Startseite Mathematik What is a proper graph Laplacian? An operator-theoretic framework for graph diffusion
Artikel Open Access

What is a proper graph Laplacian? An operator-theoretic framework for graph diffusion

  • Ernesto Estrada EMAIL logo
Veröffentlicht/Copyright: 15. Dezember 2025

Abstract

We introduce an operator-theoretic definition of a proper graph Laplacian as any matrix associated with a given graph that can be expressed as the composition of a divergence and a gradient operator, with the gradient acting between graph-related spaces and annihilating constant functions. This provides a unified framework for determining whether a matrix represents a genuine diffusive operator on a graph. Within this framework, we prove that the standard Laplacian, the fractional Laplacian, the d -path Laplacians, and the degree-attracting and degree-repelling Laplacians are all proper diffusive Laplacians. In contrast, the in-degree and out-degree Laplacians correspond to advection operators, while the signed, signless, magnetic, and deformed Laplacians are improper, as they cannot be written as the composition of a divergence and a true gradient. The magnetic Laplacian is shown to arise as the Schur complement of an extended proper Laplacian defined on a higher-dimensional space, a property also inherited by the signless Laplacian. The Lerman-Ghosh Laplacian is identified as a nonconservative diffusive operator coupled to an external reservoir. Finally, we prove that the Moore-Penrose pseudoinverse of the Laplacian is itself a proper Laplacian. This classification establishes a rigorous operator-theoretic foundation for distinguishing proper, nonconservative, and improper graph Laplacians.

MSC 2020: 05C50; 15A09; 47A05; 35R02

1 Introduction

The standard graph Laplacian L [46,62] is a cornerstone of algebraic graph theory [25,47,48,68,84]. It plays a crucial role in understanding the structural properties of graphs and in modeling a wide range of dynamical processes on graphs and networks [12,61,76]. A classic example is the diffusion equation on an undirected graph G = ( V , E ) [32]:

(1) t x ( t ) + L x ( t ) = 0 ,

with initial condition x ( 0 ) = x 0 , where x ( t ) = [ x 1 ( t ) , , x n ( t ) ] T is a column vector of concentrations. The Laplacian matrix is defined as L = K A , where K is the diagonal matrix of vertex degrees and A is the adjacency matrix of the graph.

Let C ( V ) denote the set of all complex-valued functions on V , and let 2 ( V ) be the Hilbert space of square-summable functions on V . Then, L acts as a mapping from C ( V ) into itself

(2) L f ( v ) = ( v , w ) E ( f ( w ) f ( v ) ) , f C ( V ) .

Beyond this standard definition, many alternative forms of graph Laplacians have been proposed in the literature. For directed graphs, two natural generalizations exist: the in-degree Laplacian L in = K in A and the out-degree Laplacian L out = K out A , where the diagonal matrices K in and K out contain the in- and out-degrees of the vertices, respectively. A non-exhaustive list of Laplacian variants includes the deformed Laplacian [38,71], the d -path Laplacian L d [30,33,34], the fractional Laplacian L α [64,78], the hub-biased Laplacians L rep and L attr [31,43], the Hermitian (magnetic) Laplacian L B [11,26,29], the signless Laplacian Q [2123], the Laplacian for signed graphs L s [5,57], and the Lerman-Ghosh Laplacian L χ [45,58].

Further extensions have been introduced through the Euclidean distance matrix by Balaji and Bapat [8], as well as by considering matrix functions of the standard Laplacian [72,79]. More recently, Saxena et al. [80] proposed a diffusion-like model inspired by electrical network analysis, where an impedance network with shunt inductors connects each vertex to the ground. In matrix-vector form, the dynamics of this model can be written as

(3) L x ˙ ( t ) = x ( t ) ,

and by applying the Moore-Penrose pseudoinverse L + to both sides, one obtains

(4) t x ( t ) + L + x ( t ) = 0 ,

in close analogy with equation (1).

The question of what qualifies as a graph Laplacian has been addressed differently by various authors. Balaji and Bapat [8] proposed that a Laplacian matrix should be positive semidefinite, have zero row-sum, and possess rank n 1 , where n is the number of vertices. In contrast, Perez Riascos et al. [72,79] and, more recently, Devriendt [27] required that a Laplacian must also have nonpositive off-diagonal entries. Devriendt further added that the matrix should be irreducible, a condition automatically satisfied for connected graphs and thus not discussed further here.

The standard Laplacian L satisfies all these requirements, as do L d and L α . However, several other variants do not: L in and L out may not have rank n 1 ; L B is Hermitian but not always positive semidefinite, and it lacks both the zero row-sum and rank properties; the signless Laplacian Q neither has zero row-sum nor nonpositive off-diagonal entries; the Lerman-Ghosh Laplacian fails all three conditions; and the pseudoinverse L + can have positive off-diagonal elements. Although Saxena et al. [80] regarded L + as a valid Laplacian operator within their diffusion framework – and it does satisfy the conditions proposed by Balaji and Bapat – it does not fulfill the stricter requirements imposed by Riascos et al. and Devriendt. Indeed, it has been explicitly stated that L + “is not always a Laplacian” because some of its off-diagonal elements can be positive (refer pp. 133–134 in [65]). The Laplacian pseudoinverse is an important matrix in the study of graphs [40,66], with applications in the definition of the resistance distance matrix of the graph [44,50,54,70,77,86,87], the calculation of the hitting/commuting times in Markov chains [14,18,75], as well as for analyzing the robustness of networked control systems [60,88,89]. Thus, the question about whether it is a proper Laplacian operator or not is very relevant across the disciplines. The objective of this work is therefore twofold: first, to provide a clear and pedagogical discussion of what should properly be regarded as a Laplacian operator for a graph, and second, to address the question of whether the pseudoinverse of the graph Laplacian, L + , can itself be considered a genuine Laplacian operator.

2 Preliminaries

We introduce the basic notations and operator spaces used throughout this study. Let C ( V ) denote the space of all complex-valued functions on the vertex set V , and C ( E ) (or more generally C ( Σ ) ) denote the space of edge-valued functions (or functions defined on subgraph elements). We endow 2 ( V ) and 2 ( E ) with their canonical inner products f , g V = v V f ( v ) ¯ g ( v ) and u , w E = e E u ( e ) ¯ w ( e ) .

The discrete gradient ϒ : C ( V ) C ( E ) acts as a difference operator between connected vertices, producing edgewise increments. Its adjoint ϒ * defines a canonical divergence operator, div 0 ϒ * : C ( E ) C ( V ) , which measures the net outflow from each vertex. A proper gradient satisfies ϒ ( 1 ) = 0 , ensuring that constant functions have zero gradient.

In general, a proper Laplacian is the composition L = div ϒ , where div need not coincide with ϒ * . The familiar combinatorial Laplacian L = ϒ * ϒ corresponds to the special case in which the divergence equals the adjoint of the gradient, div = ϒ * , thus yielding a self-adjoint diffusive operator. Other Laplacians, such as directed or magnetic variants, arise when div and ϒ * differ, breaking self-adjointness while preserving the diffusive structure.

Let us give here some indications about the notation used in the study. Throughout the study, we distinguish four levels of gradient notation:

  1. the abstract operator ϒ acting between function spaces,

  2. its matrix realization representing oriented vertex-edge incidence,

  3. specialized variants τ , σ , B , d , χ used for directed, signed, magnetic, distance-based, or extended gradients, and

  4. the roman symbols grad and div, which denote generic gradient/divergence operators in prose.

This hierarchy avoids ambiguity between abstract operator, matrix form, and contextual variants.

We collect a few standard facts that we will use repeatedly. Let μ 1 μ n be the set of eigenvalues of the graph Laplacian L ( G ) . We have then the following known results.

Lemma 1

(Monotonicity under edge addition) [19,39] If G = ( V , E ) is obtained from G = ( V , E ) by adding edges ( E E ), then L ( G ) L ( G ) is positive semidefinite. Consequently, μ k ( G ) μ k ( G ) for all k , and in particular μ n ( G ) μ n ( G ) .

Proof

Each added edge { i , j } contributes the rank-1 PSD Laplacian L { i , j } = ( e i e j ) ( e i e j ) . Summing such contributions yields L ( G ) L ( G ) 0 . The eigenvalue monotonicity follows from the Courant-Fischer characterization.□

Lemma 2

(Upper bound by maximum degree) [19,82] For any graph G , μ n ( G ) 2 Δ ( G ) , where Δ ( G ) is the maximum degree.

Proof

By Gershgorin disks for the Laplacian, every eigenvalue lies in [ 0 , 2 Δ ] . Equivalently, x L x = { i , j } E ( x i x j ) 2 2 Δ i x i 2 .□

Lemma 3

(Cheeger upper bound) [17,19] Let ϕ ( G ) min S V E ( S , S ¯ ) min { vol ( S ) , vol ( S ¯ ) } , where vol ( S ) = i S k i . Then, the (combinatorial) Laplacian satisfies μ 2 ( G ) 2 ϕ ( G ) .

Proof

This classical bound follows by taking the Fiedler Rayleigh quotient with a test function f that is approximately constant on S and on S ¯ with opposite signs, and optimizing over S (refer, e.g., standard Cheeger-type inequalities for the combinatorial Laplacian [4,19,82]).□

3 On Balaji-Bapat and Riascos et al./Devriendt properties of a Laplacian

Here we explain briefly what the properties proposed by Balaji and Bapat [8], Perez Riascos et al. [78,79], and Devriendt [27] mean from the perspective of an operator for diffusive dynamics on a graph. Let us discuss these properties individually.

3.1 Positive definiteness and rank equal to n 1

Let L be a candidate Laplacian operator for a graph. The requirement that L be positive semidefinite ensures that the diffusion dynamics

t x ( t ) + L x ( t ) = 0 , x ( 0 ) = x 0

converges to a steady state. Indeed, if 0 = μ 1 1 ( L ) = = μ 1 h ( L ) , i.e., there are h zero eigenvalues, and μ j > h > 0 , then

lim t x ( t ) = lim t e t L x 0 = j = 1 h ψ 1 j ( ψ 1 j T x 0 ) = κ ,

where ψ 1 j are the eigenvectors associated with the zero eigenvalues of L . The vector κ is time-independent, representing the steady state of the diffusion. In general, its components may differ, depending on the multiplicity of the zero eigenvalue.

If, in addition, rank ( L ) = n 1 , then the zero eigenvalue is simple, implying that all components of κ are identical. In this case, the diffusion converges to a unique steady state.

3.2 Zero row-sum

The zero row-sum property of L ensures that L 1 = 0 . Since we can write 1 = k = 1 c n k u k ker ( L ) for c 1 , with u k = 1 n k 1 C k ( k = 1 , , c ), we have

lim t x ( t ) = k = 1 c ( ψ 1 , k T x 0 ) ψ 1 , k = k = 1 c 1 n k i C k x i 0 1 n k 1 C k = k = 1 c 1 n k i C k x i 0 1 C k .

