networks that have It appears at first glance that matrix multiplications will be slow been of particular interest in the last few years, therefore it taking o(n2)operations each because the modularity matrix appears that the spectral method for detecting community dense, we can in fact perform them much faster by exploiting the structure may be the most effective of the methods considered particular structure of the matrix. Writing B=A-kkT/2m y matrix and k is the vector who The modularity values given in Table I provide a useful elements are the degrees of the vertices, the product of B and an quantitative measure of the success of the algorithm when arbitrary vector x can be written lied to real-world problems. It is worthwhile, he to confirm that it returns sensible divisions of networks in k(k-x Bx= Ax 7 ctice. I have given one example demonstrating such a divisi Fig. 2. I have also checked the method against many or the The first term is a standard sparse matrix multiplication taking o more examples, both involving network representations of time O(m +n). The inner product kx takes time O(n)to American politics valuate and hence the second term can be evaluated in total The first example is a network of books about politic time O(n)also. Thus the complete multiplication can be per compiled by V Krebs(personal communication ). In this net- formed in O(m +n)time. Typically O(n)such multiplications work the vertices represent 105 recent books on American olitics bought from the on-line bookseller Amazon. com, and time of o[(m +n)n]overall. Often we are concerned with sparse edges join pairs of books that are frequently purchased by the graphs with m o n, in which case the running time becomes same buyer. Books were divided(by me)according to their o( ) It is a simple matter to extend this procedure to find the stated or apparent political alignment, liberal or conservative, leading eigenvector of the generalized modularity matrix, Eq. 6, except for a small number of books that were explicitly bipartisan Although I will not go through the details here, it is straight forward to show that the fine-tuning stage of the algorithm car algorithm. The algorithm finds four communities of vertices, also be completed in o(m+ n)n] time, so that the combined denoted by the dotted lines in Fig 3, with a modularity of 0.526. running time for a single split of a graph or subgraph scales as As can be seen. one of these communities consists almost entirely O[(m n)n), or O(n)on a sparse graph of liberal books and one almost entirely of conservative books Most of the centrist books fall in the two remaining communities. damic, L. A&Glance, N, www-2005 Workshop on the Weblogging Ecosystem, May Thus the books appear to form communities of copurchasing 10-14,2005,Chiba, Japan. Newman PNAs|June6,2006|vo.103|no.23|8581

et al. (26), which optimizes modularity by using a greedy algo￾rithm; and the extremal optimization algorithm of Duch and Arenas (19), which is arguably the best previously existing method, by standard measures (8), if one discounts methods impractical for large networks, such as exhaustive enumeration of all partitions or simulated annealing. Table 1 reveals some interesting patterns. The algorithm clearly outperforms the methods of Girvan and Newman and Clauset et al. for all of the networks in the task of optimizing the modularity. The extremal optimization method, on the other hand, is more competitive. For the smaller networks, up to 1,000 vertices, there is essen￾tially no difference in performance between the method of this article and extremal optimization; the modularity values for the divisions found by the two algorithms differ by no more than a few parts in 1,000 for any given network. For the larger networks, however, the spectral algorithm does better than extremal op￾timization, and furthermore the gap widens as network size increases, to a maximum modularity difference of 6% for the largest network studied. For the