• No results found

Potential robustness measures for digraphs

N/A
N/A
Protected

Academic year: 2021

Share "Potential robustness measures for digraphs"

Copied!
40
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

C. Kwok

Potential robustness measures for digraphs

Master thesis

Thesis advisor: Dr. F.M. Spieksma

Date master exam: 28 August 2015

Mathematisch Instituut, Universiteit Leiden

(2)
(3)

3

Contents

1 Introduction . . . 5

2 E ective resistance . . . 6

3 The (weighted) Laplacian L . . . 7

3.1 Properties of the Laplacian . . . 7

4 The Laplacian pseudoinverse L+ . . . 10

4.1 L+: Example undirected graph . . . 12

5 Di erent approach for the undirected graph . . . 14

5.1 Reduced Laplacian L . . . 14

5.2 Generalized inverse . . . 15

5.3 De ning X . . . 16

5.4 X: Total e ective resistance . . . 17

5.5 X: Example undirected graph . . . 18

6 Computing e ective resistance using determinants . . . 19

6.1 Example undirected graph using determinants . . . 22

7 L+: Directed graph . . . 24

8 X: Directed graph . . . 25

8.1 E ective resistance for the directed graph is well-de ned . . . 28

8.2 The square root of the e ective resistance is a metric . . . 29

8.3 X: Examples directed graph . . . 31

8.3.1 Example: Corresponding Laplacian of X . . . 33

9 Di erent de nition for X . . . 35

9.1 Examples corresponding to the di erent de nition of X . . . 36

10 Determinants: Directed graph . . . 37

11 Conclusion and discussion . . . 38

References . . . 40

(4)
(5)

1 Introduction 5

1 Introduction

Whenever a connection in a network fails, it could have grave consequences. It is important that even if these malfunctions occur, the network keeps on per- forming well. This ability is called robustness. The more a network is resistant to failures, the more robust it is. In this thesis, we will consider graphs as a representation of networks. While there has been a lot of research done for undirected graphs, there is little we know about directed graphs even though di- rected networks play a large roll in our world. Our aim is to nd a mathematical de nition and an approach to calculate the robustness of directed graphs such that it does not contradict with our sense of robustness. The most important property is that when an edge is added, the robustness should not decrease.

The same holds if an edge weight is increased. In literature, it is also desired that a direct connection is equal or more robust than an indirect connection (triangle inequality). If there is a way to calculate the robustness of a network, it is possible to compare networks and to design networks that are more robust.

First we will show how to calculate robustness for undirected graphs. This will be done by using the concept of e ective resistance between two nodes, which is explained in Section 2. In Section 3, the de nition of the Laplacian of a graph will be given and some properties of the Laplacian will be shown. In Section 4, we will show how the robustness is calculated by using the Laplacian pseu- doinverse. In the master thesis of W. Ellens [4], it is shown that the e ective resistance function has some nice properties. That is, the total e ective resis- tance of a graph strictly decreases (the graph is more robust) when an edge is added or weights are increased. Furthermore, the e ective resistance function is a metric (distance function). The proof of these properties will not be given in this thesis. It can be found in Chapter 4.3 in [4]. Another way to nd the Lapla- cian pseudoinverse will be shown in Section 5. This approach is obtained from [10]. In Section 6, an alternative expression of the e ective resistance between two nodes which is obtained from [1] will be given.

We will then analyze directed graphs. In Section 7, it is shown how the method used in Section 4 fails for directed graphs. In Section 8, we analyze [10] and their de nition of e ective resistance for directed graphs. This is followed by Section 9 in which we propose and analyze a di erent, but similar de nition. In Section 10, we will try to extend the expression found in Section 6 to directed graphs.

(6)

2 E ective resistance 6

2 E ective resistance

A graph is denoted by G = (V; E), where V the set of N vertices and E is the set of edges or arrows of G when G is undirected respectively directed. A weighted graph has weights on the edges. The weight of edge (i; j) is denoted with wij. We assume that all weights are positive. If no weight is written next to an edge, it is assumed that the edge weight is 1.