This means that the steady state of the diffusion dynamics on the graph consists of one or several “consensus” states among groups of vertices, each characterized by the average of the initial condition over those vertices. Clearly, if the rank of L is n 1 , there is a unique steady state, corresponding to the global average of the initial condition over all vertices, which justifies the name of “consensus” given to this dynamics in engineering contexts (refer for instance [74]).

3.3 Non-positive off-diagonal entries

Assuming that the diagonal entries of L are positive, the condition that off-diagonal entries are nonpositive ensures that when the operator is applied to a function defined on the vertices of a graph, it produces at each vertex the sum of the differences of the function values along the edges incident to that vertex. Indeed, let L = [ i j ] i , j = 1 n be a symmetric matrix satisfying

i i > 0 , i j 0 for i j , j i j = 0 .

Then, for any vector x C n ,

( L x ) i = j i ( i j ) ( x i x j ) .

Thus, the action of L on x at vertex i consists of adding up the weighted differences between the value of x at i and those at its neighboring vertices. The positive diagonal term i i = j i ( i j ) balances the outgoing and incoming contributions, ensuring that L 1 = 0 .

3.4 Additional condition for conservation of mass

A diffusive dynamics is said to be conservative if i = 1 n x i ( t ) = i = 1 n x i 0 for any t 0 , where the sum is carried out over all vertices of the graph. Then, if L is symmetric, the condition that it has zero row-sum is sufficient to guaranty that the diffusive process is conservative. If the matrix is not symmetric, then the conservation of mass is obtained if the matrix is also zero column-sum. That is, let us consider the sum of all the values of x ( t ) at a given t > 0

(5) 1 T x ( t ) = 1 T e t L x 0 = 1 T I + t ( L ) + t 2 ( L ) 2 2 ! + + t k ( L ) k k ! + x 0 = 1 T x 0 + t 1 T ( L ) x 0 + t 2 1 T ( L ) ( ( L ) x 0 ) + + t k 1 T ( L ) ( ( L ) k 1 x 0 ) + = 1 T x 0 ,

due to the fact that the condition 1 T ( L ) = 0 has been imposed.

4 General definition of graph Laplacians

The notion of a Laplacian on a graph should not be restricted to specific matrix sign patterns or spectral properties, but should instead be grounded on its functional construction as the composition of a divergence and a gradient operator. Accordingly, we propose the following general definition.

Definition 4

(Graph Laplacian). A matrix L C n × n associated with a graph G = ( V , E ) is said to be a Laplacian operator on G if there exist two linear operators

grad : C ( V ) C ( Σ ) , div : C ( Σ ) C ( V ) ,

acting between the space of complex-valued functions on vertices and that on a family of subgraph structures Σ of G (for instance, edges, paths, cycles, or higher-order motifs), such that

L = div grad , grad ( 1 ) = 0 .

Here Σ is a family of graph elements that connect groups of vertices, such as edges, paths, cycles, or any other subgraph structures of G . The gradient operator grad maps vertex functions into functions defined on Σ , thus quantifying the local variations in the field between connected graph elements, while div aggregates those variations back onto the vertices.

Remark 5

The structural definition above differs fundamentally from the usual algebraic characterizations of a Laplacian, such as those of Balaji-Bapat [8] and Perez Riascos et al. [79]/Devriendt [27]. Because L = div grad and grad ( 1 ) = 0 , every proper Laplacian automatically satisfies L 1 = 0 , i.e., it has zero row-sum, and its nullspace always contains the constants, so that rank ( L ) n 1 for a connected graph. No other matrix property of Balaji and Bapat [8] and Perez Riascos et al. [79]/Devriendt [27] is required. In particular, positive semidefiniteness, symmetry, and nonpositive off-diagonal entries are not implied by the operator form div grad . They arise only when one further demands div = grad * with respect to some inner product, in which case L = grad * grad becomes self-adjoint and positive semidefinite in that metric. Without this additional assumption, L may be nonsymmetric or even indefinite, as exemplified by the degree-biased Laplacians or other anisotropic diffusions. Finally, the requirement that grad acts on a family Σ of subgraph structures ensures that L is not an arbitrary zero-row-sum matrix, but one generated by local or higher-order interactions prescribed by the graph G .

Definition 6

(Graph gradient). The operator grad is a proper gradient if it satisfies the condition

grad ( 1 ) = 0 ,

where 1 is the constant function on V . This condition ensures that the gradient measures differences between values of the function across connected elements of the graph and vanishes for constant functions. If grad ( 1 ) 0 , the operator does not represent variations in the field and cannot be associated with a Laplacian of diffusive nature.

These conditions are well known to be obeyed by the standard graph Laplacian as can be seen in the following result, shown for completeness of the current work.

Proposition 7

The standard graph Laplacian L = K A for an undirected graph is a Laplacian operator in the above sense.

Proof

Let denote the oriented vertex-edge incidence matrix of the graph, defined by

e v = + 1 , v = head ( e ) , 1 , v = tail ( e ) , 0 , otherwise.

Define the discrete gradient as

grad x = x ,

so that ( grad x ) e = x head ( e ) x tail ( e ) , and its adjoint (negative divergence) as

div = .

Then,

L = div grad = .

By construction,

grad ( 1 ) = 1 = 0 ,

so the gradient of a constant function vanishes. Therefore, L = satisfies the defining property of a Laplacian operator.□

5 Are other “Laplacians” proper graph Laplacians?

Based on the previous definition, we now prove a series of results about several known “Laplacian” on graphs. Let us start by the out- and in-degree Laplacians. The out-degree Laplacian has been previously studied by Chapman [15,16] who effectively defines it as an “advection” operator. The in-degree Laplcian is commonly used as a “diffusive” (consensus) operator on directed graphs (refer Chapter 3.2 in [63]). Hereafter we clarify the conceptual meaning of these two operators.

Proposition 8

On a directed graph G = ( V , E ) with adjacency matrix A, the out- and in-degree Laplacians

L out K out A , L in K in A

are not proper Laplacians, i.e., they cannot be written as L = div grad with a diffusive gradient grad satisfying grad ( 1 ) = 0 . Instead, they admit transport-type factorizations:

L out = U τ , L in = div ρ V ,

where U and V are vertex-edge and edge-vertex “tail” selectors (transport fields), and τ , div ρ are oriented discrete derivatives along edges. Hence, L out acts as an advection operator (outflow) and L in as its inflow/accumulation conjugate.

Proof

(1) Out- and in-degree Laplacians are not diffusive. Any operator of the form L = div grad with grad ( 1 ) = 0 and div = grad * is Hermitian and positive semidefinite. For a genuinely directed graph ( A A ),

( L out ) * = K out A K out A = L out , ( L in ) * = K in A L in ,

so it is neither self-adjoint nor diffusive in the standard inner product. Thus, they are not proper Laplacians.

(2) Advection factorization of L out . Choose an orientation of edges and let be the vertex-edge incidence (difference) matrix,

( x ) e = x head ( e ) x tail ( e ) , 1 = 0 .

Let U be the vertex-edge “tail” selector,

U v e = 1 , v = tail ( e ) , 0 , otherwise.

Then, for any x C n ,

( U ( ) x ) v = e : tail ( e ) = v ( x tail ( e ) x head ( e ) ) = k out ( v ) x v w A v w x w = ( L out x ) v .

Hence,

L out = U τ , τ ,

which represents a discrete advection (outflow) operator with velocity field U .

(3) Inflow/accumulation factorization of L in . Define the discrete divergence div and the edge-vertex “tail” selector

V e v = 1 , v = tail ( e ) , 0 , otherwise.

Then, componentwise

( V x ) v = e : head ( e ) = v ( x tail ( e ) ) + e : tail ( e ) = v x tail ( e ) = u : u v x u + k out ( v ) x v .

Reversing the orientation (equivalently, using the incidence of the reversed digraph) yields

( rev V rev x ) v = k in ( v ) x v w A v w x w = ( L in x ) v ,

so that

L in = div ρ V ,

with div ρ the divergence in the reversed orientation. This operator aggregates incoming flux and thus acts as an inflow or accumulation operator – the natural conjugate of advection.□

Remark 9

The pair ( L out , L in ) realizes a discrete transport duality

u f (advection/outflow) vs ( u f ) (inflow/continuity) .

Neither operator is a proper Laplacian, since they lack the diffusive-conservative form div grad with grad ( 1 ) = 0 ; they instead describe directed transport processes.

Let us now consider the so-called deformed Laplacian in its undirected graph-version [38,71] which was used for studying “consensus” (diffusive) dynamics on graphs.

Proposition 10

Let G = ( V , E ) be a simple undirected graph with degree matrix K = diag ( k v ) , adjacency A, and identity I. For a real parameter r, define the deformed Laplacian

L ( r ) = ( K I ) r 2 + A r + I .

Then, except in two trivial regular cases, L ( r ) is neither diffusive (positive semidefinite) nor a proper Laplacian of the form L = div grad with a proper gradient satisfying grad ( 1 ) = 0 .

Proof

If L ( r ) was a proper Laplacian, the condition grad ( 1 ) = 0 would imply L ( r ) 1 = 0 . For each vertex v ,

( L ( r ) 1 ) v = ( k v 1 ) r 2 + k v r + 1 q v ( r ) .

For a fixed r , q v ( r ) = 0 for all v only if all degrees are equal ( G is k -regular), since q v depends on k v . In that case, the roots of ( k 1 ) r 2 + k r + 1 = 0 are r = 1 and r = 1 ( k 1 ) . Only for these two values do we have

L ( 1 ) = K A = L , L 1 k 1 = 1 k 1 L ,

which are (multiples of) the standard Laplacian and hence, proper in the strict diffusive-self-adjoint sense. For every other value of r , L ( r ) 1 0 , so L ( r ) cannot be expressed as div grad with grad ( 1 ) = 0 .

Regarding diffusivity, for general r , the off-diagonal entries of L ( r ) on edges equal + r , which become positive when r > 0 . Consequently, x L ( r ) x need not be nonnegative, and the matrix L ( r ) fails to be positive semidefinite except at the conservative points r = 1 and r = 1 ( k 1 ) for regular graphs. In general, L ( r ) is indefinite and therefore non-diffusive. Hence, except for the two regular conservative cases, L ( r ) is neither diffusive nor a proper Laplacian.□

Remark 11

Whenever the diagonal quantities

s v ( k v 1 ) r 2 + k v r + 1 ( 0 )

are nonnegative (e.g., for r 1 or for r [ 1 ( Δ 1 ) , 0 ] with Δ = max v k v ), one may write

L ( r ) = ( r ) + diag ( s v ) = r diag ( s v ) r diag ( s v ) .

Here is any oriented incidence matrix. This shows that L ( r ) is diffusive (positive semidefinite), but the added unary rows do not compare pairs of vertices, so the corresponding “gradient” does not annihilate constants. Therefore, even in this case, L ( r ) is not a proper Laplacian under the definition L = div grad with grad ( 1 ) = 0 .

We now consider one of the emergent lines of research for the study of directed and mixed graphs: the use of complex-valued Hermitian Laplacian matrices [1,13,20,42,49,51,59,69,91], which possess many valuable algebraic properties. Additionally, complex-valued Hermitian Laplacian matrices can be related to the magnetic Schrödinger operator [7,11,26,35,37,52,53,55,56,81], such that they are known as magnetic Laplacians in the literature.