very large networks that have been of particular interest in the last few years, therefore, it appears that the spectral method for detecting community structure may be the most effective of the methods considered here. The modularity values given in Table 1 provide a useful quantitative measure of the success of the algorithm when applied to real-world problems. It is worthwhile, however, also to confirm that it returns sensible divisions of networks in practice. I have given one example demonstrating such a division in Fig. 2. I have also checked the method against many of the example networks used in previous studies (10, 17). Here I give two more examples, both involving network representations of American politics. The first example is a network of books about politics, compiled by V. Krebs (personal communication). In this net￾work the vertices represent 105 recent books on American politics bought from the on-line bookseller Amazon.com, and edges join pairs of books that are frequently purchased by the same buyer. Books were divided (by me) according to their stated or apparent political alignment, liberal or conservative, except for a small number of books that were explicitly bipartisan or centrist, or had no clear affiliation. Fig. 3 shows the result of feeding this network through the algorithm. The algorithm finds four communities of vertices, denoted by the dotted lines in Fig. 3, with a modularity of 0.526. As can be seen, one of these communities consists almost entirely of liberal books and one almost entirely of conservative books. Most of the centrist books fall in the two remaining communities. Thus the books appear to form communities of copurchasing that align closely with political views, a finding that encourages us to believe that the algorithm is capable of extracting mean￾ingful results from raw network data. It is particularly interesting to note that the centrist books belong to their own communities and are not, in most cases, merely lumped in with the liberals or conservatives; this finding may indicate that political moderates form their own purchasing community. For the second example, we consider a network of political commentary web sites, also called ‘‘weblogs’’ or ‘‘blogs,’’ com￾piled from online directories by Adamic and Glance,† who also assigned a political alignment, conservative or liberal, to each blog based on content. The 1,225 vertices in the network studied here correspond to the 1,225 blogs in the largest component of Adamic and Glance’s network, and undirected edges connect vertices if either of the corresponding blogs contained a hyper￾link to the other on its front page. On feeding this network through the algorithm we discover that the network divides cleanly into conservative and liberal communities and, remark￾ably, the optimal modularity of Q  0.426 is found for a division into just two communities. One community has 638 vertices of which 620 (97%) represent conservative blogs. The other has 587 vertices of which 548 (93%) represent liberal blogs. The algo￾rithm found no division of either of these two groups that gives any positive contribution to the modularity; as near as the algorithm is able to tell, these groups are ‘‘indivisible’’ in the sense defined here. This behavior is unique in my experience among networks of this size and is perhaps a testament not only to the widely noted polarization of the current political land￾scape in the United States but also to the strong cohesion of the two factions. Implementation This algorithm is fast as well as accurate. The most time￾consuming part of the algorithm is the evaluation of the leading eigenvector of the modularity matrix. The fastest method for finding this eigenvector is the simple power method, the re￾peated multiplication of the matrix into a trial vector. Although it appears at first glance that matrix multiplications will be slow, taking O(n2) operations each because the modularity matrix is dense, we can in