community detectionin quantum complex networksmodularity functions: the first is the coherent...

16
arXiv:1310.6638v3 [quant-ph] 22 Oct 2014 Community Detection in Quantum Complex Networks Mauro Faccin, 1, Piotr Migda l, 2, 1 Tomi H. Johnson, 3, 4, 5, 1 Ville Bergholm, 1 and Jacob D. Biamonte 1 1 ISI Foundation, Via Alassio 11/c, 10126 Torino, Italy 2 ICFO–Institut de Ci` encies Fot`oniques, 08860 Castelldefels (Barcelona), Spain 3 Centre for Quantum Technologies, National University of Singapore, 3 Science Drive 2, 117543, Singapore 4 Clarendon Laboratory, University of Oxford, Parks Road, Oxford OX1 3PU, United Kingdom 5 Keble College, University of Oxford, Parks Road, Oxford OX1 3PG, United Kingdom Determining community structure is a central topic in the study of complex networks, be it tech- nological, social, biological or chemical, in static or interacting systems. In this paper, we extend the concept of community detection from classical to quantum systems — a crucial missing compo- nent of a theory of complex networks based on quantum mechanics. We demonstrate that certain quantum mechanical effects cannot be captured using current classical complex network tools, and provide new methods that overcome these problems. Our approaches are based on defining close- ness measures between nodes, and then maximizing modularity with hierarchical clustering. Our closeness functions are based on quantum transport probability and state fidelity, two important quantities in quantum information theory. To illustrate the effectiveness of our approach in detect- ing community structure in quantum systems, we provide several examples, including a naturally occurring light harvesting complex, LHCII. The prediction of our simplest algorithm, semi-classical in nature, mostly agrees with a proposed partitioning for the LHCII found in quantum chemistry literature, whereas our fully quantum treatment of the problem of uncover a new, consistent and appropriately quantum community structure. The identification of the community structure within a network addresses the problem of characterizing the mesoscopic boundary between the microscopic scale of basic network components (herein called nodes) and the macroscopic scale of the whole network [13]. In non- quantum networks, the detection of community struc- tures dates back to Rice [4]. Such analysis has revealed countless important hierarchies of community groupings within real-world complex networks. Salient examples can be found in social networks such as human [5] or an- imal relationships [6], biological [710], biochemical [11] and technological [12, 13] networks, as well as numerous others (see Ref. [1]). In quantum networks, as researchers explore networks of an increasingly non-trivial geometry and large size [1417], their analysis and understanding will involve identifying non-trivial community structures. In this article, we devise methods to perform this task, providing an important missing component in the recent drive to unite quantum physics and complex network sci- ence [18, 19]. For quantum systems, beyond being merely a tool for analysis following simulations, community partitioning is closely related to performing the simulations themselves. Simulation is generally a difficult task [20], e.g. simulat- ing exciton transport in dissipative quantum biological networks [2126]. The amount of resources required to exactly simulate such processes scales exponentially with the number of nodes. To overcome this, one must in gen- eral seek to describe only limited correlations between certain parts of the network [27]. Mean-field [2831] and tensor-network methods [32, 33] assume correlations Electronic address: [email protected] between bi-partitions of the system along some node- structure to be zero or limited by an area law. Hartree- Fock methods assume limited correlations between parti- cles [34, 35]. Thus planning a simulation involves identi- fying a partitioning of a system for which it is appropri- ate to limit inter-community correlations, i.e. is a type of community detection. We apply our detection methods to artificial networks and the real-world light harvesting complex II (LHCII) network. In past works, researchers have divided the LHCII by hand in order to gain more insight into the system dynamics [3638]. Meanwhile, our methods opti- mize the task of identifying communities within a quan- tum network ab initio and, as we will show, the resulting communities consistently point towards a structure that is different to those previously identified for the LHCII. For larger networks, as with our artifical examples, au- tomatic methods would appear to be the only feasible option. Our specific approach is to generate a hierarchical community structure [39] by defining both inter-node and inter-community closeness. The optimum level in the hierarchy is determined by a modularity-based mea- sure, which quantifies how good a choice of communi- ties is for the quantum network on average relative to an appropriately randomized version of the network. Al- though modularity-based methods are known to struggle with large sparse networks [40, 41], this work focuses on quantum systems whose size remains much smaller than the usual targets of classical community detection algo- rithms. While the backbone of our quantum community method is shared with classical methods, the physical properties used to characterize a good community in a quantum must necessarily be very different to the prop-

Upload: others

Post on 20-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

arX

iv:1

310.

6638

v3 [

quan

t-ph

] 2

2 O

ct 2

014

Community Detection in Quantum Complex Networks

Mauro Faccin,1, ∗ Piotr Migda l,2, 1 Tomi H. Johnson,3, 4, 5, 1 Ville Bergholm,1 and Jacob D. Biamonte1

1ISI Foundation, Via Alassio 11/c, 10126 Torino, Italy2ICFO–Institut de Ciencies Fotoniques, 08860 Castelldefels (Barcelona), Spain

3Centre for Quantum Technologies, National University of Singapore, 3 Science Drive 2, 117543, Singapore4Clarendon Laboratory, University of Oxford, Parks Road, Oxford OX1 3PU, United Kingdom

5Keble College, University of Oxford, Parks Road, Oxford OX1 3PG, United Kingdom

Determining community structure is a central topic in the study of complex networks, be it tech-nological, social, biological or chemical, in static or interacting systems. In this paper, we extendthe concept of community detection from classical to quantum systems — a crucial missing compo-nent of a theory of complex networks based on quantum mechanics. We demonstrate that certainquantum mechanical effects cannot be captured using current classical complex network tools, andprovide new methods that overcome these problems. Our approaches are based on defining close-ness measures between nodes, and then maximizing modularity with hierarchical clustering. Ourcloseness functions are based on quantum transport probability and state fidelity, two importantquantities in quantum information theory. To illustrate the effectiveness of our approach in detect-ing community structure in quantum systems, we provide several examples, including a naturallyoccurring light harvesting complex, LHCII. The prediction of our simplest algorithm, semi-classicalin nature, mostly agrees with a proposed partitioning for the LHCII found in quantum chemistryliterature, whereas our fully quantum treatment of the problem of uncover a new, consistent andappropriately quantum community structure.

The identification of the community structure withina network addresses the problem of characterizing themesoscopic boundary between the microscopic scale ofbasic network components (herein called nodes) and themacroscopic scale of the whole network [1–3]. In non-quantum networks, the detection of community struc-tures dates back to Rice [4]. Such analysis has revealedcountless important hierarchies of community groupingswithin real-world complex networks. Salient examplescan be found in social networks such as human [5] or an-imal relationships [6], biological [7–10], biochemical [11]and technological [12, 13] networks, as well as numerousothers (see Ref. [1]). In quantum networks, as researchersexplore networks of an increasingly non-trivial geometryand large size [14–17], their analysis and understandingwill involve identifying non-trivial community structures.In this article, we devise methods to perform this task,providing an important missing component in the recentdrive to unite quantum physics and complex network sci-ence [18, 19].

For quantum systems, beyond being merely a tool foranalysis following simulations, community partitioning isclosely related to performing the simulations themselves.Simulation is generally a difficult task [20], e.g. simulat-ing exciton transport in dissipative quantum biologicalnetworks [21–26]. The amount of resources required toexactly simulate such processes scales exponentially withthe number of nodes. To overcome this, one must in gen-eral seek to describe only limited correlations betweencertain parts of the network [27]. Mean-field [28–31]and tensor-network methods [32, 33] assume correlations

∗Electronic address: [email protected]

between bi-partitions of the system along some node-structure to be zero or limited by an area law. Hartree-Fock methods assume limited correlations between parti-cles [34, 35]. Thus planning a simulation involves identi-fying a partitioning of a system for which it is appropri-ate to limit inter-community correlations, i.e. is a type ofcommunity detection.

We apply our detection methods to artificial networksand the real-world light harvesting complex II (LHCII)network. In past works, researchers have divided theLHCII by hand in order to gain more insight into thesystem dynamics [36–38]. Meanwhile, our methods opti-mize the task of identifying communities within a quan-tum network ab initio and, as we will show, the resultingcommunities consistently point towards a structure thatis different to those previously identified for the LHCII.For larger networks, as with our artifical examples, au-tomatic methods would appear to be the only feasibleoption.

Our specific approach is to generate a hierarchicalcommunity structure [39] by defining both inter-nodeand inter-community closeness. The optimum level inthe hierarchy is determined by a modularity-based mea-sure, which quantifies how good a choice of communi-ties is for the quantum network on average relative toan appropriately randomized version of the network. Al-though modularity-based methods are known to strugglewith large sparse networks [40, 41], this work focuses onquantum systems whose size remains much smaller thanthe usual targets of classical community detection algo-rithms.