Proposition 12

Let G = ( V , E ) be a directed or mixed graph with complex edge phases ω i j C satisfying ω i j = 1 and ω j i = ω i j ¯ whenever both orientations ( i , j ) and ( j , i ) are present. Define the magnetic (Hermitian) Laplacian

L B = K A B , ( A B ) i j = ω i j , ( i , j ) E , 0 , otherwise,

where K = diag ( k v ) is the degree matrix with k v = j ( A B ) v j . Then, L B is Hermitian and diffusive (it admits the factorization L B = B * B ), but it is not a proper Laplacian in the sense L = div grad with a proper gradient satisfying grad ( 1 ) = 0 , unless all phases are trivial ( ω i j 1 ). As a limiting real case, the signless Laplacian Q = K + A (corresponding to ω i j 1 ) is also diffusive but not proper.

Proof

Fix an orientation of E and define the magnetic gradient B : C ( V ) C ( E ) by

( B x ) ( i , j ) = x i ω i j x j .

A direct computation yields the Hermitian diffusive factorization

L B = B * B .

However, for the constant vector 1 , we have

( B 1 ) ( i , j ) = 1 ω i j ,

which is nonzero whenever ω i j 1 . Thus, B fails to be a proper gradient since it does not annihilate constants. Consequently, L B , while diffusive and self-adjoint, cannot be expressed as div grad with a proper gradient satisfying grad ( 1 ) = 0 unless all phases are trivial.

To exclude the possibility of another proper gradient ϒ generating L B , suppose L B = ϒ * ϒ = B * B . By the polar decomposition, there exists a partial isometry U such that B = U ϒ . Hence, B 1 = U ϒ 1 = 0 , contradicting the fact that B 1 0 . Therefore, no proper gradient can yield L B unless ω i j 1 , in which case B reduces to the ordinary incidence matrix and L B becomes the standard Laplacian.

For the uniform phase ω i j 1 , one has A B = A and thus, L B = K ( A ) = K + A Q . The associated incidence operator is

( + x ) ( i , j ) = x i + x j ,

for which + 1 0 . Repeating the same argument shows that no proper gradient ϒ with ϒ 1 = 0 can satisfy Q = ϒ * ϒ . Hence, Q is also diffusive but not proper.□

Remark 13

The magnetic Laplacian L B can be interpreted as the Schur complement of a proper Laplacian on an extended vertex-edge space. Given an oriented or mixed graph G = ( V , E ) and magnetic phases ω i j with ω i j = 1 , introduce auxiliary edge variables z e C and define the extended gradient

ϒ : C ( V ) × C ( E ) C 2 E , ϒ ( x , z ) has, for each e = ( i , j ) , two rows 2 ( x i z e ) , 2 ( ω i j x j z e ) .

In block form,

ϒ = M D , M = 2 e = ( i , j ) e i ω i j e j , D = I m I m .

The corresponding extended Laplacian

ext = ϒ * ϒ = M * M M * D D * M D * D

is Hermitian positive semidefinite and of the diffusive form grad * grad on the extended space V E . Eliminating the edge variables z gives the Schur complement

L B = ext ( D * D ) = M * M M * D ( D * D ) 1 D * M .

Thus, the magnetic Laplacian arises as the projection of a proper Laplacian defined on a higher-dimensional space. The signless Laplacian Q = K + A corresponds to the real case with a uniform magnetic phase ω i j = 1 (a phase shift of π ) along all oriented edges.

Example 14

Consider the undirected 3-cycle with edges { 1 , 2 } , { 2 , 3 } , { 3 , 1 } and magnetic phases ω 12 , ω 23 , ω 31 C satisfying ω i j = 1 and ω j i = ω i j ¯ . The magnetic Laplacian is

L B = K A B = 2 ω 12 ω 31 ¯ ω 12 ¯ 2 ω 23 ω 31 ω 32 ¯ 2 , K = 2 I 3 .

We introduce one auxiliary variable z i j for each oriented edge ( i j ) { ( 1 , 2 ) , ( 2 , 3 ) , ( 3 , 1 ) } . Define the extended magnetic gradient ϒ : C 3 × C 3 C 6 by

ϒ = 1 0 0 1 0 0 0 ω 12 0 1 0 0 0 1 0 0 1 0 0 0 ω 23 0 1 0 0 0 1 0 0 1 ω 31 0 0 0 0 1 , ϒ x 1 x 2 x 3 z 12 z 23 z 31 = x 1 z 12 ω 12 x 2 z 12 x 2 z 23 ω 23 x 3 z 23 x 3 z 31 ω 31 x 1 z 31 .

In matrix form, this can be written as

ϒ = [ M D ] ,

where

M = 1 0 0 0 ω 12 0 0 1 0 0 0 ω 23 0 0 1 ω 31 0 0 , D = 1 0 0 1 0 0 0 1 0 0 1 0 0 0 1 0 0 1 .

Hence,

D * D = 2 I 3 , ext = ϒ * ϒ = M * M M * D D * M D * D ,

which for the example in question is given by

ext = ϒ * ϒ = 2 ω 12 ω 13 ¯ 1 0 0 ω ¯ 12 2 ω 23 0 1 0 ω 31 ω 23 ¯ 2 0 0 1 1 0 0 2 0 0 0 1 0 0 2 0 0 0 1 0 0 2 .

Eliminating the edge variables z gives the Schur complement

ext ( D * D ) = M * M 1 2 M * D D * M = { i , j } E ( e i e i * + e j e j * ω i j ¯ e i e j * ω i j e j e i * ) = L B .

Therefore, the Hermitian magnetic Laplacian L B is precisely the Schur complement of the proper extended Laplacian ext = ϒ * ϒ . Note that in this particular example, due to the specific choice of orientation and normalization, L B is also the upper-left block of ext .

Remark 15

Let L B = K A B be the magnetic Laplacian as defined before, such that A B ( i , j ) = ω i j C if ( i , j ) E and ω j i = ω i j ¯ . Let ϕ = Re ( ω i j ) and let φ = Im ( ω i j ) . Let R be the matrix whose entries are defined as

R i j = 1 iff i j 1 iff i j 0 otherwise .

Then, we can write

L B = ϕ K ϕ A + φ i R ϕ L ϕ + φ i R .

Let us plug this operator on the diffusion equation:

t x ( t ) + ϕ L ϕ x ( t ) φ i R x ( t ) = 0 .

Then, L ϕ is a diffusive Laplacian with diffusion coefficient ϕ and i R can be considered as an imaginary reaction potential, which account for wave-like reactions. The equation is similar to the one proposed by Varshney et al. [85] to model a diffusive process plus a chemical potential in neuronal systems. The difference is that the model of Varshney et al. [85] does not include the term φ i multiplying the chemical potential, which makes that their operator is not symmetric, its eigenvalues may be complex, and it could be nondiagonalizable. Additionally, the semigroup obtained from their model is neither stochastic nor sub-Markovian.

We now turn our attention toward the Laplacian matrix of a graph with edge signs, known as signed graphs [90]. The study of the signed Laplacian has received some attention in the literature for the analysis of real-world systems in which friendship-enmity relations are present [5,57], particularly interesting is the concept of dissensus found in some graphs by Altafini [5].

Proposition 16

Let G σ = ( V , E , σ ) be a signed graph with edge signs σ i j { ± 1 } on { i , j } E , and let

L σ = D A σ , ( A σ ) i j = σ i j , { i , j } E , 0 , otherwise.

Then, L σ is Hermitian and diffusive (it admits the factorization L σ = σ * σ ), but it is not a proper Laplacian in the sense L = div grad with a proper gradient satisfying grad ( 1 ) = 0 , unless the signing is trivial ( σ i j + 1 on all edges).

Proof

Fix an orientation of E and define the signed (“twisted”) gradient σ : C ( V ) C ( E ) by

( σ x ) ( i , j ) = x i σ i j x j .

A direct computation gives the standard factorization

L σ = σ * σ .

Evaluating this operator on the constant vector 1 yields

( σ 1 ) ( i , j ) = 1 σ i j .

Hence, σ 1 = 0 if and only if σ i j = + 1 on every edge. Whenever a negative edge is present, σ fails to be a proper gradient, since it does not annihilate constants.

Suppose there existed another proper gradient ϒ with ϒ 1 = 0 such that L σ = ϒ * ϒ = σ * σ . Then, by the polar decomposition, there would exist a partial isometry U such that σ = U ϒ . Applying both sides 1 gives σ 1 = U ϒ 1 = 0 , a contradiction when any negative edge exists. Thus, L σ can only be proper when all σ i j = + 1 , in which case it reduces to the standard Laplacian.□

Remark 17

The signed Laplacian L σ = D A σ can be decomposed as the difference of two proper Laplacians,

L σ = L + L ,

where L + and L are the ordinary Laplacians of the subgraphs consisting of positive and negative edges, respectively. Hence, L σ represents the superposition of two diffusive processes of opposite sign: one promoting alignment across positive links, and the other promoting anti-alignment across negative ones. It is therefore not a purely diffusive operator but the generator of a competition between cooperative and antagonistic diffusion.

The so-called Lerman-Ghosh Laplacian L χ was introduced by Lerman and Ghosh in 2012 [58] as an attempt to explore the nonconservative diffusion processes occurring in online social media. It was then further studied in [32,45,67].

Proposition 18

Let G = ( V , E ) be a (simple) undirected graph with degree matrix K = diag ( k 1 , , k n ) and standard incidence (edge-vertex) matrix (so L = = K A ). Fix a parameter χ R and define the augmented (“hairy”) incidence

χ [ ] R n × ( m + n ) , diag ( max { χ k i , 0 } ) , 2 = diag ( max { χ k i , 0 } ) ,

interpreting each diagonal entry of as a semiedge connecting vertex i to an external reservoir. Then, the Lerman-Ghosh operator

L χ χ ( χ )

is self-adjoint and positive semidefinite (diffusive) and satisfies

L χ = L + ( χ I K ) .

However, L χ is not a proper Laplacian in the sense L = grad * grad with a proper gradient such that grad ( 1 ) = 0 .

Proof

Block multiplication gives

L χ = [ ] = + = L + 2 ,

where 2 = diag ( max { χ k i , 0 } ) . Since L χ is a Gram operator, it is symmetric and positive semidefinite:

x L χ x = χ x 2 0 , x R n .