To determine the e ective resistance, the graph is viewed as an electrical circuit, where an edge (i; j) corresponds to a resistor Rij. For each pair of vertices the e ective resistance between these vertices can be calculated with the well- known series and parallel manipulations. Two edges, corresponding to resistors with resistance R1 and R2, connected in series can be replaced by one edge with e ective resistance R1+ R2. On the other hand, two resistors connected in parallel can be replaced by a resistor with resistance 1 1

R1+R21 . Notice that from these series and parallel manipulations, the e ective resistance takes both the number of paths and their length into account, intuitively measuring the presence and quality of back-up possibilities.

Now using Kirchho 's rst law and Ohm's law, an expression for the e ective resistance rab between vertices a and b is found. Kirchho 's rst law states that at any vertex in an electrical circuit, the sum of currents owing into that vertex is equal to the sum of currents owing out of that vertex. In other words, connect a voltage between vertices a and b and let I be the net current out of a and into b. Then the current yij between vertices i and j must satisfy

X

j2N(j)

yij = 8>

<

>:

I if i = a;

I if i = b;

0 otherwise,

(1)

where N(i) is the set of rst neighbors of vertex i. Now Ohm's law allows us to associate a potential vi to any vertex i, such that for all edges (i; j) we have

yijRij = vi vj: (2)

De nition 2.1. The e ective resistance between vertex a and b is de ned as rab=va vb

I :

De nition 2.2. The total e ective resistance for undirected graphs, also known as the Kirchho index Kf, is de ned as the sum of the e ective resistances over all pairs of vertices:

Kf =XN

i=1

XN j=i+1

rij:

Note that if one graph is more robust than another, the former's Kirchho index is smaller.

(7)

3 The (weighted) Laplacian L 7

3 The (weighted) Laplacian L

For every graph G, directed or undirected, we can nd the associated Laplacian matrix.

De nition 3.1. The Laplacian L is given by L = D A;

where D is the diagonal matrix of the node out-degrees and A is the adjacency matrix corresponding to G.

De nition 3.2. For a graph with non-negative edge weights wij, the analogue of the adjacency matrix is the matrix of weights W = (wij). The weighted Laplacian, Lw, is given by

Lw= S W;

where S is the diagonal matrix of strengths, with Sii=PN

j=1wij.

Notice the Laplacian matrix is the weighted Laplacian matrix where the weights are either zero or one. For notational convenience we will write L for the weighted Laplacian.

We de ne the resistance of an edge Rij as w1ij.

Theorem 3.1. The e ective resistance of an edge (i; j) is computed as follows rij= (e(i)N e(j)N )>Y (e(i)N e(j)N ) = Yii+ Yjj Yij Yji; (3) where Y 2 RNN acts as the inverse of the Laplacian matrix on 1?N and as the zero map on spf1Ng.

Proof. This proof is similar to the proof of Theorem 4.1 in [4].

We will see in Section 4 that the Laplacian pseudoinverse satis es these pro- perties mentioned in Theorem 3.1.

3.1 Properties of the Laplacian

In this section some properties of the Laplacian are shown.

Lemma 3.1. The (weighted) Laplacian corresponding to an undirected graph is symmetric, while the (weighted) Laplacian corresponding to a directed graph can be non-symmetric.

Proof. For undirected graphs, we have

(i; j) 2 E , (j; i) 2 E

Thus, Lij = Lji for all i; j 2 V . On the otherhand, if G is a directed graph, it is possible that (i; j) 2 E, while (j; i) =2 E i.e. Lij = wij and Lji= 0.

(8)

3 The (weighted) Laplacian L 8

Whether or not the Laplacian is symmetric, the sum of each row is always equal to zero, thus L1N = 0 i.e. 0 is always an eigenvalue of L and 1N its corresponding eigenvector.

The following lemma shows that the eigenvalues of the Laplacian L of an undi- rected graph are all positive.

Lemma 3.2. Let G be an undirected graph. The (weighted) Laplacian L of G is a positive semi-de nite matrix.