While the backbone of our quantum communitymethod is shared with classical methods, the physicalproperties used to characterize a good community in aquantum must necessarily be very different to the prop-

Page 2: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

2

erties used for a classical system. Here we show howtwo quantum properties are used to obtain closeness andmodularity functions: the first is the coherent transportbetween communities and the second is the change inthe states of individual communities during a coherentevolution.

In Section I we will begin by recalling several commonnotions from classical community detection that we relyon in this work. This sets the stage for the developmentof a quantum treatment of community detection in Sec-tion II. We then turn to several examples in Section IIIincluding the LHCII complex mentioned previously, be-fore concluding in Section IV.

I. COMMUNITY DETECTION

Community detection is the partitioning of a set ofnodes N into non-overlapping 1 and non-empty subsetsA, B, C, . . . ⊆ N , called communities, that togethercover N .

There is usually no agreed upon optimal partitioningof nodes into communities. Instead there is an array ofapproaches that differ in both the definition of optimal-ity and the method used to achieve, exactly or approxi-mately, this optimality (see Ref. [3] for a recent review).In classical networks optimality is, for example, definedstatistically [42], e.g. in terms of connectivity [1] or com-

10

20

30

40

50

60

10 20 30 40 50

n

o

d

e

s

nodes

1

00.20.40.60.81

n

o

d

e

s

merging loseness (10−3

)

0

0.1

0.2

0.3

5

10

15

modularitydegree (10−3

)

Figure 1: Hierarchical community structure arising from aquantum evolution. On the left are the closenesses c(i, j) be-tween n = 60 nodes. On the right is the dendrogram showingthe resulting hierarchical community structure. The dashedline shows the optimum level within this hierarchy, accordingto the modularity. The particular example shown here is theone corresponding to Fig. 3d.

1 We do not consider generalizations to overlapping communitieshere.

municability [43, 44], or increasingly, and sometimes re-latedly [45], in terms of stochastic random walks [46–48].Our particular focus is on the latter, since the conceptof transport (e.g. a quantum walk) is central to nearlyall studies conducted in quantum physics. As for achiev-ing optimality, methods include direct maximization viasimulated annealing [10, 40] or, usually faster, iterativedivision or agglomeration of communities [49]. We focuson the latter since it provides a simple and effective wayof revealing a full hierarchical structure of the network,requiring only the definition of the closeness of a pair ofcommunities.

Formally, hierarchical community structure detectionmethods are based on a (symmetric) closeness functionc(A,B) = c(B,A) of two communities A 6= B. In the ag-glomerative approach, at the lowest level of the hierarchy,the nodes are each assigned their own communities. Aniterative procedure then follows, in each step of which theclosest pair of communities (maximum closeness c) aremerged. This procedure ends at the highest level, whereall nodes are in the same community. To avoid instabil-ities in this agglomerative procedure, the closeness func-tion is required to be non-increasing under the mergingof two communities, c(A ∪ B, C) ≤ max(c(A, C), c(B, C)),which allows the representation of the community struc-ture as a linear hierarchy indexed by the merging close-ness. The resulting structure is often represented as adendrogram (as shown in Fig. 1) 2.

This leaves open the question of which level of the hi-erarchy yields the optimal community partitioning. If apartitioning is desired for simulation, for example, thenthere may be a desired maximum size or minimum num-ber of communities. However, without such constraints,one can still ask what is the best choice of communitieswithin those given by the hierarchical structure.

A type of measure that is often used to quantify thequality of a community partitioning choice for this pur-pose is modularity [50–52], denoted Q. It was originallyintroduced in the classical network setting, in which anetwork is specified by a (symmetric) adjacency matrixof (non-negative) elements Aij = Aji ≥ 0 (Aii = 0), eachoff-diagonal element giving the weight of connections be-tween nodes i and j 6= i 3. The modularity attempts tomeasure the fraction of weights connecting elements inthe same community, relative to what might be expected.Specifically, one takes the fraction of intra-communityweights and subtracts the average fraction obtained when

2 In general it may happen that more than one pair of communitiesare at the maximum closeness. In this case the decision on whichpair merges first can influence the structure of the dendrogram,see [39, 69]. In [39] a permutation invariant formulation of theagglomerative algorithm is given, where more than two clusterscan be merged at once. In our work we use this formulationunless stated otherwise.

3 As will become apparent, we need only consider undirected net-works without self-loops.

Page 3: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

3

the start and end points of the connections are reassignedrandomly, subject to the constraint that the total connec-tivity ki =

∑j Aij of each node is fixed. The modularity

is then given by

Q =1

2mtr{CTBC

}, (1)

where m = 12

∑i ki is the total weight of connections,

B is the modularity matrix with elements Bij = Aij −kikj/2m, and C is the community matrix, with elementsCiA equal to unity if i ∈ A, otherwise zero. The mod-ularity then takes values strictly less than one, possiblynegative, and exactly zero in the case that the nodes forma single community.

As we will see, there is no natural adjacency matrixassociated with the quantum network and so for the pur-poses of modularity we use Aij = c(i, j) for i 6= j. Themodularity Q thus measures the fraction of the closenessthat is intra-community, relative to what would occurif the inter-node closeness c(i, j) were randomly mixedwhile fixing the total closeness ki =

∑j 6=i c(i, j) of each

node to all others. Thus both the community structureand optimum partitioning depend solely on the choice ofthe closeness function.

Modularity-based methods such as above are intuitive,fast and on the most part effective, yet we must note thatfor classical systems it has been shown that modularity-based methods suffer from a number of flaws that influ-ence the overall efficacy of those approaches. In Refs. [40,41] modularity-based methods show a poor performancein large, sparse real-world and model networks. This isdue mainly to the resolution limit problem [53], wheresmall communities can be overlooked, and modularitylandscape degeneracy, which strongly influence accuracyin large networks. Another modularity-related problem isthe so-called detectability/undetectability threshold [54–56] where an approximate bi-partition of the system be-comes undetectable in some cases, in particular in pres-ence of degree homogeneity. However, in the presentwork we focus on quantum networks whose size typicallyremains small compared to classical targets of commu-nity detection algorithms, and for which the derived ad-jacency matrices are not sparse. These characteristicshelp to limit the known flaws of our modularity-basedapproach, making it adequate for our purposes.

Finally, once a community partitioning is obtained itis often desired to compare it against another. Herewe use the common normalized mutual information(NMI) [57–59] as a measure of the mutual dependenceof two community partitionings. Each partitioning X ={A,B, . . .} is represented by a probability distributionPX = {|A|/|N |}A∈X , where |A| =

∑i CiA is the num-

ber of nodes in community A. The similarity of twocommunity partitionings X and X ′ depends on the jointdistribution PXX′ = {|A ∩ A′|/|N |}A∈X,A′∈X′ , where|A ∩ A′| =

∑i CiACiA′ is the number of nodes that be-

long to both communities A and A′. Specifically, NMI is

defined as

NMI(X,X ′) =2 I(X,X ′)

H(X) +H(X ′). (2)

Here H(X) is the Shannon entropy of PX , and the mu-tual information I(X,X ′) = H(X) +H(X ′) −H(X,X ′)depends on the entropy H(X,X ′) of the joint distribu-tion PXX′ . The mutual information is the average of theamount of information about the community of a nodein X obtained by learning its community in X ′. The nor-malization ensures that the NMI has a minimum valueof zero and takes its maximum value of unity for twoidentical community partitionings. The symmetry of thedefinition of NMI follows from that of mutual informationand Eq. (2).

II. QUANTUM COMMUNITY DETECTION

The task of community detection has a particular in-terpretation in a quantum setting. The state of a quan-tum system is described in terms of a Hilbert space H,spanned by a complete orthonormal set of basis states{|i〉}i∈N . Each basis state |i〉 can be associated with anode i in a network and often, as in the case of single ex-citon transport, there is a clear choice of basis states thatmakes this abstraction to a spatially distributed networknatural.

The partitioning of nodes into communities then cor-responds to the partitioning of the Hilbert space H =⊕

A∈X VA into mutually orthogonal subspaces VA =spani∈A{|i〉}. As with classical networks, one can thenimagine an assortment of optimality objectives for com-munity detection, for example, to identify a partitioninginto subspaces in which inter-subspace transport is small,or in which the state of the system remains relativelyunchanged within each subspace. In the next two sub-sections we introduce two classes of community closenessmeasures that correspond to these objectives. Technicaldetails can be found in the Supplemental Material.