However,

χ 1 = ( 1 ; 1 ) = ( 0 ; 1 ) 0 ,

whenever 0 , which implies that the augmented gradient χ does not annihilate constants. Therefore, L χ cannot be expressed as grad * grad with a proper gradient grad ( 1 ) = 0 . In this sense, the Lerman-Ghosh operator is diffusive but nonconservative, and not a proper Laplacian.□

Remark 19

If the system is extended by including an additional vertex representing the reservoir and defining the hair terms as true differences x i y , the corresponding augmented gradient

ext χ x = ( x ; ( x 1 y ) )

satisfies ext χ 1 = 0 . The associated extended operator

L χ ext = ext χ ( ext χ ) = L + 2 2 1 1 2 1 2 1

is then a proper conservative Laplacian on the enlarged graph ( V { y } , E { ( i , y ) } ) . Therefore, the Lerman-Ghosh operator is: L χ = L + 2 , which means it is precisely the principal submatrix of the extended conservative Laplacian obtained by removing the row and column corresponding to the reservoir node. Hence, L χ describes the open (nonconservative) subsystem of the proper diffusive operator L χ ext acting on the extended graph ( V { y } , E { ( i , y ) } ) .

Remark 20

It is important to distinguish between the properness of a Laplacian operator and the stability of the diffusion dynamics it generates. In the extended-graph construction of Remark 19, where an additional vertex is introduced to represent the reservoir and the coupling parameter is denoted by χ , the resulting operator L ext = div grad indeed satisfies grad ( 1 ) = 0 . Hence, it qualifies as a proper Laplacian in the structural sense adopted throughout this work.

However, properness alone does not guarantee that the diffusion

x ˙ = L ext x

converges to a finite steady state. If the coupling strength χ is smaller than the spectral radius ρ ( A ) of the adjacency matrix A of the original graph, the eigenvalues of L ext may acquire negative real parts, leading to an unbounded evolution in which some components of x ( t ) diverge to + and others to . The diffusion remains structurally proper but becomes dynamically unstable.

Stability is restored when χ ρ ( A ) , since in this regime, the spectrum of L ext is nonnegative and the dynamics converge to the consensus (or average) state. This observation underscores that properness – the existence of a valid div-grad factorization with grad ( 1 ) = 0 – is a geometric property of the operator, while stability is a dynamical property depending on the spectral location of its eigenvalues.

Remark 21

Setting χ = 0 gives L 0 = L K = A , so the (negative) adjacency operator appears as the χ -Laplacian with zero reservoir strength. Allowing complex hair entries or phases on edges turns χ into a complex (magnetic-type) incidence operator, as studied in related works on complex gradients.

The study of roots of the standard Laplacian matrix, a.k.a. fractional graph Laplacian has been developed by Riascos and collaborators [64,78] as a way to account for nonlocal spatial effects on diffusive processes on graphs.

Proposition 22

Let G = ( V , E ) be an undirected graph and let L = be its standard Laplacian (with any fixed orientation of E), acting on C ( V ) . For any α > 0 , define the fractional Laplacian by spectral calculus

L α = U Λ α U * ,

where L = U Λ U * is a unitary diagonalization with Λ = diag ( λ 1 , , λ n ) , λ k 0 and λ k = 0 for k indexing the nullspace (spanned by constants on each connected component). Then, L α is a proper graph Laplacian in the sense that there exists a (possibly nonlocal) gradient

grad α : C ( V ) C n , grad α ( 1 ) = 0 , L α = grad α * grad α .

Proof

Let L = U Λ U * with U unitary and Λ = diag ( λ 1 , , λ n ) , λ k 0 . Define

grad α Λ α 2 U * : C ( V ) C n .

Then,

grad α * grad α = U Λ α 2 Λ α 2 U * = U Λ α U * = L α ,

so L α is of the form grad * grad . Moreover, if 1 denotes the (componentwise) constant vector on V , then 1 lies in the nullspace of L (one copy per connected component), hence U * 1 has zero entries on all coordinates where λ k > 0 . Since Λ α 2 has zeros on those null coordinates, we obtain

grad α ( 1 ) = Λ α 2 U * 1 = 0 .

Thus, grad α is a proper gradient (it annihilates constants). Consequently, L α = grad α * grad α is self-adjoint, positive semidefinite, and conservative on V : L α 1 = 0 (indeed, L α acts as 0 on the nullspace of L by functional calculus).□

Remark 23

The operator grad α maps vertex functions to coefficients along the spectral (Fourier) modes of L , scaled by λ k α 2 . Hence, for α 1 , the induced Laplacian L α is nonlocal in the vertex domain, but remains a proper (diffusive-conservative) Laplacian under the general definition where the gradient may act between vertices and an abstract family of graph elements (here, the spectral modes). For a connected graph, the nullspace is span { 1 } ; for a graph with c components, grad α annihilates the c indicator-constants of the components.

A different approach to account for nonlocal spatial effects on the diffusion on graphs has been developed under the name of d -path Laplacians [30,33,34]. In a series of papers, it was proved analytically that a Mellin transformation of the d -path Laplacians produce superdiffusive behavior on graphs [33,34], which combined with fractional temporal derivatives have conduced to a generalization of the diffusion equation on graph [28] capable to describe sub-, normal, and super-diffusive dynamics on graphs.

Proposition 24

Let G = ( V , E ) be a (simple, undirected) graph with geodesic (shortest-path) distance dist ( , ) and diameter diam ( G ) . For d { 1 , , diam ( G ) } , the d-path Laplacian L d : C ( V ) C ( V ) is defined by

( L d f ) ( v ) = w V dist ( v , w ) = d ( f ( v ) f ( w ) ) .

Then,

  1. Each L d is a proper conservative-diffusive Laplacian, i.e., there exist linear operators

    grad d : C ( V ) C ( E d ) , div d : C ( E d ) C ( V ) ,

    with grad d ( 1 ) = 0 and L d = div d grad d = grad d * grad d . In particular, L d is self-adjoint, positive semidefinite, and conservative: L d 1 = 0 .

  2. For any nonnegative coefficients ( c d ) d = 1 diam ( G ) , the operator

    L d = 1 diam ( G ) c d L d

    is also a proper conservative-diffusive Laplacian. This covers, in particular, exponentially decaying weights c d = e λ d and polynomially decaying weights c d = d s with λ > 0 or s > 0 .

Proof

(i) Fix d and form the d-sphere adjacency on V : connect v and w iff dist ( v , w ) = d . Let E d { ( v w ) : dist ( v , w ) = d } denote the set of oriented d -pairs. Define the discrete d -gradient and d -divergence by

( grad d x ) ( v w ) x v x w , ( div d y ) v w V dist ( v , w ) = d y ( v w ) .

Then, for any x C ( V ) ,

( div d grad d x ) v = w dist ( v , w ) = d ( x v x w ) = ( L d x ) v .

Moreover, grad d ( 1 ) = 0 since 1 1 = 0 on every oriented pair. Thus, L d = div d grad d with a proper gradient. With the standard inner products on C ( V ) and C ( E d ) , we also have div d = grad d * , hence

L d = grad d * grad d , x , L d x = grad d x 2 0 ,

which proves self-adjointness and positive semidefiniteness. Finally, L d 1 = div d grad d 1 = 0 , so the operator is conservative.

(ii) Consider the block gradient

grad : C ( V ) d = 1 diam ( G ) C ( E d ) , grad x d = 1 diam ( G ) c d grad d x ,

with the convention c d = 0 if c d = 0 . Then, grad ( 1 ) = d c d grad d ( 1 ) = 0 and

grad * grad = d = 1 diam ( G ) c d grad d * grad d = d = 1 diam ( G ) c d L d = L .

Thus, L is of the form grad * grad with a proper gradient, hence self-adjoint, positive semidefinite, and conservative: L 1 = grad * grad 1 = 0 . For the specific choices c d = e λ d ( λ > 0 ) or c d = d s ( s > 0 ), all c d 0 , so the conclusion applies.□

Remark 25

Each L d is the standard (combinatorial) Laplacian of the d-path graph whose edges connect pairs at geodesic distance d ; hence d -path Laplacians are nonlocal (for d > 1 ) yet remain proper Laplacians in the sense of div grad with grad ( 1 ) = 0 . Weighted superpositions with c d 0 simply rescale the corresponding edge bundles and preserve the diffusive-conservative structure.

The hub-biased Laplacians L rep and L attr have been developed to account for the realistic scenarios in which a diffusive particle at a given vertex of a graph is biased away or toward nearest neighbor vertices of higher degree [31,43]. They have found applications not only in diffusive processes but in nonlinear synchronization models (refer references cited earlier).

Proposition 26

Let G = ( V , E ) be a simple undirected graph with vertex degrees k v > 0 , and let

( L R f ) ( v ) = w N v k v k w ( f ( v ) f ( w ) ) , ( L A f ) ( v ) = w N v k w k v ( f ( v ) f ( w ) ) ,

be respectively the hubs-repelling and hubs-attracting Laplacians acting on vertex functions f C ( V ) . Then, there exist linear operators

grad α : C ( V ) C ( E ) , div α : C ( E ) C ( V ) ,

where E denotes the set of oriented edges and

c α ( v , w ) = k v k w α , α { + 1 , 1 } ,

such that

L α = div α grad α , grad α ( 1 ) = 0 .

Hence, both L R and L A are proper diffusive Laplacians in the sense of Definition 4.

Proof

For each oriented edge ( v w ) E , the degree-weighted gradient is defined by

( grad α x ) ( v w ) = c α ( v , w ) ( x v x w ) = k v k w α 2 ( x v x w ) , x C ( V ) .

Then, grad α ( 1 ) = 0 , since all vertex differences vanish for the constant function. Define the corresponding divergence as the aggregation of edge fluxes leaving each vertex,

( div α y ) v = w N v c α ( v , w ) y ( v w ) , y C ( E ) .

By direct substitution,

( div α grad α x ) v = w N v c α ( v , w ) [ c α ( v , w ) ( x v x w ) ] = w N v c α ( v , w ) ( x v x w ) = ( L α x ) v .

Thus, L α = div α grad α with grad α ( 1 ) = 0 . Since every term acts as a weighted difference between neighboring vertices, L R and L A are diffusive operators, although generally asymmetric.□

Remark 27

The asymmetry of L R and L A reflects degree-dependent anisotropy of the diffusion process rather than a failure of the Laplacian structure. Indeed, both operators are similar to the symmetric matrix W A , where L R = K ( W A ) K 1 and L A = K 1 ( W A ) K , so they are diagonalizable with real, nonnegative eigenvalues. Hence, the degree-biased Laplacians remain proper within the operator-theoretic definition of this study, even though they are not self-adjoint in the Euclidean metric. They represent biased diffusions in which probability fluxes are modulated by local vertex degrees.

6 Moore-Penrose pseudoinverse of L as a proper Laplacian

The motivation of this section is twofold. On one hand we are interested in exploring the physically-sounded model proposed by Saxena et al. [80], which is a diffusion-like model inspired by electrical network based on the pseudoinverse L + of the graph Laplacian. On the other hand, is the statement in a spectral graph theory book that L + “is not always a Laplacian” because some of its off-diagonal elements can be positive (refer pp. 133–134 in [65]). Therefore, we investigate analytically whether the pseudoinverse of the graph Laplacian is a proper diffusive Laplacian or not.

Let G = ( V , E ) be a connected undirected graph with n vertices and m edges. Denote by R n × m its vertex-edge incidence matrix and by L = the combinatorial Laplacian of G . Let the singular value decomposition of be

= U Σ V , U R n × n , V R m × m , Σ = diag ( σ 1 , , σ n 1 , 0 ) ,

where σ i > 0 for i n 1 because G is connected.