Proof. Taken from [4]. First convert G in a directed graph, by choosing an arbitrary direction for each edge in G. Now de ne an arc-vertex incidence matrix M, i.e. for an arc a and a vertex v:

Mav = 8>

<

>:

pwij if a = (i; j) and v = i;

pwij if a = (i; j) and v = j;

0 otherwise.

Then the Laplacian L satis es L = M>M and we have for all z 2 RN : z>Lz = z>M>Mz

= (Mz)>(Mz)  0;

where the last inequality holds, since it is the inner product of the vector Mz with itself.

Lemma 3.3. Let G be a directed graph. Then all eigenvalues of the Laplacian L of G have non-negative real part.

Proof. It follows from Gerschgorin circles [8] that all eigenvalues of L have non- negative real part.

As mentioned before, zero is always an eigenvalue of L. The algebraic multi- plicity of zero is given in the next theorem. However, we introduce a de nition rst.

De nition 3.3. A component C of a directed graph G is called a strongly connected component if for all vertices i; j of C, there is a directed path from i to j and a directed path from j to i. A directed path from i 6= j to j is a sequence of arcs which connects a sequence of vertices starting from i and ending in j such that all arcs have the same direction. Assume that there is always a directed path from i to i for every i 2 G. A strongly connected closed component C is a strongly connected component such that for all vertices j not belonging to C we have wij = 0 for all vertices i in C.

Theorem 3.2. Let G be an undirected graph. The algebraic multiplicity of the eigenvalue zero of the (weighted) Laplacian L of G is equal to the number of connected components of G. If G were a directed graph, then it is equal to the number of strongly connected closed components of G.

(9)

3 The (weighted) Laplacian L 9

Proof. The following proof is from [4].

Let G be an undirected graph and L its (weighted) Laplacian. For every con- nected component C, let yC2 Rnbe the vector with yiC= 1 if i is a vertex of C and yCi = 0 otherwise. Notice that yC is an eigenvector of L with eigenvalue 0, since LyC = 0. Furthermore, the set of these eigenvectors is linearly indepen- dent. Thus, we will prove that all eigenvectors with eigenvalue zero are linear combinations of these eigenvectors.

Now suppose x is an eigenvector of L with eigenvalue zero, i.e. Lx = 0, then we have for all i 2 V

xi

Xn j=1

wij = Xn j=1

xjwij: (4)

Now let C be an arbitrary component and i such that xi = maxj2Cxj, then using (4) we nd

xiXn

j=1

wij =Xn

j=1

xjwij Xn

j=1

xiwij = xiXn

j=1

wij:

Thus xi = xj for all neighbours j of i, since in that case we have wij > 0.

Similarly, we see that xj = xk for every neighbor k of j. By repeating this argument, we see that all eigenvectors x have xi= xj when i and j are part of the same component. Therefore, all eigenvectors with eigenvalue zero are linear combinations of the vectors yC.

If G were a directed graph, we let C be a strongly connected closed component and then the proof still holds.

For example, notice in Example 8.1 there is only one strongly connected closed component, that is, f3g.

(10)

4 The Laplacian pseudoinverse L+ 10

4 The Laplacian pseudoinverse L

+

As mentioned before there are multiple ways to calculate the e ective resistance for undirected graphs. In [4] it is shown how one can calculate the e ective resistance for undirected graphs using the Laplacian. Leaving the proofs aside, we will repeat some of the results.

Since the Laplacian has an eigenvalue zero, it is not invertible. However, if we restrict the linear transformation to the subspace orthogonal to the null space spf1Ng, the matrix can be inverted. Let L+be the matrix that corresponds to this inverse on 1?N and to the zero map on spf1Ng. In other words:

De nition 4.1. The Laplacian pseudoinverse L+ is de ned as the matrix sa- tisfying

L+1N = 0 and for every w?1N:

L+w = v , Lv = w and v?1N:

This de nition is a speci c case of the Moore-Penrose pseudoinverse for general mn matrices provided that the graph is undirected. In Section 5.2, it is shown that it is unique.