fact perform them much faster by exploiting the particular structure of the matrix. Writing B  A kkT2m, where A is the adjacency matrix and k is the vector whose elements are the degrees of the vertices, the product of B and an arbitrary vector x can be written Bx  Ax kkTx 2m . [7] The first term is a standard sparse matrix multiplication taking time O(m  n). The inner product kTx takes time O(n) to evaluate and hence the second term can be evaluated in total time O(n) also. Thus the complete multiplication can be per￾formed in O(m  n) time. Typically O(n) such multiplications are needed to converge to the leading eigenvector, for a running time of O[(m  n)n] overall. Often we are concerned with sparse graphs with m n, in which case the running time becomes O(n2). It is a simple matter to extend this procedure to find the leading eigenvector of the generalized modularity matrix, Eq. 6, also. Although I will not go through the details here, it is straight￾forward to show that the fine-tuning stage of the algorithm can also be completed in O[(m  n)n] time, so that the combined running time for a single split of a graph or subgraph scales as O[(m  n)n], or O(n2) on a sparse graph. †Adamic, L. A. & Glance, N., WWW-2005 Workshop on the Weblogging Ecosystem, May 10 –14, 2005, Chiba, Japan. Fig. 3. Krebs’ network of books on American politics. Vertices represent books and edges join books frequently purchased by the same readers. Dashed lines divide the four communities found by the algorithm, and shapes represent the political alignment of the books (circles are liberal, squares are conservative, and triangles are centrist or unaligned). Newman PNAS  June 6, 2006  vol. 103  no. 23  8581 APPLIED MATHEMATICS

We then repeat the division into two parts until the network mization task in which one searches for the maximal value of the is reduced to its component indivisible subgraphs. The running quantity known modularity over possible divisions of a time of the entire process depends on the depth of the tree or network. We have shown that this problem can be rewritten in the worst case the dendrogram hapeateth linear in n, but only modularity matrix, and by exploiting this transformation created a small fraction of possible dendrograms realize this worst algorithm detection that demons case. A more realistic figure for running time is given by the bly outperforms the best previous general-purpose algorithms in average depth of the dendrogram, which goes an average running time for the whole algorithm of o(n log n pgn, giving terms of both quality of results and speed of execution. The algorithm has been applied to a variety of real-world network In the sparse case. This is considerably better than the o(n) data sets, including social and biological examples, the results algorithm(19). It is not as good as the O(nlog2n)running time works and to give both intuitively reasonable divisions of net- the extremal optimization for the greedy algorithm of ref. 26, but the results are of far quantitatively better divisions as measured by the better quality than those for the greedy algorithm. In practice nning times are reasonable for networks up to 100,000 After this article was submitted. I was informed of a recent vertices with current computers. For the largest of the net works studied here. the collaboration network. which has ar to Eq. 2 was derived and used as the basis for a modularity imization algorithm quite different from the one presented here. I 27,000 vertices, the algorithm takes s20 min to run on a thank Christian Pich for br standard personal computer(circa 20 06) Conclusions I thank Lada Adamic. Alex Arenas and valdis Krebs for oviding network data. This work was funded in part b In this article we have examined the problem of detecting Science Foundation Grant DMS-0405348 and the James S McDonnell community structure in networks, which is framed as an opti- Foundation. 1. Watts D.J. tzS.H.(198)Nan393,440-442. 