Definition 28

Define the pseudogradient of G as

˜ U ( Σ + ) V ,

where Σ + = diag ( σ 1 1 , , σ n 1 1 , 0 ) is the Moore-Penrose pseudoinverse of Σ .

Lemma 29

The matrix ˜ satisfies

  1. L ˜ ˜ L = L ;

  2. ˜ ˜ L ˜ ˜ = ˜ ˜ ;

  3. ( L ˜ ˜ ) = L ˜ ˜ and ( ˜ ˜ L ) = ˜ ˜ L .

Proof

Each statement follows directly by substitution of = U Σ V , ˜ = U ( Σ + ) V , and the orthogonality of U , V . The first equality reduces to U Σ V V Σ U U ( Σ + ) V V Σ + U U Σ V V Σ U = U Σ Σ U = L , and the remaining ones are analogous.□

Theorem 30

Let L + be the Moore-Penrose pseudoinverse of the Laplacian L. Then,

L + = ˜ ˜ = ( + ) + ,

and L + satisfies the defining properties of a proper Laplacian operator

L + = grad * grad with grad ( 1 ) = 0 .

Proof

The Moore-Penrose pseudoinverse of any matrix product A A satisfies ( A A ) + = ( A + ) A + . Applying this to A = gives L + = ( + ) + = ( U ( Σ + ) V ) ( V ( Σ + ) U ) = U ( Σ + ) Σ + U . Setting grad + , we obtain L + = grad * grad .

Since G is connected, Ran ( ) = 1 and therefore, 1 Ker ( ) = Ran ( ) . The Moore-Penrose properties imply that + 1 = 0 , i.e., grad ( 1 ) = 0 .□

Remark 31

(Spectral form and geometric meaning.) In the orthonormal basis of Laplacian eigenvectors L = U Λ U , where Λ = diag ( λ 1 , , λ n ) with 0 = λ 1 < λ 2 λ n , one has

L + = U Λ + U , Λ + = diag ( 0 , λ 2 1 , , λ n 1 ) .

Thus, L + plays the role of the inverse diffusion operator restricted to 1 : it corresponds to the integral kernel of the Green’s function for the diffusion equation t x + L x = 0 .

Remark 32

(On nonlocality and diffusion.) Although L + is diffusive (being positive semidefinite and self-adjoint), its associated pseudogradient ˜ is nonlocal: each row mixes all vertices rather than single edge differences. This agrees with the idea that L + represents the steady-state response or long-range coupling induced by the diffusion generated by L . Hence, L + is a proper (but nonlocal) Laplacian under the general definition L = div grad with grad ( 1 ) = 0 .

Example 33

(Pseudogradient for the pseudoinverse Laplacian of a tree) Let T = ( V , E ) be the tree with V = { 1 , 2 , 3 , 4 , 5 } and E = { ( 1 , 2 ) , ( 2 , 3 ) , ( 3 , 4 ) , ( 3 , 5 ) } . The pseudogradient ˜ = ˜ ( T ) such that L + ( T ) = ˜ ˜ is given by

˜ ( T ) = 1 5 4 3 1 1 1 3 1 1 1 2 1 1 1 2 4 1 1 2 1 4 .

This construction follows the method of Bapat [10], later extended by Alazemi et al. [2,3] (refer also [24]), who derived combinatorial and signed-graph generalizations of the Moore-Penrose pseudoinverse of graph Laplacians.

These results can be interpreted combinatorially as follows. For any edge e = ( i , j ) E , removing e splits T into two subtrees T + ( i , j ) and T ( i , j ) . Let n + and n be their respective numbers of vertices, and n = V . Then, following Bapat [10],

˜ v k , e ( i , j ) = n n , v k T + ( i , j ) , n + n , v k T ( i , j ) .

In words, each entry of the pseudogradient measures the size imbalance between the two subtrees created by deleting e , normalized by n . The sign encodes whether the vertex v k belongs to the same component as the head or the tail of the oriented edge.

We can also give a path-count interpretation to the entries of the gradient, which is intimately related to that of Bapat [10] for the case of trees, but which we present here for the sake of interpretability. Let π ( i , j ) k denote the number of paths (of any length) in T that start at vertex v k and traverse the oriented edge ( i , j ) . Then,

˜ v k , e ( i , j ) = 1 n π ( i , j ) k , if the path orientation agrees with ( i , j ) , π ( i , j ) k , if it is opposite .

Thus, the pseudogradient ˜ has a purely combinatorial interpretation: it encodes, for each vertex-edge pair, the normalized signed count of all paths passing through that edge.

In closing, the pseudogradient ˜ ( T ) is nonlocal: every vertex interacts with every edge of T , with weights determined by the sizes of the complementary subtrees. Its entries are both positive and negative, reflecting the orientation and relative position of each vertex with respect to each cut defined by an edge. Therefore, although L + ( T ) represents a signed complete graph, it remains a proper Laplacian under our operator-based definition, since

L + ( T ) = ˜ ˜ and ˜ ( 1 ) = 0 .

Bapat’s formula [10] shows that the entries of the pseudoinverse of a tree Laplacian depend only on the sizes of the subtrees obtained by deleting an edge. This provides a concrete realization of the abstract operator framework introduced in this work: the pseudogradient ˜ is a nonlocal, signed gradient satisfying ˜ ( 1 ) = 0 and yielding a diffusive Gram form L + = ˜ ˜ . Hence, the pseudoinverse of the Laplacian of a tree is a proper Laplacian in the generalized sense. Alazemi et al. [2,3] extended these constructions to threshold and signed graphs, demonstrating that such nonlocal pseudogradients arise naturally beyond trees.

6.1 Diffusive dynamics with L and L +

We finally compare the standard diffusion governed by the graph Laplacian L with that induced by its Moore-Penrose pseudoinverse L + in undirected graphs. Both are proper Laplacians in the sense of this study, but they operate in complementary geometric spaces: L diffuses over the adjacency (metric) structure of the graph, whereas L + diffuses over the effective-resistance (dual) geometry. Let us then consider { L , L + } and the diffusive process

x ˙ ( t ) = x ( t ) , x ( 0 ) = x 0 ,

which has solution x ( t ) = e t x 0 . Because is positive semidefinite we have

lim t x ( t ) = 1 x 0 n 1 ,

for both Laplacians. However, because L has eigenvalues 0 = μ 1 < μ 2 μ n , the rate of convergence to the steady state is dictated by μ 2 (the algebraic connectivity): the larger, μ 2 , the faster the convergence. On the other hand, L + = U Λ + U , where Λ + = diag ( 0 , 1 μ 2 , , 1 μ n ) , and the rate of convergence of the diffusion is now controlled by 1 μ n , the reciprocal of the largest eigenvalue of L . Hence, while L -diffusion emphasizes the tightest connected modes of the graph, L + -diffusion emphasizes its loosest (resistance-based) modes. Hence, L + -diffusion is faster than L -diffusion iff

(⋆) 1 μ n > μ 2 .

The quadratic form of L + is x L + x = 1 2 i , j Ω i j ( x i x j ) 2 , with Ω i j the effective resistance between i and j . Thus, L + acts as the Laplacian of the complete weighted graph whose edge weights are the resistive distances of ϒ . In stretched graphs (long paths, narrow bottlenecks), many Ω i j are large, so L + strongly penalizes global discrepancies and accelerates homogenization relative to adjacency-based diffusion by L . That is, basically we need graphs with very small algebraic connectivity, which also has relatively small Laplacian spectral radius. Graphs with small algebraic connectivity have been previously studied in the literature (refer Section 6.8 in [83]). These graphs tend to be elongated graphs with large diameter. In the case of trees, the minimum μ 2 for graph with fixed size and diameter is obtained by attaching an almost equal number of pendant vertices at each end of a path [36]. On the other hand, some of the graphs with small Laplacian spectral radius are the path and the cycle with n vertices (refer Section 6.3 in [83]). Then, we have the following example.

Example 34

For the path P n , the Laplacian eigenvalues are μ k = 2 1 cos k π n + 1 , so μ 2 π 2 n 2 and μ n 4 as n , and thus, 1 μ n 0.25 > μ 2 for n 6 . For the cycle C n , μ k = 2 1 cos 2 π k n , so μ 2 4 π 2 n 2 and μ n = 4 , giving 1 μ n = 0.25 > μ 2 for n 13 .

Inspired by the results of [36] we explore other types of graph which we prove to display μ n μ 2 < 1 .

6.1.1 Lollipop graphs L ( m , )

Let L ( m , ) be the graph obtained by attaching a path P to a clique K m by a single bridge edge. Denote by v the clique vertex incident to the bridge. Then, k v = m and k v = m 1 for other clique vertices; the path has degrees 1 (ends) or 2 (internal). Hence, Δ ( L ( m , ) ) = m .

Proposition 35

(Rates on lollipops) For L ( m , ) , one has

m + 1 μ n 2 m , μ 2 2 vol min ,

where vol min min { vol ( clique s i d e ) , vol ( path s i d e ) } for the cut that removes the single bridge edge. In particular, if the path side has larger volume (e.g., sufficiently large compared to m ), then vol min = vol ( clique s i d e ) = m + ( m 1 ) 2 = m 2 m + 1 , and hence,

μ 2 2 m 2 m + 1 2 m 2 m .

Consequently, for all m 2 ,

1 μ n 1 2 m > 2 m 2 m + 1 μ 2

whenever m is sufficiently large; in particular, ( ) holds for all large m uniformly in as soon as vol ( path s i d e ) vol ( clique s i d e ) .

Proof

Bounds on μ n . The subgraph induced by v and its m neighbors contains the star K 1 , m . By Lemma 1, adding edges cannot decrease μ n , hence μ n ( L ( m , ) ) μ n ( K 1 , m ) = m + 1 . The upper bound μ n 2 Δ = 2 m follows from Lemma 2.

Bound on μ 2 . Consider the cut that removes the unique bridge edge; then E ( S , S ¯ ) = 1 and ϕ ( ϒ ) = 1 min { vol ( S ) , vol ( S ¯ ) } . If the path side has larger volume than the clique side, then vol min = vol ( clique side ) = k v + v K m v v ( m 1 ) = m + ( m 1 ) 2 = m 2 m + 1 . Hence, ϕ 1 ( m 2 m + 1 ) and Lemma 3 yields μ 2 2 ϕ 2 ( m 2 m + 1 ) . The final comparison with 1 μ n 1 ( 2 m ) gives the claim for all sufficiently large m .□

Remark 36

If the clique side was smaller in volume, one would obtain μ 2 2 vol ( path side ) , which can be even smaller for very long paths. Thus, the conclusion ( ) is robust: any lollipop where the path side is not the bottleneck in volume satisfies faster L + -diffusion.

Example 37

We consider here a graph formed by one clique of 20 vertices connected by a single edge to a chain of 80 vertices. In this case, 1 μ n 0.0476 and μ 2 0.0010 . In Figure 1, we illustrate the progress of the diffusive dynamics on this graph using the standard Laplacian (panel a) and its pseudoinverse (panel b). The simulations are performed with exactly the same initial condition.