In what follows, we focus our analysis one an isolatedquantum system governed by Hamiltonian H , which en-ables us to derive convenient closed-form expressions forthe closeness measures. We may expand H in the nodebasis {|i〉}i∈N :

H =∑

ij

Hij |i〉〈j|. (3)

A diagonal element Hii is a real value denoting the en-ergy of state |i〉, whilst an off-diagonal element Hij , i 6= j,is a complex weight denoting the change in the ampli-tude of the wave function during a transition from state|j〉 to |i〉. The matrix formed by these elements can bethought of as a |N | × |N | complex, hermitian adjacencymatrix. In quantum mechanics, complex elements in theHamiltonian lead to a range of phenomena not capturedby real matrices, such as time-reversal symmetry break-ing [19, 60]. In the case where each state |i〉 corresponds

Page 4: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

4

to a particle being localised at a spatially distinct node i,the Hamiltonian describes a spinless single-particle walkwith an energy landcape given by the diagonal elements,and transition amplitudes by the off-diagonal elements.Any quantum evolution can be viewed in this picture,making the single particle spiness walk scenario rathergeneral.

A community partitioning based on a Hamiltonian Hcould be used, among other things, to guide the simula-tion or analysis of a more complete model in the presenceof an environment, where this more complete model maybe much more difficult to describe. Additionally, ourmethod could be generalized to use closeness measuresbased on open-system dynamics obtained numerically.

A. Inter-community transport

Several approaches to detecting communities in clas-sical networks are based on the flow of probabilitythrough the network during a classical random walk [45–48, 61, 62]. In particular, many of these methods seekcommunities for which the inter-community probabilityflow or transport is small. A natural approach to quan-tum community detection is thus to consider the flow ofprobability during a continuous-time quantum walk, andto investigate the change in the probability of observingthe walker within each community:

TX(t) =∑

A∈X

TA(t) =∑

A∈X

1

2|pA {ρ(t)} − pA {ρ(0)}| ,

(4)

where ρ(t) = e−iHtρ(0)eiHt is the state of the walker, attime t, during the walk generated by H , and

pA {ρ} = tr {ΠAρ} , (5)

is the probability of a walker in state ρ being found incommunity A upon a von Neumann-type measurement 4.ΠA =

∑i∈A |i〉〈i| denotes the projector to the A sub-

space.The initial state ρ(0) can be chosen freely. The change

in inter-community transport is clearest when the pro-cess begins either entirely inside or entirely outside eachcommunity. Because of this, we choose the walker to beinitially localized at a single node ρ(0) = |i〉〈i| and then,for symmetry, sum TX(t) over all i ∈ N . This results inthe particularly simple expression

TA(t) =∑

i∈A,j /∈A

Rij(t) +Rji(t)

2=

i∈A,j /∈ARij(t), (6)

4 Equivalently, pA {ρ} is the norm of the projection (performed byprojector ΠA) of the state ρ onto the community subspace VA.

where R(t) is the doubly stochastic transfer matrix whoseelements Rij(t) = |〈i|e−iHt|j〉|2 give the probability of

transport from node j to node i, and R(t) its symmetriza-tion. This is reminiscent of classical community detectionmethods, e.g. [48], using closeness measures based on thetransfer matrix of a classical random walk.

We can thus build a community structure that seeksto reduce TX(t) at each hierarchical level by using thecloseness function

cTt (A,B) =TA(t) + TB(t) − TA∪B(t)

|A||B|

=2

|A||B|∑

i∈A,j∈BRij(t), (7)

where the numerator is the decrement in TX(t) caused bymerging communities A and B. The normalizing factor inEq. (7) avoids the effects due to the uninteresting scalingof the numerator with the community size.

Since a quantum walk does not converge to a stationarystate, a time-average of the closeness defined in Eq. (7)is needed to obtain a quantity that eventually convergeswith increasing time. Given the linearity of the formula-tion, this corresponds to replacing the transport proba-bility Rij(t) in Eq. (7) with its time-average

Rij(t) =1

t

∫ t

0

Rij(t′) dt′. (8)

It follows that, as with similar classical community de-tection methods [46], our method is in fact a class of ap-proaches, each corresponding to a different time t. Theappropriate value of t will depend on the specific appli-cation, for example, a natural time-scale might be thedecoherence time. Not wishing to lose generality and fo-cus on a particular system, we focus here on the shortand long time limits.

In the short time limit t → 0, relevant if tHij ≪ 1

for i 6= j, the averaged transfer matrix Tij(t) is simplyproportional to |Hij |2. Note that in the short time limitthere is no interference between different paths from |i〉 to|j〉, and therefore for short times cTt (i, j) does not dependon the on-site energies Hii or the phases of the hoppingelements Hij . This is because, to leading order in time,interference does not play a role in the transport out of asingle node. For this reason we can refer to this approachas “semi-classical”.

In the long time limit t → ∞, relevant if t is muchlarger than the inverse of the smallest gap between dis-tinct eigenvalues of H , the probabilities are elements ofthe mixing matrix [63],

limt→∞

Rij(t) =∑

k

| 〈i |Λk| j〉 |2, (9)

where Λk is the projector onto the k-th eigenspace of H .This thus provides a simple spectral method for buildingthe community structure.

Page 5: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

5

Note that, unlike in a classical infinitesimal stochastic

walk where each Rij(t) eventually becomes proportionalto the connectivity kj of the final node j, the long timelimit in the quantum setting is non-trivial and, as we will

see, Rij(t) retains a strong impression of the communitystructure for large t 5.

B. Intra-community fidelity

Classical walks, and the community detection meth-ods based on them, are fully described by the evolutionof the probabilities of the walker occupying each node.The previous quantum community detection approach isbased on the evolution of the same probabilities but fora quantum walker. However, quantum walks are richerthan this, they are not fully described by the evolutionof the node-occupation probabilities. We therefore intro-duce another community detection method that capturesthe full quantum dynamics within each community sub-space.

Instead of reducing merely the change in probabilitywithin the community subspaces, we reduce the changein the projection of the quantum state in the commu-nity subspaces. This change is measured using (squared)fidelity, a common measure of distance between two quan-tum states. For a walk beginning in state ρ(0) we there-fore focus on the quantity

FX(t) =∑

A∈X

FA(t) =∑

A∈X

F 2 {ΠAρ(t)ΠA,ΠAρ(0)ΠA} ,

(10)

where ΠAρΠA is the projection of the state ρ onto thesubspace VA and

F {ρ, σ} = tr

{√√ρσ

√ρ

}∈ [0,

√tr{ρ} tr{σ}] (11)

is the fidelity, which is symmetric between ρ and σ.We build a community structure that seeks to maxi-

mize the increase in FX(t) at each hierarchical level byusing the closeness measure

cFt (A,B) =FA∪B(t) − FA(t) − FB(t)

|A||B| ∈ [−1, 1], (12)

i.e. the change in FX(t) caused by merging communitiesA and B. Our choice for the denominator prevents unin-teresting size scaling, as in Eq. (7).

The initial state ρ(0) can be chosen freely. Herewe choose the pure uniform superposition state ρ(0) =

5 Note that, apart from small or large times t, there is no guaranteeof symmetry Rij(t) = Rji(t) in the transfer matrix for a givenHamiltonian. See [19]. Hamiltonians featuring this symmetry,e.g., those with real Hij , are called time-symmetric.

|ψ0〉〈ψ0| satisfying 〈 i |ψ0 〉 = 1/√n for all i. This state

was used to investigate the effects of the connectivity onthe dynamics of a quantum walker in Ref. [18].

As for our other community detection approach, weconsider the time-average of Eq. (12), which yields

cFt (A,B) =2

|A||B|∑

i∈A,j∈BRe(ρij(t)ρji(0)), (13)

where ρij(t) = 1t

∫ t

0 dt′ρij(t′). In the long time limit, thetime-average of the density matrix takes a particularlysimple expression:

limt→∞

ρij(t) =∑

k

Λkρij(0)Λk, (14)

where Λk is as in the previous Sec. II A.The definition of community closeness given in Eq. (12)

can exhibit negative values. In this case the usual defini-tion of modularity fails [64] and one must extended it. Inthis work we use the definition of modularity proposedin Ref. [64], which coincides with Eq. (1) in the case ofnon-negative closeness. The extended definition treatsnegative and positive links separately, and tries to min-imize intra-community negative links while maximizingintra-community positive links.