18. Newman, M.EJ(2004)Plys. Rev. E 69, 066133. 2. Barabasi, A.-L& Albert, R(1999)Science 286, 509-512. 19. Duch, J& Arenas, A.(2005) Pinys Rev. E 72, 027104 3. Milo, R, Shen-Orr, S Itzkovitz, S, Kashtan, N, Chklovski, D. Alon, U. 20. Chung. F R K(1997)Spectral Graph Theory, CBMS Regional Conference (2002) Science298824-827. Mathematics(Am. Math. Soc. Providence. RD). no g Rev. Mod. Phys. 74. 47-9 21. Fiedler, M.(1973)Czech. Math. J. 5. Dorogovtsev, s. N.& Men F.(2002)Ad.Phys.51,1079-1187 22 Pothen, A, Simon, H.& Liou, K-P.(1990)SLAM J Matrix Anal. Appl. 11, an.上2 430-452 J.B38.321-330 8. Danon, L Duch. J. Dia 23. Zachary, w. W.(1977). Anthropol. Res 33, 452-473 A.& Arenas, A.(2005)J Star. Mech., 24. Radicchi, F, Castellano, C, Cecconi, F, Loreto, V& Parisi, D.(2004)Proc. Natl. Acad. Sci. USA 101. 2658-2663 9. Flake, G. W, Lawrence, S.R, Giles, C. L.& Coetzee, F M.(2002)IEEE 25. Kernighan, B. W.& Lin, S(1970) Bell System TechJ.49, 291-307. 26. Clauset, A, Newman, M. E J& Moore, C. (2004)P/rys. Rev. E 70, 066111 10. Girvan. M.& Newman, M. E. J.(2002)Proc. NatL. Acad. Sci. USA 99, 27. Gleiser. P.& Danon, L (2003)Adv Complex Systems 6, 565-573 7821-782 11. Holme, P, Huss, M. Jeong. H. (2003)Bioin cs19,532-538 28. Jeong. H, Tombor, B, Albert, R, Oltvai, Z N. Barabasi, A-L (2000)Nature 13. Elsner, U. (1997) Graph Partitioning: A Survey (Technische Universitat Chem- 29. Guimera, R, Danon, L, Diaz-Guilera, A, Giralt, F. Arenas, A(2003)P/. 97-27 14. Fjallstrom, P-O.(1998)Linkoping Electronic Articles in Computer and as, A. Diaz-Guilera, A, Streib. D. Amaral. mationScienceVol.3.avAilableatwww.ep.liu.se/ea/cis/1998/006.Accessed La.N.(2002)e-printArchivehttp://xxx.lanl-gow/abs/cond-mat/0206240 31. Newman, M. E.J.(2001)Proc. Natl. Acad. Sci. USA 98, 404-409 15.White, HC,Boorman,.A& Breiger, RL(1976)Am J Socio/81, 730-779. 32. White, S.& Smyth, P.(2005)in Proceedings of the Sth SIAM Intenational 16. Wasserman, S& Faust, K(1994)Social Network Analysis(Cambridge Univ Conference on Data Mining, eds. Kargupta, H, Srivastava, J, Kamath, C& Press, Cambridge, U. K). Goodman, A. Society for Industrial and Applied 17. Newman, M.E.J.& Girvan, M.(2004) Plys. Rev. E 69, 026113 8582Iwww.pnas.org/egi/doi/10.1073/pnas.0601602103

We then repeat the division into two parts until the network is reduced to its component indivisible subgraphs. The running time of the entire process depends on the depth of the tree or ‘‘dendrogram’’ formed by these repeated divisions (10, 16). In the worst case the dendrogram has depth linear in n, but only a small fraction of possible dendrograms realize this worst case. A more realistic figure for running time is given by the average depth of the dendrogram, which goes as log n, giving an average running time for the whole algorithm of O(n2log n) in the sparse case. This is considerably better than the O(n3) running time of the betweenness algorithm (10), and slightly better than the O(n2log2 n) of the extremal optimization algorithm (19). It is not as good as the O(nlog2 n) running time for the greedy algorithm of ref. 26, but the results are of far better quality than those for the greedy algorithm. In practice, running times are reasonable for networks up to 100,000 vertices with current computers. For the largest of the net￾works studied here, the collaboration network, which has 27,000 vertices, the algorithm takes 20 min to run on a standard personal computer (circa 2006). Conclusions In this article we have examined the problem of detecting community structure in networks, which is framed as an opti￾mization task in which one searches for the maximal value of the quantity known as modularity over possible divisions of a network. We have shown that this problem can be rewritten in terms of the eigenvalues and eigenvectors of a matrix called the modularity matrix, and by exploiting this transformation created a computer algorithm for community detection that demonstra￾bly outperforms the best previous general-purpose algorithms in