Figure 1 
                     Illustration of the progress of the diffusive dynamics on a graph formed by one clique of 20 vertices connected by a single edge to a chain of 80 vertices. (a) Standard diffusion. (b) Diffusion based on the pseudoinverse of the Laplacian.
Figure 1

Illustration of the progress of the diffusive dynamics on a graph formed by one clique of 20 vertices connected by a single edge to a chain of 80 vertices. (a) Standard diffusion. (b) Diffusion based on the pseudoinverse of the Laplacian.

6.1.2 Two cliques joined by a chain (barbell-with-handle)

Let B ( m , ) be the graph obtained by connecting two disjoint cliques of size m by a simple path P whose endpoints attach by a single edge to one vertex in each clique. The two junction vertices have degree m ; all other clique vertices have degree m 1 ; path vertices have degree 1 or 2. Thus, Δ ( B ( m , ) ) = m .

Proposition 38

(Rates on two cliques joined by a chain) For B ( m , ) , one has

m + 1 μ n 2 m , μ 2 2 min { vol ( S ) , vol ( S ¯ ) } ,

where the cut S is chosen across any single edge of the interconnecting chain. In particular, cutting the chain near its middle gives E ( S , S ¯ ) = 1 and

min { vol ( S ) , vol ( S ¯ ) } m 2 m + 1

(because one side contains an entire K m ), hence μ 2 2 ( m 2 m + 1 ) . Therefore, for all sufficiently large m,

1 μ n 1 2 m > 2 m 2 m + 1 μ 2 ,

so ( ) holds and L + -diffusion is faster than L-diffusion.

Proof

The bounds on μ n are as in Proposition 35: the induced star around a junction vertex gives μ n m + 1 , and Lemma 2 gives μ n 2 m . For μ 2 , choose S so that E ( S , S ¯ ) = 1 across the chain. Then, ϕ = 1 min { vol ( S ) , vol ( S ¯ ) } . Either side contains at least one full K m , whose volume is m + ( m 1 ) 2 = m 2 m + 1 , ignoring the chain contribution which only increases volume. Hence ϕ 1 ( m 2 m + 1 ) and μ 2 2 ϕ by Lemma 3.□

Remark 39

If the chain length equals the clique size ( = m ), one still has μ 2 = O ( m 2 ) while μ n [ m + 1,2 m ] , hence ( ) holds for all large m . As the chain grows, the minimal side volume in the middle cut only increases, so the upper bound on μ 2 remains O ( m 2 ) (or smaller), further favoring L + .

Example 40

We consider here a graph formed by two cliques of 33 vertices each connected by a chain of 33 vertices. The connection between the cliques and the chain are made by a single edge to one vertex in each clique. In this case 1 μ n 0.0476 and μ 2 0.0041 . In Figure 2 we illustrate the progress of the diffusive dynamics on this graph using the standard Laplacian (panel a) and its pseudoinverse (panel b). The simulations were carried out with exactly the same initial condition.

Figure 2 
                     Illustration of the progress of the diffusive dynamics on a barbell graph formed by two cliques of 33 vertices each connected by a linear chain of 33 vertices. (a) Standard diffusion. (b) Diffusion based on the pseudoinverse of the Laplacian.
Figure 2

Illustration of the progress of the diffusive dynamics on a barbell graph formed by two cliques of 33 vertices each connected by a linear chain of 33 vertices. (a) Standard diffusion. (b) Diffusion based on the pseudoinverse of the Laplacian.

In all these stretched families (lollipops, clique-chain-clique, paths, cycles), the bottleneck structure depresses μ 2 to O ( m 2 ) or O ( n 2 ) , while μ n remains Θ ( Δ ) ; consequently, ( ) holds and L + -diffusion equilibrates strictly faster than standard L -diffusion.

Remark 41

Both L and L + are proper Laplacians in the operator sense: they can be written as div grad with grad ( 1 ) = 0 , but they correspond to dual notions of locality. L acts on physical adjacency, while L + acts on resistance space. Together they form a complementary pair of diffusion generators on a graph and its dual metric.

6.1.3 Example of a real-world network

Is it possible to find real world networks in which μ n μ 2 < 1 ? Here we consider the call network of the software MySQL consisting of 1,480 vertices representing software modules and 4,190 edges among them. We have computed that for this network

μ 2 ( L ) = 0.0039 and 1 μ n ( L ) = 0.0045 ,

so the slowest exponential rate of L + -diffusion exceeds that of L -diffusion by 0.0045 0.0039 1 15 % . For any initial condition x 0 , writing x L ( t ) = e t L x 0 and x L + ( t ) = e t L + x 0 , the deviations from the mean satisfy

x L ( t ) x ¯ 1 e μ 2 t x 0 x ¯ 1 , x L + ( t ) x ¯ 1 e 1 μ n t x 0 x ¯ 1 ,

with x ¯ = 1 n 1 x 0 . Thus, the observed trajectories should display a visibly faster decay for L + , consistent with our theory that L + diffuses in effective-resistance geometry and can outpace adjacency-based diffusion on stretched or bottlenecked graphs.

We verify these predictions by simulating the temporal evolution of the diffusion in this network. As initial condition we select here that all the initial concentrations were allocated in a single vertex. We perform the simulations using L and L + with exactly the same initial condition. As can be seen in Figure 3, the rate of convergence – measured here as the standard deviation of all values of the vector x ( t ) – is faster when the diffusion is controlled by L + than by L . The main reason of this difference in the studied network is that it consists of two main clusters, which are very poorly connected among them. Therefore, the algebraic connectivity of the network is very small and μ n ( L ) is not so big, such that μ 2 ( L ) μ n ( L ) < 1 .

Figure 3 
                     Illustration of the temporal evolution of the diffusive dynamics toward the steady state on the network of MySQL software.
Figure 3

Illustration of the temporal evolution of the diffusive dynamics toward the steady state on the network of MySQL software.

7 Conclusion

We have introduced a unifying definition of a proper graph Laplacian as any operator that can be expressed in the form L = div grad with grad ( 1 ) = 0 , acting between vertex functions and functions defined on suitable graph structures. This operator-theoretic perspective recovers the classical Laplacian as a special case but extends naturally to nonlocal, weighted, and higher-order settings. It also allows a systematic distinction between proper and improper Laplacians on structural grounds, independent of ad hoc algebraic properties such as symmetry, sign patterns, or zero row-sum.

Within this framework we have shown that several matrices traditionally referred to as “Laplacians” are, in fact, not proper Laplacians. This includes the in- and out-degree Laplacians (which describe advection rather than diffusion), the signless and magnetic Laplacians (which arise as Schur complements of proper extended operators), and the deformed Laplacian. In contrast, the signed Laplacian, the Lerman–Ghosh Laplacian, and the fractional and degree-biased Laplacians were shown to behave as proper diffusive operators; however, certain members of this family fail to conserve total mass in diffusive processes and are therefore nonconservative. Finally, we demonstrated that the Moore-Penrose pseudoinverse L + of the standard Laplacian is itself a proper (nonlocal) Laplacian, governing diffusion in the effective-resistance geometry of the graph. For certain classes of stretched or bottlenecked networks (paths, cycles, lollipops, and clique-chain-clique graphs), the L + -based diffusion equilibrates faster than the standard Laplacian flow, revealing a duality between physical and resistive diffusion. These results may be extended to other concepts of Laplacians not considered here like the “distance Laplacian” [6,9,73] as to the pseudoinverse of signed Laplacians just to mention a couple of examples [41].

Overall, this work establishes an operator-theoretic foundation for graph Laplacians that unifies classical and generalized constructions, clarifies long-standing ambiguities in terminology, and identifies a minimal structural criterion – existence of a genuine gradient annihilating constants for Laplacian-type operators. This approach opens the way to new forms of graph dynamics beyond the local edge-based paradigm, while preserving the geometric essence of diffusion on complex networks.

Acknowledgments

The author thanks the Editor and referees for valuable comments and suggestions which improved significantly the presentation of the results.

  1. Funding information: Financial support from Agencia Estatal de Investigación (AEI, MCI, Spain) MCIN/AEI/ 10.13039/501100011033 under grant PID2023-149473NB-I00 and by Agencia Estatal de Investigación (AEI, MCI, Spain) MCIN/AEI/10.13039/501100011033 and Fondo Europeo de Desarrollo Regional (FEDER, UE) under the María de Maeztu Program for units of Excellence in R&D, grant CEX2021-001164-M) are also acknowledged.

  2. Author contributions: The author confirms the sole responsibility for the conception of the study, presented results, and manuscript preparation.

  3. Conflict of interest: The author states no conflict of interest.

  4. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] M. Abudayah, O. Alomari, and O. AbuGhneim, Inverse of Hermitian adjacency matrix of a mixed graph, Appl. Math. 16 (2022), no. 5, 823–828, https://doi.org/10.18576/amis/160516.Suche in Google Scholar

[2] A. Alazemi, M. Anđelić, and S. Mallik, Signed graphs and inverses of their incidence matrices, Linear Algebra Appl. 694 (2024), 78–100, https://doi.org/10.1016/j.laa.2024.04.012. Suche in Google Scholar

[3] A. Alazemi, M. Anđelić, J. Milenković, and J. N. Radenković, Combinatorial versus algebraic formulae for the Moore-Penrose inverse of a Laplacian matrix of a threshold graph, J. Comput. Appl. Math. 442 (2024), 115714, https://doi.org/10.1016/j.cam.2023.115714.Suche in Google Scholar

[4] N. Alon and V. Milman, λ1, isoperimetric inequalities for graphs, and superconcentrators, J. Combinat. Theory Series B 38 (1985), no. 1, 73–88, https://doi.org/10.1016/0095-8956(85)90092-9. Suche in Google Scholar

[5] C. Altafini, Consensus problems on networks with antagonistic interactions, IEEE Trans. Automatic Control 58 (2013), no. 4, 935–946, https://doi.org/10.1109/TAC.2012.2224251. Suche in Google Scholar

[6] M. Aouchiche and P. Hansen, Two Laplacians for the distance matrix of a graph, Linear Algebra Appl. 439 (2013), no. 1, 21–33, https://doi.org/10.1016/j.laa.2013.02.030. Suche in Google Scholar

[7] N. Athmouni, H. Baloudi, M. Damak, and M. Ennaceur, The magnetic discrete Laplacian inferred from the Gauss-Bonnet operator and application, Ann. Funct. Anal. 12 (2021), no. 2, 33, https://doi.org/10.1007/s43034-021-00119-8. Suche in Google Scholar

[8] R. Balaji and R. B. Bapat, On Euclidean distance matrices, Linear Algebra Appl. 424 (2007), no. 1, 108–117, https://doi.org/10.1016/j.laa.2006.05.013. Suche in Google Scholar

[9] R. Balaji and V. Gupta, On distance Laplacian matrices of weighted trees, Operat. Matrices 18 (2024), no. 1, 97–114, https://doi.org/10.7153/oam-2024-18-07. Suche in Google Scholar