III. PERFORMANCE ANALYSIS

To analyze the performance of our quantum commu-nity detection methods we apply them to three differentnetworks. The first one (Sec. III A) is a simple quantumnetwork, which we use to highlight how some intuitive no-tions in classical community detection do not necessarilytransfer over to quantum systems. The second example(Sec. III B) is an artificial quantum network designed toexhibit a clear classical community structure, which weshow is different from the quantum community structureobtained and fails to capture significant changes in thisstructure induced by quantum mechanical phases on thehopping elements of the Hamiltonian. The final network(Sec. III C) is a real world quantum biological network, de-scribing the LHCII light harvesting complex, for whichwe find a consistent quantum community structure dif-fering from the community structure cited in the litera-ture. These findings confirm that a quantum mechanicaltreatment of community detection is necessary as classi-cal and semi-classical methods cannot be reproduce thestructures that appropriately capture quantum effects.

Below we will compare quantum community structuresagainst more classical community structures, such theone given by the semi-classical method based on theshort time transport and, in the case of the example ofSec. III B, the classical network from which the quantumnetwork is constructed. Additionally we use a traditionalclassical community detection algorithm, OSLOM [42],an algorithm based on the maximization of the statisti-cal significance of the proposed partitioning, whose input

Page 6: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

6

2 4

1 6

3 5

Disconnected components:

clo

se

far

(a) Transport (b) Fidelity (c) Fidelity (Perturbed)

Phases’ effect on transport:

clo

se

far

(d) Coherent phases (e) Random phases (f) Cancelling phases

Phases’ effect on fidelity:

clo

se

far

(g) Coherent phases (h) Random phases (i) Cancelling phases

Figure 2: Simple quantum network — a graph with six nodes.Each solid line represents transition amplitude Hij = 1. Fordashed and dotted lines the transition amplitude can be eitherzero (a, b and c) or the absolute value is the same |Hij | = 1but phase is (d and g) coherent (all ones), (e and h) randomexp(iϕk) for each link, (f and i) canceling (ones for dashed redand minus one for dotted green). Plots show the node close-ness for both methods based on transport and fidelity (onlythe long-time-averages are considered, in plots (g), (h) and(i) we used a perturbed Hamiltonian to solve the eigenvaluesdegeneracy, this explains the non-symmetric closeness in (i)).

adjacency matrix A must be real. For this purpose weuse the absolute values of the Hamiltonian elements inthe site basis: Aij = |Hij |.

A. Simple quantum network

Here we use a simple six-site network model to studyways in which quantum effects lead to non-intuitive re-sults, and how methods based on different quantum prop-erties can, accordingly, lead to very different choices ofcommunities.

We begin with two disconnected cliques of three nodeseach, where all Hamiltonian matrix elements within the

groups are identical and real. Fig. 2 illustrates this highlysymmetric topology. The community detection methodbased on quantum transport identifies the two fully-connected groups as two separate communities (Fig. 2a),as is expected. Contrastingly, the methods based on fi-delity predict counter-intuitively only a single commu-nity; two disconnected nodes can retain coherence and,by this measure, be considered part of the same commu-nity (Fig. 2b).

This symmetry captured by the fidelity-based commu-nity structure breaks down if we introduce random per-turbations into the Hamiltonian. Specifically, the fidelity-based closeness cFt is sensitive to perturbations of the or-der t−1, above which the community structure is dividedinto the two groups of three (Fig. 2c) expected from trans-port considerations. Thus we may tune the resolution ofthis community structure method to asymmetric pertur-bations by varying t.

Due to quantum interference we expect that the Hamil-tonian phases should significantly affect the quantumcommunity partitioning. The same toy model can beused to demonstrate this effect. For example, consideradding four elements to the Hamiltonian correspondingto hopping from nodes 2 and 3 to 4 and 5 (see diagram inFig. 2). If these hopping elements are all identical to theothers, it is the two nodes, 1 and 6, that are not directlyconnected for which the inter-node transport is largest(and thus their inter-node closeness is the largest). How-ever, when the phases of the four additional elements arerandomized, this transport is decreased. Moreover, whenthe phases are canceling, the transport between nodes 1and 6 is reduced to zero, and the closeness between themis minimized (see Figs. 2d–2f).

The fidelity method has an equally strong dependenceon the phases (see Figs. 2g–2i), with variations in thephases breaking up the network from a large central com-munity (with nodes 1 and 6 alone) into the two previouslyidentified communities.

B. Artificial quantum network

The Hamiltonian of our second quantum network isconstructed from the adjacency matrix A of a classicalunweighted, undirected network exhibiting a clear classi-cal partitioning, using the relation Hij = Aij . We con-struct A using the algorithm proposed by Lancichinettiet al. in Ref. [65], which provides a method to constructa network with heterogeneous distribution both for thenode degree and for the communities dimension and acontrollable inter-community connection. We start witha rather small network of 60 nodes with average intra-community connectivity 〈k〉 = 6, and only 5% of theedges are rewired to join communities. The networkis depicted in Fig. 3a. To confirm the expected, theknown classical community structure is indeed obtained

Page 7: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

7

(a) Original data (b) Transport; t→ 0 (c) Transport; t→ ∞ (d) Fidelity; t→ ∞

(e) OSLOM

Dependence on phases:

0.0

0.2

0.4

0.6

0.8

1.0

0 π/4 π/2 3π/2 π

N

M

I

Phases σ

cT0cT∞cF∞OSLOM

0.0

0.2

0.4

0.6

0.8

1.0

0 π/4 π/2 3π/2 π

N

M

I

Phases σ

(f) Original partitioning (g) Classical model

Figure 3: Artificial community structure. (a) Classical community structure used in creating the network. (b–e) Communitypartitionings found using the three quantum methods and OSLOM. (f,g) Behavior of the approaches as the phases of theHamiltonian elements are randomly sampled from a Gaussian distribution of width σ. The mean NMI, compared with zerophase partitioning (f) and the classical model data (g), over 200 samplings of the phase distribution is plotted. The standarddeviation is indicated by the shading. Both OSLOM and cT0 are insensitive to phases and thus do not respond to the changesin the Hamiltonian.

by the semi-classical short-time-transport algorithm 6

and the OSLOM algorithm (see Figs. 3b–3e), achievingNMI = 0.953 and NMI = 0.975 with the known structure,respectively.

The quantum methods based on the long-time averageof both transport and fidelity reproduce the main fea-tures of the original community structure while unveil-ing new characteristics. The transport-based long-timeaverage method (NMI = 0.82 relative to the classical par-titioning) exhibits disconnected communities, i.e. the cor-responding subgraph is disconnected. This behavior canbe explained by interference-enhanced quantum walkerdynamics, as exhibited by the toy model in the previ-ous subsection. The long-time average fidelity method(NMI = 0.85) returns the four main classical communi-ties plus a number of single-node communities. Both

6 In the case of short-time transport, a small perturbation was alsoadded to the closeness function in order to break the symmetriesof the system.

methods demonstrate that the quantum and classicalcommunity structures are unsurprisingly different, withthe quantum community structure clearly dependent onthe quantum property being optimized, more so than thedifferent classical partitionings.

Adjusted phases

As shown in Sec. III A, due to interference the dynam-ics of the quantum system can change drastically if thephases of the Hamiltonian elements are non-zero. Thisis known as a chiral quantum walk [19]. Such walks ex-hibit, for example, time-reversal symmetry breaking oftransport between sites [19] and it has been proposedthat nature might actually make use of phase controlledinterference in transport processes [66]. OSLOM, oursemi-classical short-time transport algorithm and otherclassical community partitioning methods are insensitiveto changes in the hopping phases. Thus, by establish-ing that the quantum community structure is sensitive

Page 8: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

8

to such changes in phase, as expected from above, weshow that classical methods are inadequate for findingquantum community structure.

To analyze this effect we take the previous network andadjust the phases of the Hamiltonian terms while pre-serving their absolute values. Specifically, the phases aresampled randomly from a normal distribution with meanzero and standard deviation σ. We find that, typically,as the standard deviation σ increases, when comparingquantum communities and the corresponding communi-ties without phases the NMI between them decreases, asshown in Fig. 3f. A similar deviation reflects on the com-parison with the classical communities used to constructthe system, shown in Fig. 3g. This sensitivity of thequantum community structures to phases, as revealed bythe NMI, confirms the expected inadequacy of classicalmethods. The partitioning based on long-time averagefidelity seems to be the most sensitive to phases.

C. Light-harvesting complex

An increasing number of biological networks of non-trivial topology are being described using quantum me-chanics. For example, light harvesting complexes havedrawn significant attention in the quantum informationcommunity.

One of these is the LHCII, a two-layer 14-chromophorecomplex embedded into a protein matrix (see Fig. 4 fora sketch) that collects light energy and directs it towardthe reaction center where it is transformed into chemicalenergy. The system can be described as a network of 14sites connected with a non-trivial topology. The single-exciton subspace is spanned by 14 basis states, each corre-sponding to a node in the network, and the Hamiltonianin this basis was found in Ref. [38].