terms of both quality of results and speed of execution. The algorithm has been applied to a variety of real-world network data sets, including social and biological examples, the results showing it to give both intuitively reasonable divisions of net￾works and quantitatively better divisions as measured by the modularity. Note. After this article was submitted, I was informed of a recent conference presentation by White and Smyth (32) in which a result similar to Eq. 2 was derived and used as the basis for a modularity maximization algorithm quite different from the one presented here. I thank Christian Pich for bringing this presentation to my attention. I thank Lada Adamic, Alex Arenas, and Valdis Krebs for graciously providing network data. This work was funded in part by National Science Foundation Grant DMS-0405348 and the James S. McDonnell Foundation. 1. Watts, D. J. & Strogatz, S. H. (1998) Nature 393, 440–442. 2. Baraba´si, A.-L. & Albert, R. (1999) Science 286, 509–512. 3. Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N., Chklovskii, D. & Alon, U. (2002) Science 298, 824–827. 4. Albert, R. & Baraba´si, A.-L. (2002) Rev. Mod. Phys. 74, 47–97. 5. Dorogovtsev, S. N. & Mendes, J. F. F. (2002) Adv. Phys. 51, 1079–1187. 6. Newman, M. E. J. (2003) SIAM Rev. 45, 167–256. 7. Newman, M. E. J. (2004) Eur. Phys. J. B 38, 321–330. 8. Danon, L., Duch, J., Diaz-Guilera, A. & Arenas, A. (2005) J. Stat. Mech., P09008. 9. Flake, G. W., Lawrence, S. R., Giles, C. L. & Coetzee, F. M. (2002) IEEE Computer 35, 66–71. 10. Girvan, M. & Newman, M. E. J. (2002) Proc. Natl. Acad. Sci. USA 99, 7821–7826. 11. Holme, P., Huss, M. & Jeong, H. (2003) Bioinformatics 19, 532–538. 12. Guimera`, R. & Amaral, L. A. N. (2005) Nature 433, 895–900. 13. Elsner, U. (1997) Graph Partitioning: A Survey (Technische Universita¨t Chem￾nitz, Chemnitz, Germany), Technical Report 97-27. 14. Fja¨llstro¨m, P.-O. (1998) Linko¨ping Electronic Articles in Computer and Infor￾mation Science, Vol. 3. Available at www.ep.liu.seeacis1998006. Accessed May 10, 2006. 15. White, H. C., Boorman, S. A. & Breiger, R. L. (1976) Am. J. Sociol. 81, 730–779. 16. Wasserman, S. & Faust, K. (1994) Social Network Analysis (Cambridge Univ. Press, Cambridge, U.K.). 17. Newman, M. E. J. & Girvan, M. (2004) Phys. Rev. E 69, 026113. 18. Newman, M. E. J. (2004) Phys. Rev. E 69, 066133. 19. Duch, J. & Arenas, A. (2005) Phys. Rev. E 72, 027104. 20. Chung, F. R. K. (1997) Spectral Graph Theory, CBMS Regional Conference Series in Mathematics (Am. Math. Soc., Providence, RI), no. 92. 21. Fiedler, M. (1973) Czech. Math. J. 23, 298–305. 22. Pothen, A., Simon, H. & Liou, K.-P. (1990) SIAM J. Matrix Anal. Appl. 11, 430–452. 23. Zachary, W. W. (1977) J. Anthropol. Res. 33, 452–473. 24. Radicchi, F., Castellano, C., Cecconi, F., Loreto, V. & Parisi, D. (2004) Proc. Natl. Acad. Sci. USA 101, 2658–2663. 25. Kernighan, B. W. & Lin, S. (1970) Bell System Tech. J. 49, 291–307. 26. Clauset, A., Newman, M. E. J. & Moore, C. (2004) Phys. Rev. E 70, 066111. 27. Gleiser, P. & Danon, L. (2003) Adv. Complex Systems 6, 565–573. 28. Jeong, H., Tombor, B., Albert, R., Oltvai, Z. N. & Baraba´si, A.-L. (2000) Nature 407, 651–654. 29. Guimera`, R., Danon, L., Dı´az-Guilera, A., Giralt, F. & Arenas, A. (2003) Phys. Rev. E 68, 065103. 30. Guardiola, X., Guimera`, R., Arenas, A., Diaz-Guilera, A., Streib, D. & Amaral, L. A. N. (2002) e-Print Archive, http:xxx.lanl.govabscond-mat0206240. 31. Newman, M. E. J. (2001) Proc. Natl. Acad. Sci. USA 98, 404–409. 32. White, S. & Smyth, P. (2005) in Proceedings of the 5th SIAM International Conference on Data Mining, eds. Kargupta, H., Srivastava, J., Kamath, C. & Goodman, A. (Society for Industrial and Applied Mathematics, Philadelphia), pp. 274–285. 8582  www.pnas.orgcgidoi10.1073pnas.0601602103 Newman

点击下载完整版文档(PDF)VIP每日下载上限内不扣除下载券和下载次数;
按次数下载不扣除下载券;
24小时内重复下载只扣除一次;
顺序:VIP每日次数-->可用次数-->下载券;
已到末页,全文结束
相关文档

关于我们|帮助中心|下载说明|相关软件|意见反馈|联系我们

Copyright © 2008-现在 cucdc.com 高等教育资讯网 版权所有