[10] R. B. Bapat, Moore-Penrose inverse of the incidence matrix of a tree, Linear Multilinear Algebra 42 (1997), no. 2, 159–167, https://doi.org/10.1080/03081089708818496. Suche in Google Scholar

[11] G. Berkolaiko, Nodal count of graph eigenfunctions via magnetic perturbation, Anal. PDE 6 (2013), no. 5, 1213–1233, https://doi.org/10.2140/apde.2013.6.1213. Suche in Google Scholar

[12] E. Blåsten, H. Isozaki, M. Lassas, and J. Lu, Inverse problems for discrete heat equations and random walks for a class of graphs, SIAM J. Discrete Math. 37 (2023), no. 2, 831–863, https://doi.org/10.1137/21M1439936. Suche in Google Scholar

[13] L. Böttcher and M. A. Porter, Complex networks with complex weights, Phys. Rev. E 109 (2024), no. 2, 024314, https://doi.org/10.1103/PhysRevE.109.024314. Suche in Google Scholar PubMed

[14] A. K. Chandra, P. Raghavan, W. L. Ruzzo, and R. Smolensky, The electrical resistance of a graph captures its commute and cover times, Proceedings of the Twenty-First Annual ACM Symposium on Theory of Computing (STOC), ACM, 1989, pp. 574–586, https://doi.org/10.1145/73007.73062. Suche in Google Scholar

[15] A. Chapman, Advection on graphs, Semi-Autonomous Networks. Effective Control of Networked Systems through Protocols, Design, and Modeling, Springer International Publishing, Switzerland, 2015, pp. 3–16, https://doi.org/10.1007/978-3-319-15010-9_1. Suche in Google Scholar

[16] A. Chapman, E. Schoof, and M. Mesbahi, Advection on networks with an application to decentralized load balancing, 2012 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), IEEE, 2012, pp. 2680–2681, https://doi.org/10.1109/IROS.2012.6386284. Suche in Google Scholar

[17] J. Cheeger, A lower bound for the smallest eigenvalue of the Laplacian, in R. C. Gunning, eds., Problems in Analysis, Princeton University Press, Princeton, New Jersey, 1970, pp. 195–199. 10.1515/9781400869312-013Suche in Google Scholar

[18] H. Chen, Random walks and the effective resistance sum rules, Discrete Appl. Math. 158 (2010), no. 15, 1691–1700, https://doi.org/10.1016/j.dam.2010.05.020. Suche in Google Scholar

[19] F. R. K. Chung, CBMS Regional Conference Series in Mathematics, no. 92, American Mathematical Society, 1997, https://doi.org/10.1090/cbms/092.Suche in Google Scholar

[20] M. Cucuringu, H. Li, H. Sun, and L. Zanetti, Hermitian matrices for clustering directed graphs: Insights and applications, Proceedings of AISTATS, PMLR, 2020, pp. 983–992, https://proceedings.mlr.press/v108/cucuringu20a.html. Suche in Google Scholar

[21] D. Cvetković and S. K. Simić, Towards a spectral theory of graphs based on signless Laplacian I, Publ. Inst. Math. (Beograd) 85 (2009), no. 99, 19–33, https://doi.org/10.2298/PIM0999019C. Suche in Google Scholar

[22] D. Cvetković and S. K. Simić, Towards a spectral theory of graphs based on the signless Laplacian, ii, Linear Algebra Appl. 432 (2010), no. 9, 2257–2272, https://doi.org/10.1016/j.laa.2009.05.020. Suche in Google Scholar

[23] D. Cvetković and S. K. Simić, Towards a spectral theory of graphs based on the signless Laplacian, iii, Appl. Anal. Discrete Math. 4 (2010), no. 1, 156–166, https://doi.org/10.2298/AADM1000001C. Suche in Google Scholar

[24] Y. Dai and S. Chen, Invariants for incidence matrix of a tree, J. Algebraic Combinat. 60 (2024), no. 4, 1011–1029, https://doi.org/10.1007/s10801-024-01361-8. Suche in Google Scholar

[25] K. C. Das, The Laplacian spectrum of a graph, Comput. Math. Appl. 48 (2004), no. 5–6, 715–724, https://doi.org/10.1016/j.camwa.2004.05.005. Suche in Google Scholar

[26] Y. Colin de Verdière, Magnetic interpretation of the nodal defect on graphs, Anal. PDE 6 (2013), no. 5, 1235–1242, https://doi.org/10.2140/apde.2013.6.1235. Suche in Google Scholar

[27] K. Devriendt, Effective resistance is more than distance: Laplacians, simplices and the Schur complement, Linear Algebra Appl. 639 (2022), 24–49, https://doi.org/10.1016/j.laa.2022.01.002. Suche in Google Scholar

[28] F. Diaz-Diaz and E. Estrada, Time and space generalized diffusion equation on graph/networks, Chaos Solitons Fractals 156 (2022), 111791, https://doi.org/10.1016/j.chaos.2022.111791. Suche in Google Scholar

[29] J. Dodziuk and V. Mathai, Kato’s inequality and asymptotic spectral properties for discrete magnetic Laplacians, Contemp. Math. 398 (2006), 69–81, https://doi.org/10.48550/arXiv.math/0312450. Suche in Google Scholar

[30] E. Estrada, Path Laplacian matrices: Introduction and application to the analysis of consensus in networks, Linear Algebra Appl. 436 (2012), no. 9, 3373–3391, https://doi.org/10.1016/j.laa.2011.11.032. Suche in Google Scholar

[31] E. Estrada, “hubs-repelling” Laplacian and related diffusion on graphs/networks, Linear Algebra Appl. 596 (2020), 256–280, https://doi.org/10.1016/j.laa.2020.03.012. Suche in Google Scholar

[32] E. Estrada, Conservative versus non-conservative diffusion toward a target in a networked environment, Target Search Problems, Springer, Switzerland AG, 2024, pp. 511–540, https://doi.org/10.1007/978-3-031-67802-8_21. Suche in Google Scholar

[33] E. Estrada, E. Hameed, N. Hatano, and M. Langer, Path Laplacian operators and superdiffusive processes on graphs. i. one-dimensional case, Linear Algebra Appl. 523 (2017), 307–334, https://doi.org/10.1016/j.laa.2017.02.027. Suche in Google Scholar

[34] E. Estrada, E. Hameed, M. Langer, and A. Puchalska, Path Laplacian operators and superdiffusive processes on graphs. ii. two-dimensional lattice, Linear Algebra Appl. 555 (2018), 373–397, https://doi.org/10.1016/j.laa.2018.06.026. Suche in Google Scholar

[35] J. S. Fabila-Monroy, F. Lledó, and O. Post, Matching number, Hamiltonian graphs and discrete magnetic Laplacians, Linear Algebra Appl. 648 (2022), 86–100, https://doi.org/10.1016/j.laa.2022.02.006. Suche in Google Scholar

[36] S. Fallat and S. Kirkland, Extremizing algebraic connectivity subject to graph theoretic constraints, Electronic J. Linear Algebra 3 (1998), 48–74, https://doi.org/10.13001/1081-3810.1014. Suche in Google Scholar

[37] M. Fanuel, C. M. Alaiz, and J. A. K. Suykens, Magnetic eigenmaps for community detection in directed networks, Phys. Rev. E 95 (2017), no. 2, 022302, https://doi.org/10.1103/PhysRevE.95.022302. Suche in Google Scholar PubMed

[38] M. Fanuel and J. A. K. Suykens, Deformed Laplacians and spectral ranking in directed networks, Appl. Comput. Harmonic Anal. 47 (2019), no. 2, 397–422, https://doi.org/10.1016/j.acha.2017.09.002. Suche in Google Scholar

[39] M. Fiedler, Algebraic connectivity of graphs, Czechoslovak Math. J. 23 (1973), no. 2, 298–305, https://doi.org/10.21136/CMJ.1973.101168. Suche in Google Scholar

[40] A. Fontan and C. Altafini, On the properties of Laplacian pseudoinverses, Proc. 60th IEEE Conference on Decision and Control (CDC), IEEE, 2021, pp. 5538–5543, https://doi.org/10.1109/CDC45484.2021.9683525. Suche in Google Scholar

[41] A. Fontan and C. Altafini, Pseudoinverses of signed Laplacian matrices, SIAM J. Matrix Anal. Appl. 44 (2023), no. 2, 622–647, https://doi.org/10.1137/22M1493392. Suche in Google Scholar

[42] S. Furutani, T. Shibahara, M. Akiyama, K. Hato, and M. Aida, Graph signal processing for directed graphs based on the Hermitian Laplacian, Machine Learning and Knowledge Discovery in Databases: ECML PKDD 2019, Part I, Springer, 2020, pp. 447–463, https://doi.org/10.1007/978-3-030-46150-8_27. Suche in Google Scholar

[43] L. V. Gambuzza, M. Frasca, and E. Estrada, Hubs-attracting Laplacian and related synchronization on networks, SIAM J. App. Dyn. Syst. 19 (2020), no. 2, 1057–1079, https://doi.org/10.1137/19M1287663. Suche in Google Scholar

[44] A. Ghosh, S. Boyd, and A. Saberi, Minimizing effective resistance of a graph, SIAM Review 50 (2008), no. 1, 37–66, https://doi.org/10.1137/050645452. Suche in Google Scholar

[45] R. Ghosh, K. Lerman, T. Surachawala, K. Voevodski, and S. Teng, Non-conservative diffusion and its application to social network analysis, J. Complex Netw. 12 (2024), no. 1, cnae006, https://doi.org/10.48550/arXiv.1102.4639. Suche in Google Scholar

[46] C. Godsil and G. Royle, The Laplacian of a Graph, Springer, New York, NY, 2001, pp. 279–306, https://doi.org/10.1007/978-1-4613-0163-9_13. Suche in Google Scholar

[47] R. Grone and R. Merris, The Laplacian spectrum of a graph ii, SIAM J. Discrete Math. 7 (1994), no. 2, 221–229, https://doi.org/10.1137/S0895480191222653. Suche in Google Scholar

[48] R. Grone, R. Merris, and V. S. Sunder, The Laplacian spectrum of a graph, SIAM J. Matrix Anal. Appl. 11 (1990), no. 2, 218–238, https://doi.org/10.1137/0611016. Suche in Google Scholar

[49] K. Guo and B. Mohar, Hermitian adjacency matrix of digraphs and mixed graphs, J. Graph Theory 85 (2017), no. 1, 217–248, https://doi.org/10.1002/jgt.22057. Suche in Google Scholar

[50] I. Gutman and W. Xiao, Generalized inverse of the Laplacian matrix and some applications, Bulletin de laAcadémie Serbe des Sciences et des Arts. Classe des Sciences Mathématiques et Naturelles. Sciences Mathématiques, 2004, pp. 15–23, https://www.jstor.org/stable/44095641. 10.2298/BMAT0429015GSuche in Google Scholar

[51] K. Hayashi, S. G. Aksoy, and H. Park, Skew-symmetric adjacency matrices for clustering directed graphs, Proc. 2022 IEEE International Conference on Big Data (Big Data), IEEE, 2022, pp. 555–564, https://doi.org/10.1109/BigData55660.2022.10020413. Suche in Google Scholar