In a widely adopted chromophore community struc-ture [37], the sites are partitioned by hand into communi-ties according to their physical closeness (e.g. there areno communities spanning the two layers of the complex),and the strength of Hamiltonian couplings (see the topright of Fig. 4). Here, we apply our ab initio automatedquantum community detection algorithms to the sameHamiltonian.

All of our approaches predict a modified partitioning tothat commonly used in the literature. The method basedon short-time transport returns communities that do notconnect the two layers. This semi-classical approach re-lies only on the coupling strength of the system, withoutconsidering interference effects, and provides the closestpartitioning to the one provided by the literature (alsorelying only on the coupling strengths). Meanwhile, themethods based on the long-time transport and fidelityreturn very similar community partitionings, in whichnode 6 on one layer and node 9 on the other are in thesame community. These two long-time community parti-tionings are identical, except one of the communities pre-dicted by the fidelity based method is split when using

the transport based method. It is therefore a differencein modularity only.

The classical OSLOM algorithm fails spectacularly: itgives only one significant community involving nodes 11and 12 which exhibit the highest coupling strength. Ifassigning a community to each node is forced, a uniquecommunity with all nodes is provided.

Note that here we have used the LHCII closed-systemdynamics, valid only for short times, to partition it. Asexplained in Sec. II, for the purpose of analysis one couldalternatively use the less tractable open-system dynam-ics to obtain a partitioning that reflects the environmentof the LHCII [26]. However, we argue that communitypartitioning, e.g. that based on the closed-system dynam-ics, is essential in devising approaches to simulating thefull open-system dynamics.

IV. DISCUSSION

We have developed methods to detect communitystructure in quantum systems, thereby extending thepurview of community detection from classical networksto include quantum networks. Our approach involves thedevelopment of a number of methods that focus on dif-ferent characteristics of the system and return a commu-nity structure reflecting that specific characteristic. Thevariation of the quantum community structure with theproperty on which this structure is based seems greaterthan for classical community structures.

All our methods are based on the full unitary dynam-ics of the system, as described by the Hamiltonian, andaccount for quantum effects such as coherent evolutionand interference. In fact phases are often fundamental tocharacterizing the system evolution. For example, Harelet al. [66] have shown that in light harvesting complexesinterference between pathways is important even at roomtemperature. In our light harvesting complex example(see Sec.III C), the ab-initio community structures pro-vided by the long-time measures propose consistent com-munities that stretch across the lumenal and stromal lay-ers of the complex, absent in the structure proposed bythe community.

Since we consider time evolution, the averaging time tacts as a tuning parameter for the partitioning methods.In the case of transport it transforms the method from asemi-classical approach (t→ 0) to a fully quantum-awaremeasure (t → ∞), For all times, the complexity of our al-gorithms scales polynomially in the number of nodes |N |,at worst O(|N |3) if the diagonalization of H is required.This allows the study of networks with node numbers upthe thousands and tens of thousands, which is appropri-ate for the real-world quantum networks currently beingconsidered.

As with classical community structure, there are manypossible definitions of a quantum community. We re-stricted ourselves to two broad classes based on transportand fidelity under coherent evolution, both based on dy-

Page 9: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

9

12

3

4

5

6

78

9

10

11

12

13

14

12

3

4

5

6

78

9

10

11

12

13

141

2

3

4

5

6

78

9

10

11

12

13

141

2

3

4

5

6

78

9

10

11

12

13

14

(a) Transport; t → 0 (b) Transport; t → ∞ (c) Fidelity; t → ∞

Figure 4: Light harvesting complex II (LHCII). (top left) Monomeric subunit of the LHCII complex with pigments Chl-a(red) and Chl-b (green) packed in the protein matrix (gray). (top center) Schematic representation of Chl-a and Chl-b inthe monomeric subunit, here the labeling follows the usual nomenclature (b601, a602. . . ). (top right) Network representationof the pigments in circular layout, colors represent the typical partitioning of the pigments into communities. The widths ofthe links represent the strength of the couplings |Hij | between nodes. Here the labels maintain only the ordering (b601→1,a602→2,. . . ). (a,b,c) Quantum communities as found by the different quantum community detection methods. Link widthdenotes the pairwise closeness of the nodes.

namics, though in the limits considered in this paper theclosenesses and thus quantum community structure canbe expressed purely in terms of static properties. Weend by briefly discussing some other possible definitionsbased on statics (the earliest classical community defini-tions were based on statics [67]). The first type is basedon some quantum state |ψ〉, e.g. the ground state of H .We might wish to partition the network by repeatedlydiving the network in two based on minimally entangledbipartitions. This could be viewed as identifying opti-mum communities for some cluster-based mean-field-likesimulation [31] whose entanglement structure is expectedto be similar to |ψ〉. The second type is based directly onthe spectrum of the Hamiltonian H . We might partitionthe Hilbert space into unions of the eigenspaces of H bytreating the corresponding eigenvalues as 1D coordinatesand applying a traditional agglomerative or divisive clus-tering algorithm on them. Note that the resulting parti-tioning would normally not be in the position basis.

The use of community detection in quantum systemsaddresses an open challenge in the drive to unite quan-tum physics and complex network science, and we expectsuch partitioning, based on our definitions or extensionssuch as above, to be used extensively in making the largequantum systems currently being targeted by quantum

physicists tractable to numerical analysis. Conversely,quantum measures have also been shown to add novelperspectives to classical network analysis [68].

Acknowledgments

We thank Michele Allegra, Leonardo Banchi, Gio-vanni Petri and Zoltan Zimboras for fruitful discussions.MF, THJ and JDB completed part of this study whilevisiting the Institute for Quantum Computing, at theUniversity of Waterloo. PM acknowledges the SpanishMINCIN/MINECO project TOQATA (FIS2008-00784),EU Integrated Projects AQUTE and SIQS, and HIS-TERA project DIQUIP. THJ acknowledges the Euro-pean Research Council under the European Union’s Sev-enth Framework Programme (FP7/2007-2013) / ERCGrant Agreement No. 319286, and the National ResearchFoundation and the Ministry of Education of Singaporefor support. JDB acknowledges the Foundational Ques-tions Institute (under grant FQXi-RFP3-1322) for finan-cial support. All authors acknowledge the Q-ARACNEproject funded by the Fondazione Compagnia di SanPaolo.

Page 10: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

10

[1] M. Girvan and M. E. J. Newman. Community structurein social and biological networks. Proceedings of the Na-tional Academy of Sciences, 99(12):7821–7826, 2002.

[2] Mason A Porter, Jukka-Pekka Onnela, and Peter JMucha. Communities in networks. Notices of the AMS,56(9):1082–1097, 2009.

[3] Santo Fortunato. Community detection in graphs.Physics Reports, 486(3):75–174, 2010.

[4] Stuart A Rice. The identification of blocs in small po-litical bodies. The American Political Science Review,21(3):619–627, 1927.

[5] Wayne W Zachary. An information flow model for conflictand fission in small groups. Journal of anthropologicalresearch, pages 452–473, 1977.

[6] David Lusseau and Mark EJ Newman. Identifying therole that animals play in their social networks. Proceed-ings of the Royal Society of London. Series B: BiologicalSciences, 271(Suppl 6):S477–S481, 2004.

[7] Pall F Jonsson, Tamara Cavanna, Daniel Zicha, andPaul A Bates. Cluster analysis of networks generatedthrough homology: automatic identification of importantprotein communities involved in cancer metastasis. BMCbioinformatics, 7(1):2, 2006.

[8] Stuart L Pimm. The structure of food webs. TheoreticalPopulation Biology, 16(2):144–158, 1979.

[9] Ann E Krause, Kenneth A Frank, Doran M Mason,Robert E Ulanowicz, and William W Taylor. Compart-ments revealed in food-web structure. Nature, 426:282–285, 2003.

[10] Roger Guimera and Luis A Nunes Amaral. Functionalcartography of complex metabolic networks. Nature,433(7028):895–900, 2005.

[11] Petter Holme, Mikael Huss, and Hawoong Jeong. Sub-network hierarchies of biochemical pathways. Bioinfor-matics, 19(4):532–538, 2003.

[12] Gary William Flake, Steve Lawrence, C Lee Giles, andFrans M Coetzee. Self-organization and identification ofweb communities. Computer, 35(3):66–70, 2002.

[13] L. Gauvin, A. Panisson, and C. Cattuto. Detecting thecommunity structure and activity patterns of temporalnetworks: a non-negative tensor factorization approach.arXiv:1308.0723, August 2013.