One way to construct the Laplacian pseudoinverse is as follows. Since the Lapla- cian is symmetric, it has an orthonormal set of eigenvectors v1; : : : ; vN. Let U be the matrix that has these eigenvectors as its columns, consequently U 1= U>, and let D be the matrix with the corresponding eigenvalues on its diagonal and zero elsewhere. Notice that i= 0 for some i = 1; : : : ; N, because 0 is an eigen- value of L. Without loss of generality, assume 1 = 0. Furthermore i  0 for all i, since L is a positive semi-de nite matrix, see Lemma 3.2. Thus, D is of the form:

D = 0 BB B@

0 0 : : : 0 0 2 : : : 0 ... ... ... ...

0 0 : : : N

1 CC CA:

Then the Laplacian satis es L = UDU 1= UDU>. The pseudoinverse is then as follows

L+= UD+U 1= UD+U>; with D+:

D+= 0 BB B@

0 0 : : : 0 0 12 : : : 0 ... ... ... ...

0 0 : : : 1N 1 CC CA:

Furthermore, L+ is similar to D+, thus the eigenvalues of L+ are 0;12; : : : ;1n. Also, L+ is symmetric:

(L+)>= (UD+U>)>= U(D+)>U>

= UD+U>= L+

(11)

4 The Laplacian pseudoinverse L+ 11

Lemma 4.1. The Laplacian pseudoinverse L+ satis es L+=

Z 1

0



e Lt 1 N1N1>N

 dt:

Proof. Let L = UDU 1 with U and D as above. Note the rst column of U is

p1

N; : : : ;p1N>

and we have an orthonormal set of eigenvectors. Observe:

e Lt=X1

k=0

( 1)kLktk k! =X1

k=0

( 1)k(UDU 1)ktk k!

=X1

k=0

( 1)kUDkU 1tk

k! = UX1

k=0

( 1)kDktk

k! U 1

= U 0 BB B@

1 0 : : : 0

0 e 2t : : : 0 ... ... ... ...

0 0 : : : e Nt 1 CC CAU 1

= 0 B@

N1 : : : N1 ... ... ...

N1 : : : N1 1

CA + e 2tv2>v2+ : : : + e NtvN>vN

Thus we nd Z 1

0



e Lt 1 N1N1>N

 dt =

Z 1

0 e 2tv2>v2+ : : : + e Ntv>NvN dt

= 1

2v2>v2+ : : : + 1

NvN>vN

= UD+U 1= L+:

Since the Laplacian pseudoinverse is given by L+ = UD+U>, it acts as the inverse of the Laplacian matrix on 1?N and as the zero map on spf1Ng. This leads to the following corollary.

Corollary 4.1. The e ective resistance of an edge (i; j) is computed as follows rij = (e(i)N e(j)N )>L+(eN(i) e(j)N ) = L+ii+ L+jj 2L+ij:

Now the total e ective resistance is given by the following theorem.

Theorem 4.1. The Kirchho index Kf satis es

Kf = N XN

i=2

1

Li ; (5)

with Li's the eigenvalues of L.

(12)

4 The Laplacian pseudoinverse L+ 12

Proof.

Kf =XN

i=1

XN j=i+1

rij =1 2

XN i=1

XN j=1

rij

=1 2

XN i=1

XN j=1

(L)+ii+ (L)+jj 2(L)+ij

= NXN

i=1

(L+ii) 1>NL+1N

= N Tr(L+) = NXN

i=2

1

Li :

4.1 L

+

: Example undirected graph

Using the Laplacian pseudoinverse, the total e ective resistance can be calcu- lated. Observe a rather easy example.