[52] Y. Higuchi, S. Kubota, and E. Segawa, On symmetric spectra of Hermitian adjacency matrices for non-bipartite mixed graphs, Discrete Math. 347 (2024), no. 5, 113911, https://doi.org/10.1016/j.disc.2024.113911. Suche in Google Scholar

[53] Y. Higuchi and T. Shirai, A remark on the spectrum of magnetic Laplacian on a graph, Yokohama Math. J. 47 (1999), no. Special, 129–141. Suche in Google Scholar

[54] D. J. Klein and M. Randić, Resistance distance, J. Math. Chem. 12 (1993), no. 1, 81–95, https://doi.org/10.1007/BF01164627. Suche in Google Scholar

[55] E. Korotyaev and N. Saburova, Magnetic Schrödinger operators on periodic discrete graphs, J. Funct. Anal. 272 (2017), no. 4, 1625–1660, https://doi.org/10.1016/j.jfa.2016.12.015. Suche in Google Scholar

[56] E. Korotyaev and N. Saburova, Trace formulas for magnetic schrödinger operators on periodic graphs and their applications, Linear Algebra Appl. 676 (2023), 395–440, https://doi.org/10.1016/j.laa.2023.07.025. Suche in Google Scholar

[57] J. Kunegis, S. Schmidt, A. Lommatzsch, J. Lerner, E. W. De Luca, and S. Albayrak, Spectral analysis of signed graphs for clustering, prediction and visualization, Proc. 2010 SIAM International Conference on Data Mining, SIAM, 2010, pp. 559–570, https://doi.org/10.1137/1.9781611972801.49. Suche in Google Scholar

[58] K. Lerman and R. Ghosh, Network structure, topology, and dynamics in generalized models of synchronization, Phys. Rev. E 86 (2012), no. 2, 026108, https://doi.org/10.1103/PhysRevE.86.026108. Suche in Google Scholar PubMed

[59] S. Li and Y. Yu, Hermitian adjacency matrix of the second kind for mixed graphs, Discrete Math. 345 (2022), no. 5, 112798, https://doi.org/10.1016/j.disc.2022.112798. Suche in Google Scholar

[60] G. Lindmark and C. Altafini, Investigating the effect of edge modifications on networked control systems: Stability analysis, Automatica 149 (2023), 110801, https://doi.org/10.1016/j.automatica.2022.110801. Suche in Google Scholar

[61] N. Masuda, M. A. Porter, and R. Lambiotte, Random walks and diffusion on networks, Phys. Reports 716 (2017), 1–58, https://doi.org/10.1016/j.physrep.2017.07.007. Suche in Google Scholar

[62] R. Merris, Laplacian matrices of graphs: a survey, Linear Algebra Appl. 197 (1994), 143–176, https://doi.org/10.1016/0024-3795(94)90486-3. Suche in Google Scholar

[63] M. Mesbahi and M. Egerstedt, Graph Theoretic Methods in Multiagent Networks, Princeton University Press, Princeton, New Jersey, 2010, https://www.torrossa.com/en/resources/an/5573598. 10.1515/9781400835355Suche in Google Scholar

[64] T. Michelitsch, A. Pérez Riascos, B. Collet, A. Nowakowski, and F. Nicolleau, Fractional dynamics on networks and lattices, John Wiley & Sons, London, UK, 2019, https://doi.org/10.1002/9781119608165. Suche in Google Scholar

[65] P. Van Mieghem, Graph spectra for complex networks, Cambridge University Press, Cambridge, UK, 2023, https://doi.org/10.1017/CBO9780511921681. Suche in Google Scholar

[66] P. Van Mieghem, K. Devriendt, and H. Cetinay, Pseudoinverse of the Laplacian and best spreader node in a network, Phys. Rev. E 96 (2017), no. 3, 032311, https://doi.org/10.1103/PhysRevE.96.032311. Suche in Google Scholar PubMed

[67] M. Miranda, P. Moreno-Spiegelberg, and E. Estrada, Heterogeneous consensus dynamics through reaction-diffusion models on graphs/networks, Commun. Nonlinear Sci. Numer. Simulation 153 (2025), 109160, https://hal.science/hal-04996182v1/document. 10.1016/j.cnsns.2025.109160Suche in Google Scholar

[68] B. Mohar, The Laplacian spectrum of graphs, Graph Theory, Combinatorics, and Applications, Y. Alavi, G. Chartrand, O. R. Oellermann, and A. J. Schwenk, eds., vol. 2, Wiley, Hoboken, NJ, USA, 1991, pp. 871–898. Suche in Google Scholar

[69] B. Mohar, A new kind of Hermitian matrices for digraphs, Linear Algebra Appl. 584 (2020), 343–352. 10.1016/j.laa.2019.09.024Suche in Google Scholar

[70] P. P. Mondal, R. B. Bapat, and F. Atik, On the inverse and Moore-Penrose inverse of resistance matrix of graphs with more general matrix weights, J. Appl. Math. Comput. 69 (2023), no. 6, 4805–4820, https://doi.org/10.1007/s12190-023-01945-w. Suche in Google Scholar

[71] F. Morbidi, The deformed consensus protocol, Automatica 49 (2013), no. 10, 3049–3055, https://doi.org/10.1016/j.automatica.2013.07.006. Suche in Google Scholar

[72] F. Morbidi, Functions of the Laplacian matrix with application to distributed formation control, IEEE Trans. Control Network Syst. 9 (2021), no. 3, 1459–1467, https://doi.org/10.1109/TCNS.2021.3113263. Suche in Google Scholar

[73] M. Nath and S. Paul, On the distance Laplacian spectra of graphs, Linear Algebra Appl. 460 (2014), 97–110, https://doi.org/10.1016/j.laa.2014.07.025. Suche in Google Scholar

[74] R. Olfati-Saber, J. A. Fax, and R. M. Murray, Consensus and cooperation in networked multi-agent systems, Proc. IEEE 95 (2007), no. 1, 215–233, https://doi.org/10.1109/JPROC.2006.887293. Suche in Google Scholar

[75] J. L. Palacios, Resistance distance in graphs and random walks, Int. J. Quantum Chem. 81 (2001), no. 1, 29–33, https://doi.org/10.1002/1097-461X(2001)81:1<29::AID-QUA6>3.0.CO;2-Y. Suche in Google Scholar

[76] L. Pan, H. Shao, and M. Mesbahi, Laplacian dynamics on signed networks, Proceedings of the 55th IEEE Conference on Decision and Control (CDC), IEEE, 2016, pp. 891–896, https://doi.org/10.1109/CDC.2016.7798380. Suche in Google Scholar

[77] B. Ramamurthy, R. B. Bapat, and S. Goel, On resistance matrices of weighted balanced digraphs, Linear Multilinear Algebra 71 (2023), no. 13, 2222–2248, https://doi.org/10.1080/03081087.2022.2094866. Suche in Google Scholar

[78] A. Pérez Riascos and J. L. Mateos, Fractional dynamics on networks: Emergence of anomalous diffusion and Lévy flights, Phys. Rev. E 90 (2014), no. 3, 032809, https://doi.org/10.1103/PhysRevE.90.032809. Suche in Google Scholar

[79] A. Pérez Riascos, T. M. Michelitsch, B. A. Collet, A. F. Nowakowski, and F. C. G. A. Nicolleau, Random walks with long-range steps generated by functions of Laplacian matrices, J. Stat. Mech Theory Experiment 2018 (2018), no. 4, 043404, https://doi.org/10.1088/1742-5468/aab04c. Suche in Google Scholar

[80] A. Saxena, T. Tripathy, and R. Anguluri, Are the flows of complex-valued Laplacians and their pseudoinverses related?, 2025 European Control Conference (ECC), IEEE, 2025, pp. 653–658, https://doi.org/10.23919/ECC65951.2025.11186905. Suche in Google Scholar

[81] M. A. Shubin, Discrete magnetic Laplacian, Commun. Math. Phys. 164 (1994), no. 2, 259–275, https://doi.org/10.1007/BF02101702. Suche in Google Scholar

[82] D. A. Spielman, Spectral graph theory and its applications, Proc. 48th Annual IEEE Symposium on Foundations of Computer Science (FOCS), IEEE, 2007, pp. 29–38, https://doi.org/10.1109/FOCS.2007.56. Suche in Google Scholar

[83] Z. Stanić, Inequalities for Graph Eigenvalues, vol. 423, Cambridge University Press, Cambridge, UK, 2015, https://doi.org/10.1017/CBO9781316341308. Suche in Google Scholar

[84] T. Stephen, A majorization bound for the eigenvalues of some graph Laplacians, SIAM J. Discrete Math. 21 (2007), no. 2, 303–312, https://doi.org/10.1137/040619594. Suche in Google Scholar

[85] L. R. Varshney, B. L. Chen, E. Paniagua, D. H. Hall, and D. B. Chklovskii, Structural properties of the Caenorhabditis elegans neuronal network, PLoS Comput. Biol. 7 (2011), no. 2, e1001066, https://doi.org/10.1371/journal.pcbi.1001066. Suche in Google Scholar PubMed PubMed Central

[86] W. Xiao and I. Gutman, Resistance distance and Laplacian spectrum, Theoret. Chem. Accounts 110 (2003), 284–289, https://doi.org/10.1007/s00214-003-0460-4. Suche in Google Scholar

[87] Y. Yang and H. Zhang, Some rules on resistance distance with applications, J. Phys. A Math. Theoret. 41 (2008), no. 44, 445203, https://doi.org/10.1088/1751-8113/41/44/445203. Suche in Google Scholar

[88] G. F. Young, L. Scardovi, and N. E. Leonard, Robustness of noisy consensus dynamics with directed communication, Proc. 2010 American Control Conference, IEEE, 2010, pp. 6312–6317, https://doi.org/10.1109/ACC.2010.5531506. Suche in Google Scholar

[89] G. F. Young, L. Scardovi, and N. E. Leonard, Rearranging trees for robust consensus, Proc. 50th IEEE Conference on Decision and Control and European Control Conference, IEEE, 2011, pp. 1000–1005, https://doi.org/10.1109/CDC.2011.6161270. Suche in Google Scholar

[90] T. Zaslavsky, Signed graphs, Discrete Appl. Math. 4 (1982), no. 1, 47–74, https://doi.org/10.1016/0166-218X(82)90033-6. Suche in Google Scholar

[91] X. Zhang, Y. He, N. Brugnone, M. Perlmutter, and M. Hirn, Magnet: A neural network for directed graphs, Adv. Neural Inform. Proces. Syst. 34 (2021), 27003–27015, https://pmc.ncbi.nlm.nih.gov/articles/PMC9425115. Suche in Google Scholar

Received: 2025-07-14
Revised: 2025-10-22
Accepted: 2025-11-03
Published Online: 2025-12-15

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 8.1.2026 von https://www.degruyterbrill.com/document/doi/10.1515/spma-2025-0047/html
Button zum nach oben scrollen