[14] Michele Allegra and Paolo Giorda. Topology and energytransport in networks of interacting photosynthetic com-plexes. Phys. Rev. E, 85:051917, May 2012.

[15] Martin B Plenio and Susana F Huelga. Dephasing-assisted transport: quantum networks and biomolecules.New Journal of Physics, 10(11):113019, 2008.

[16] J. Adolphs and T. Renger. How proteins trigger excita-tion energy transfer in the FMO complex of green sulfurbacteria. Biophys. J., 91(8):2778–2797, 2006.

[17] J. C. Baez and J. Biamonte. A Course on QuantumTechniques for Stochastic Mechanics. arXiv:1209.3632,September 2012.

[18] Mauro Faccin, Tomi Johnson, Jacob Biamonte, SabreKais, and Piotr Migda l. Degree distribution in quantumwalks on complex networks. Phys. Rev. X, 3:041007, Oct2013.

[19] Zoltan Zimboras, Mauro Faccin, Zoltan Kadar, JamesWhitfield, Ben Lanyon, and Jacob Biamonte. Quantumtransport enhancement by time-reversal symmetry break-

ing. Sci. Rep., 3:2361, 2013.[20] Tomi H Johnson, Stephen R Clark, and Dieter Jaksch.

What is a quantum simulator? arXiv:1405.2831, 2014.[21] AK Ringsmuth, GJ Milburn, and TM Stace. Multiscale

photosynthetic and biomimetic excitation energy trans-fer. Nature Physics, 8(7):562–567, 2012.

[22] Mohan Sarovar, Akihito Ishizaki, Graham R. Fleming,and K. Birgitta Whaley. Quantum entanglement in pho-tosynthetic light-harvesting complexes. Nature Physics,6(6):462–467, April 2010.

[23] A. Ishizaki and G. R. Fleming. Quantum coherence inphotosynthetic light harvesting. Annu. Rev. Cond. Mat.Phys., 3(1):333–361, 2012.

[24] Yuan-Chung Cheng and Graham R. Fleming. Dynam-ics of light harvesting in photosynthesis. Annual Re-view of Physical Chemistry, 60(1):241–262, 2009. PMID:18999996.

[25] F. Caruso, A. W. Chin, A. Datta, S. F. Huelga, andM. B. Plenio. Highly efficient energy excitation trans-fer in light-harvesting complexes: The fundamental roleof noise-assisted transport. The Journal of ChemicalPhysics, 131(10):105106, 2009.

[26] M. Mohseni, P. Rebentrost, S. Lloyd, and A. Aspuru-Guzik. Environment-assisted quantum walks inphotosynthetic energy transfer. J. Chem. Phys.,129(17):174106, 2008.

[27] Leonardo Banchi, Gianluca Costagliola, Akihito Ishizaki,and Paolo Giorda. An analytical continuation approachfor evaluating emissionlineshapes of molecular aggregatesand the adequacy of multichromophoricforster theory.The Journal of Chemical Physics, 138(18):–, 2013.

[28] Daniel S Rokhsar and BG Kotliar. Gutzwiller projectionfor bosons. Physical Review B, 44(18):10328, 1991.

[29] Werner Krauth, Michel Caffarel, and Jean-PhilippeBouchaud. Gutzwiller wave function for a modelof strongly interacting bosons. Physical Review B,45(6):3137, 1992.

[30] K Sheshadri, HR Krishnamurthy, Rahul Pandit, andTV Ramakrishnan. Superfluid and insulating phases inan interacting-boson model: mean-field theory and therpa. EPL (Europhysics Letters), 22(4):257, 1993.

[31] Dirk-Soren Luhmann. Cluster gutzwiller method forbosonic lattice systems. Phys. Rev. A, 87:043619, Apr2013.

[32] Frank Verstraete, Valentin Murg, and J Ignacio Cirac.Matrix product states, projected entangled pair states,and variational renormalization group methods for quan-tum spin systems. Advances in Physics, 57(2):143–224,2008.

[33] J Ignacio Cirac and Frank Verstraete. Renormaliza-tion and tensor product states in spin chains and lat-tices. Journal of Physics A: Mathematical and Theoreti-cal, 42(50):504004, 2009.

[34] Haobin Wang, Ivan Pshenichnyuk, Rainer Hartle, andMichael Thoss. Numerically exact, time-dependent treat-ment of vibrationally coupled electron transport in single-molecule junctions. The Journal of Chemical Physics,135(24), 2011.

[35] Lushuai Cao, Sven Kronke, Oriol Vendrell, and PeterSchmelcher. The multi-layer multi-configuration time-dependent hartree method for bosons: Theory, imple-

Page 11: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

11

mentation, and applications. The Journal of ChemicalPhysics, 139(13), 2013.

[36] Xiaowei Pan, Zhenfeng Liu, Mei Li, and Wenrui Chang.Architecture and function of plant light-harvesting com-plexes II. Current Opinion in Structural Biology,23(4):515 – 525, 2013.

[37] Vladimir I. Novoderezhkin, Miguel A. Palacios, Herbertvan Amerongen, and Rienk van Grondelle. Excitationdynamics in the LHCII complex of higher plants: Mod-eling based on the 2.72 a crystal structure. The Jour-nal of Physical Chemistry B, 109(20):10493–10504, 2005.PMID: 16852271.

[38] Gabriela S. Schlau-Cohen, Tessa R. Calhoun, Naomi S.Ginsberg, Elizabeth L. Read, Matteo Ballottari, RobertoBassi, Rienk van Grondelle, and Graham R. Fleming.Pathways of energy flow in LHCII from two-dimensionalelectronic spectroscopy. The Journal of Physical Chem-istry B, 113(46):15352–15363, 2009. PMID: 19856954.

[39] Gunnar Carlsson and Facundo Memoli. Characterization,stability and convergence of hierarchical clustering meth-ods. The Journal of Machine Learning Research, 99:1425–1470, 2010.

[40] Roger Guimera, Marta Sales-Pardo, and Luıs A. NunesAmaral. Modularity from fluctuations in random graphsand complex networks. Phys. Rev. E, 70:025101, Aug2004.

[41] Benjamin H Good, Yves-Alexandre de Montjoye, andAaron Clauset. Performance of modularity maximizationin practical contexts. Physical Review E, 81(4):046106,2010.

[42] Andrea Lancichinetti, Filippo Radicchi, Jose J Ramasco,and Santo Fortunato. Finding statistically significantcommunities in networks. PloS ONE, 6(4):e18961, 2011.

[43] Ernesto Estrada. Community detection based on networkcommunicability. Chaos: An Interdisciplinary Journal ofNonlinear Science, 21, 2011.

[44] Ernesto Estrada and Naomichi Hatano. Communicabilitygraph and community structures in complex networks.Applied Mathematics and Computation, 214(2):500 – 511,2009.

[45] Marina Meila and Jianbo Shi. A random walks view ofspectral segmentation. In Eighth International Workshopon Artificial Intelligence and Statistics, 2001.

[46] J.-C. Delvenne, S. N. Yaliraki, and M. Barahona. Sta-bility of graph communities across time scales. Proceed-ings of the National Academy of Sciences, 107(29):12755–12760, 2010.

[47] Martin Rosvall and Carl T Bergstrom. Multilevel com-pression of random walks on networks reveals hierarchi-cal organization in large integrated systems. PloS one,6(4):e18209, 2011.

[48] Pascal Pons and Matthieu Latapy. Computing commu-nities in large networks using random walks. In Com-puter and Information Sciences-ISCIS 2005, pages 284–293. Springer, 2005.

[49] Trevor Hastie, Robert Tibshirani, and Jerome Friedman.The elements of statistical learning theory. Springer NewYork:, 2001.

[50] M. E. J. Newman and M. Girvan. Finding and evalu-ating community structure in networks. Phys. Rev. E,69:026113, Feb 2004.

[51] M. E. J. Newman. Fast algorithm for detecting commu-nity structure in networks. Phys. Rev. E, 69:066133, Jun2004.

[52] Aaron Clauset, M. E. J. Newman, and Cristopher Moore.Finding community structure in very large networks.Phys. Rev. E, 70:066111, Dec 2004.

[53] Santo Fortunato and Marc Barthelemy. Resolution limitin community detection. Proceedings of the NationalAcademy of Sciences, 104(1):36–41, 2007.

[54] Raj Rao Nadakuditi and Mark EJ Newman. Graph spec-tra and the detectability of community structure in net-works. Physical review letters, 108(18):188701, 2012.

[55] Filippo Radicchi. Detectability of communities in hetero-geneous networks. Phys. Rev. E, 88:010801, Jul 2013.

[56] Filippo Radicchi. A paradox in community detection.EPL (Europhysics Letters), 106(3):38001, 2014.

[57] L.N.F. Ana and A.K. Jain. Robust data clustering. InComputer Vision and Pattern Recognition, 2003. Pro-ceedings. 2003 IEEE Computer Society Conference on,volume 2, pages II–128–II–133 vol.2, 2003.

[58] Alexander Strehl and Joydeep Ghosh. Clusterensembles—a knowledge reuse framework for combiningmultiple partitions. The Journal of Machine LearningResearch, 3:583–617, 2003.

[59] Leon Danon, Albert Dıaz-Guilera, Jordi Duch, and AlexArenas. Comparing community structure identification.Journal of Statistical Mechanics: Theory and Experiment,2005(09):P09008, 2005.

[60] DaWei Lu, Jacob D. Biamonte, Jun Li, Hang Li, Tomi H.Johnson, Ville Bergholm, Mauro Faccin, Zoltan Zim-boras, Raymond Laflamme, Jonathan Baugh, and SethLloyd. Chiral quantum walks. arXiv:1405.6209, 2014.

[61] Kasper Astrup Eriksen, Ingve Simonsen, Sergei Maslov,and Kim Sneppen. Modularity and extreme edges of theinternet. Physical review letters, 90(14):148701, 2003.

[62] E Weinan, Tiejun Li, and Eric Vanden-Eijnden. Op-timal partition and effective dynamics of complex net-works. Proceedings of the National Academy of Sciences,105(23):7907–7912, 2008.

[63] Chris Godsil. Average mixing of continuous quantumwalks. arXiv:1103.2578 [math.CO], 2011.

[64] V. A. Traag and Jeroen Bruggeman. Community detec-tion in networks with positive and negative links. Phys.Rev. E, 80:036115, 2009.

[65] Andrea Lancichinetti, Santo Fortunato, and FilippoRadicchi. Benchmark graphs for testing community de-tection algorithms. Phys. Rev. E, 78:046110, Oct 2008.

[66] Elad Harel and Gregory S. Engel. Quantum coherencespectroscopy reveals complex dynamics in bacterial light-harvesting complex 2 (lh2). Proceedings of the NationalAcademy of Sciences, 109(3):706–711, 2012.

[67] Mark Newman. Networks: an introduction. Oxford Uni-versity Press, 2009.

[68] Eduardo Sanchez-Burillo, Jordi Duch, Jesus Gomez-Gardenes, and David Zueco. Quantum navigation andranking in complex networks. Scientific reports, 2, 2012.

[69] Anil K Jain and Richard C Dubes. Algorithms for clus-tering data. Prentice-Hall, Inc., 1988.

Page 12: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

12

Appendix A: Definitions

1. Modularity

Assume we have a directed, weighted graph (with possibly negative weights) and self-links, described by a realadjacency matrix A. The element Aij is the weight of the link from node i to node j.

The in- and outdegrees of node i are defined as

kini =∑

j

Aji, kouti =∑

j

Aij . (A1)

For a symmetric graph A is symmetric and the indegree is equal to the outdegree. The total connection weight ism =

∑i k

ini =

∑i k

outi =

∑ij Aij .

The community matrix C defines the membership of the nodes in different communities. The element CiA is equal tounity if i ∈ A, otherwise zero.7 The size of a community is given by |A| =

∑iCiA. For strict (non-fuzzy) communities

we can define C using an assignment vector σ (the entries being the communities of each node): CiA = δA,σi. This

yields (CCT )ij = δσi,σj.

There are many different ways of partitioning a graph into communities. A simple approach is to minimize thefrustration of the partition, defined as the sum of the absolute weight of positive links between communities andnegative links within them:

F = −∑

ij

Aijδσi,σj= − tr

(CTAC

). (A2)

Frustration is inadequate as a goodness measure for partitioning nonnegative graphs (in which a single communitycontaining all the nodes minimizes it). For nonnegative graphs we can instead maximize another measure calledmodularity:

Q =1

m

A,ij

(Aij − pij)CiACjA =1

mtr(CT (A− p)C

), (A3)

where pij is the “expected” link weight from i to j, with∑

ij pij = m, and is what separates modularity fromplain frustration. Different choices of the “null model” p give different modularities. Using degrees, we can definepij = kouti kinj /m.

For graphs with both positive and negative weights the usual definitions of degrees do not make much sense, sinceusually negative and positive links should not simply cancel each other out. Also, plain modularity will fail e.g.when m = 0. This can be solved by treating positive and negative links separately [64].

2. Hierarchical clustering

All our community detection approaches share a common theme. For each (proposed) community A we have a good-ness measure MA(t) that depends on the system Hamiltonian, the initial state, and t. This induces a correspondingmeasure for a partition X :

MX(t) =∑

A∈X

MA(t). (A4)

Using this, we define a function for comparing two partitions, X and X ′, which only differ in a single merge thatcombines A and B:

MA,B(t) = MX′(t) −MX(t) = MA∪B(t) −MA(t) −MB(t). (A5)

We can make MA,B(t) into a symmetric closeness measure c(A,B) by fixing the time t and normalizing it with |A||B|.Using this closeness measure together with the agglomerative hierarchical clustering algorithm (as explained in Sec. I)

7 For a fuzzy definition of membership we could require CiA ≥ 0 and∑

A CiA = 1 instead.

Page 13: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

13

we then obtain a community hierarchy. The goodness of a specific partition in the hierarchy is given by its modularity,obtained using the adjacency matrix given by Aij = c(i, j).

The standard hierarchical clustering algorithm requires closeness to fulfill the monotonicity property

min(c(A, C), c(B, C)) ≤ c(A ∪ B, C) ≤ max(c(A, C), c(B, C)). (A6)

for any communities A,B, C. If this does not hold, we may encounter a situation where the merging closeness sometimesincreases, which in turn means that the results cannot be presented as a dendrogram indexed by decreasing closeness.The real downside of not having the monotonicity property, however, is stability-related. The clustering algorithmshould be stable, i.e. a small change in the system should not dramatically change the resulting hierarchy. Assumewe encounter a situation where all the pairwise closenesses between a subset of clusters S = {Ai}i are within a giventolerance. A small perturbation can now change the pair {A,B} chosen for the merge. If Eq. (A6) is fulfilled, thenthe rest of S is merged into the same new cluster during subsequent rounds, and hence their relative merging orderdoes not matter.

3. Notation

Let the Hamiltonian of the system have the spectral decomposition H =∑

k EkΛk. The unitary propagator of thesystem decomposes as U(t) = e−iHt =

∑k e

−iEktΛk. We denote the state of the system at time t by

ρ(t) = U(t)ρ(0)U(t)†. (A7)

Sometimes we make use of the state obtained by measuring in which community subspace VA the quantum state islocated, and then discarding the result. The resulting state is

ρX(t) =∑

A∈X

ΠAρ(t)ΠA. (A8)

This state is normally not pure even if ρ(t) is.The probability of transport from node b to node a, the transfer matrix, is given by the elements

Rab(t) = |〈a|U(t)|b〉|2. (A9)

R(t) is doubly stochastic, i.e. its rows and columns all sum up to unity. We use R = (R + RT )/2 to denote itssymmetrization.

The time average of a function f(t) is denoted using f(t):

f(t) =1

t

∫ t

0

f(t′) dt′. (A10)

Now we have

Rab(t) =∑

jk

1

t

∫ t

0

e−i(Ej−Ek)t′

dt′〈a|Λj |b〉〈b|Λk|a〉. (A11)

The tH ≪ 1 and t→ ∞ limits of this average are

Rab(t→ 0) = δab

(1 − t2

3(H2)aa

)+t2

3|Hab|2 +O(t3),

Rab(t → ∞) =∑

jk

δjk〈a|Λj |b〉〈b|Λk|a〉 =∑

k

|〈a|Λk|b〉|2. (A12)

The time average of the state of the system is given by

ρ(t) =∑

jk

1

t

∫ t

0

e−i(Ej−Ek)t′

dt′Λjρ(0)Λk. (A13)

It can be interpreted as the density matrix of a system that has evolved for a random time, sampled from the uniformdistribution on the interval [0, t]. Again, in the short- and infinite-time limits this yields

ρ(t→ 0) =ρ(0) − it

2[H, ρ(0)] +

t2

3

(Hρ(0)H − 1

2

{H2, ρ(0)

})+O(t3),

ρ(t→ ∞) =∑

k

Λkρ(0)Λk. (A14)

Page 14: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

14

Appendix B: Closeness measures

1. Inter-community transport

Considering the flow of probability during a continuous-time quantum walk, let us investigate the change in theprobability of observing the walker within a community:

TA(t) =1

2|pA {ρ(t)} − pA {ρ(0)}| , (B1)

where pA {ρ} = tr (ΠAρ) is the probability of a walker in state ρ being found in community A upon a von Neumann-type measurement.8 A good partition should intuitively minimize this change, keeping the walkers as localized tothe communities as possible. TX =

∑A∈X TA is of course minimized by the trivial choice of a single community,

X = {A}, and any merging of communities can only decrease TX . Therefore we have TA∪B(t) ≤ TA(t) + TB(t).

The initial state ρ(0) can be chosen freely. For a pure initial state ρ(0) = |ψ〉〈ψ| we obtain

TA(t) =1

2

∣∣⟨ψ∣∣U †(t)ΠAU(t)

∣∣ψ⟩− 〈ψ |ΠA|ψ〉

∣∣ . (B2)

The change in inter-community transport is clearest when the process begins either entirely inside or entirely outsideeach community. Because of this, we choose the walker to be initially localized at a single node ρ(0) = |b〉〈b| and then,for symmetry, sum (or average) TA(t) over all b ∈ N :

TA(t) =1

2

b

∣∣⟨b∣∣U(t)†ΠAU(t)

∣∣ b⟩− 〈b |ΠA| b〉

∣∣

=1

2

b

∣∣∣∣∣∑

a∈A(Rab(t) − δab)

∣∣∣∣∣

=1

2

(∑

b∈A

∣∣∣∣∣1 −∑

a∈ARab(t)

∣∣∣∣∣+∑

b/∈A

∣∣∣∣∣∑

a∈ARab(t)

∣∣∣∣∣

)

=1

2

a/∈A,b∈ARab(t) +

a∈A,b/∈ARab(t)

=∑

a∈A,b/∈A

Rab(t) +Rba(t)

2=

a∈A,b/∈ARab(t), (B3)

since R(t) is doubly stochastic. Now we have

TA,B(t) = TA(t) + TB(t) − TA∪B(t) = 2∑

a∈A,b∈BRab(t) (B4)

with 0 ≤ TA,B(t) ≤ 2 min(|A|, |B|). The short- and long-time limits of the time-averaged TA,B(t) can be found usingEqs. (A12):

T t→0A,B = 2

a∈A,b∈B

(δab +

t2

3

(|Hab|2 − δab(H

2)aa)

+O(t3)

), (B5)

T t→∞A,B = 2

a∈A,b∈B

k

|(Λk)ab|2. (B6)

8 Equivalently, pA {ρ} is the norm of the projection (performed by projector ΠA) of the state ρ onto the community subspace VA.

Page 15: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

15

2. Intra-community fidelity

Our next measure aims to maximize the “similarity” between the evolved and initial states when projected to acommunity subspace. We do this using the squared fidelity

FA(t) = F 2 {ΠAρ(t)ΠA,ΠAρ(0)ΠA} , (B7)

where ΠAρΠA is the projection of the state ρ onto the subspace VA and

F {ρ, σ} = tr

{√√ρσ

√ρ

}∈ [0,

√tr{ρ} tr{σ}], (B8)

is the fidelity, which is symmetric between ρ and σ. If either ρ or σ is rank-1, their fidelity reduces to F {ρ, σ} =√tr{ρσ}. Thus, if the initial state ρ(0) is pure, we have

FA(t) = tr (ΠAρ(t)ΠAρ(0)) . (B9)

This assumption makes FX(t) equivalent to the squared fidelity between ρX(t) and a pure ρ(0):

FX(t) =∑

A∈X

tr (ΠAρ(t)ΠAρ(0)) = tr (ρX(t)ρ(0))

= F 2{ρX(t), ρ(0)} = F 2{ρ(t), ρX(0)}, (B10)

and yields

FA,B(t) = FA∪B(t) − FA(t) − FB(t)

= 2 Re tr (ΠAρ(t)ΠBρ(0))

= 2∑

a∈A,b∈BRe (ρab(t)ρba(0)) . (B11)

We use as the initial state the uniform superposition of all the basis states with arbitrary phases: |ψ〉 = 1√n

∑k e

iθk |k〉,which gives

FA,B(t) =2

n2

a∈A,b∈B

xy

Re(ei(θx−θy+θb−θa)UaxUby

). (B12)

In this case the short-term limit does not yield anything interesting. The long-time limit of the time-average of FA,B(t)is

F t→∞A,B =

2

n2

a∈A,b∈B

xy,k

Re(ei(θx−θy+θb−θa)(Λk)ax(Λk)yb

).

We may now (somewhat arbitrarily) choose all the phases θk to be the same, or average the closeness measure overall possible phases θk ∈ [0, 2π].

3. Purity

The coherence between any communities X = {A,B, . . .} is completely destroyed by measuring in which communitysubspace VA the quantum state is located, see Eq. (A8). If the measurement outcome is not revealed, the purity ofthe measured state ρX(t) is, due to the orthogonality of the projectors,

PX(t) = tr(ρ2X(t)

)=∑

A∈X

tr((ΠAρ(t))2

)=∑

A∈X

PA(t),

where

PA(t) = tr((ΠAρ(t)ΠA)2

)= tr

((ΠAρ(t))2

). (B13)

Page 16: Community Detectionin Quantum Complex Networksmodularity functions: the first is the coherent transport between communities and the second is the change in the states of individual

16

method initial state limit Aab

T |j〉, summed over before t-average 12

(

|Uab(t)|2 + |Uba(t)|2

)

t→ 0 δab + t2

3

(

|Hab|2 − δab(H

2)aa)

+O(t3)

t→ ∞∑

k|(Λk)ab|

2

F∑

j|j〉 before t-average

xyRe

(

ei(θx−θy+θb−θa)Uax(t)Uby(t))

t→ ∞∑

x,y,kRe ((Λk)ax(Λk)yb)

Fph∑

jeiθj |j〉, before t-average Re

(

Uaa(t)Ubb(t))

+ δab(

1 − |Uaa(t)|2)

phase-averaged t→ ∞∑

k(Λk)aa(Λk)bb + δab

(

1 −∑

k((Λk)aa)2

)

P∑

j |j〉 before t-average |∑

xy ei(θx−θy)Uax(t)Uby(t)|2

t→ ∞∣

k,x,y(Λk)ax(Λk)yb

2

+∑

k 6=m

x,y(Λk)ax(Λm)yb

2

Pph ∑

jeiθj |j〉, before t-average 1 + δab −

x|Uax(t)|2|Ubx(t)|2

phase-averaged t→ ∞ 1 + δab −∑

x

(

km|(Λk)ax|

2|(Λm)bx|2 +

k 6=m(Λk)ax(Λk)xb(Λm)bx(Λm)xa

)

Table I: Adjacency matrices, based on time-averaged measures. The closeness measure in each case is c(A,B) =2

n2|A||B|

a∈A,b∈B Aab. Note that if H has purely nondegenerate eigenvalues, then all the projectors are of the form

Λk = |ψk〉〈ψk|, which makes some of the t → ∞ measures above identical. For example T∞ becomes the same as F∞,ph

outside the diagonal. This type of nondegeneracy occurs e.g. when a small random perturbation is used to break the symme-tries of H .

If ρ(t) is pure, we have (cf. Eq. (B10))

PX(t) =∑

A∈X

tr(ΠAρ(t)ΠAρ(t)) = F 2{ρX(t), ρ(t)}. (B14)

The change in purity of the state after a projective measurement locating the walker into one of the communities is

PA,B(t) = PA∪B(t) − PA(t) − PB(t)

= 2 tr (ΠAρ(t)ΠBρ(t))

= 2∑

a∈A,b∈B|ρab(t)|2 ≥ 0. (B15)

Again, we will use the initial state |ψ〉 = 1√n

∑k e

iθk |k〉:

PA,B(t) =2

n2

a∈A,b∈B

∣∣∣∣∣∑

xy

ei(θx−θy)Uax(t)Uby(t)

∣∣∣∣∣

2

. (B16)

As with the fidelity-based measure, the short-time limit is uninteresting. The long-time limit of the time-average ofPA,B(t) is

P t→∞A,B = 2

a∈A,b∈B

| 〈a |ρ(∞)| b〉 |2 +

k 6=m

| 〈a |Λkρ0Λm| b〉 |2

= 2∑

a∈A,b∈B

|∑

kxy

ei(θx−θy)(Λk)ax(Λk)yb|2 +∑

k 6=m

|∑

xy

ei(θx−θy)(Λk)ax(Λm)yb|2 . (B17)