Example 4.1. Let G be an undirected graph with weights 1:

2

1 3 with the Laplacian matrix L =

0

@ 1 1 0

1 2 1

0 1 1

1 A :

The eigenvalues of L are 1 = 0; 2 = 1; 3 = 3. Since L is symmetric, it has an orthonormal set of eigenvectors, explicitly

v1=

1 3

p3;1 3

p3;1 3

p3



; v2=

 1 2

p2; 0;1 2

p2



; v3=

1 6

p6; 1 3

p6;1 6

p6

 :

Let U be the matrix that has these eigenvectors as its columns:

U = 0

@

13

p3 12p

2 16p

1 6

3

p3 0 13p

1 6

3

p3 12p

2 16p 6

1

A , U 1= U>= 0

@

13

p3 13p 3 13p

1 3

2

p2 0 12p

1 2

6

p6 13p 6 16p

6 1 A : Then let D be the matrix with the eigenvalues on the diagonal and zero else- where:

D = 0

@ 0 0 0 0 1 0 0 0 3

1 A :

(13)

4 The Laplacian pseudoinverse L+ 13

Note L = UDU 1= UDU>. Then D+ is given by

D+= 0

@ 0 0 0 0 1 0 0 0 13

1 A ; and so the Laplacian pseudoinverse L+ is

L+= UD+U 1= UD+U>= 1 9

0

@ 5 1 4

1 2 1

4 1 5

1 A :

We have L+1 = 0 and for every w?1 :

L+w = v such that Lv = w and v?1:

For example take w = 0

@ 1 01

1

A, then v = 19 0

@ 6 33

1

A and Lv = w.

The total e ective resistance of this graph is Kf = 3  1 +13

= 4.

(14)

5 Di erent approach for the undirected graph 14

5 Di erent approach for the undirected graph

In this section, a di erent way to construct the Laplacian pseudoinverse is shown. This matrix, which we will refer to as X in this section, is thus identical to the Laplacian pseudoinverse L+for undirected graphs. However, in Section 8 the matrix X will be rede ned such that it is unique and symmetric for directed graphs, while acting as the Laplacian pseudoinverse for undirected graphs. Ef- fective resistance is then de ned using the rede ned X. Consequently, the square root of the e ective resistance is a metric on the nodes of any connected directed graph. The following results are from [10].

Let  = IN 1

N1N1>N, denote the orthogonal projection matrix onto the sub- space of RN perpendicular to 1N. This subspace will be denoted with 1?N.  is a symmetric matrix. Since the entries of the rows of L add up to 0 (i.e. the rows are perpendicular to 1), we have L1N = 0, L = L and L> = L> for any graph. Note, if the graph is balanced, i.e. the out-degree and in-degree are equal, which is true for undirected graphs, L = L>= L>= L also hold.

Let Q 2 R(N 1)N be a matrix whose rows form an orthonormal basis for 1?N. Thus, we require h

p1

N1N Q>i

to be an orthogonal matrix. It follows,

Q1N = 0; QQ>= IN 1 and Q>Q = : (6) The last result holds since h

p1

N1N Q>i

is an orthogonal matrix, thus

p1

N1N Q>

 p1

N1N Q>

>

= IN

and by adding the p1N1N column to Q, we essentially add N11N1>N to Q>Q.

X will be constructed such that X is equal on 1?N to the inverse of L and 0 otherwise. Thus, it satis es the de nition of the Laplacian pseudoinverse in Section 4. This can be achieved by taking X as the unique generalized inverse of L.

5.1 Reduced Laplacian L

We construct X by using the reduced Laplacian. This matrix acts as the Lapla- cian L on 1?N.

Let b1; : : : ; bN 1be an orthonormal basis for 1?N and let Q 2 R(N 1)N be the matrix formed with these basis vectors as rows. Then for any v 2 RN, we can write v =PN 1

i=1 cibi+ cN1N, with c1; : : : ; cN 2 R. By taking the innerproduct with any bi on both sides, we nd bi v = ci. Thus, v := Qv is a coordinate vector (with respect to the chosen basis) of the orthogonal projection of v onto 1?N. For any M 2 RNN, let M 2 RN 1N 1be the linear transformation with respect to b1; : : : ; bN 1 with M := QMQ>. Then M acts as M on 1?N, i.e.

Mv = (QMQ>)Qv = QMv = QMv = Mv for any v 2 1?N:

(15)

5 Di erent approach for the undirected graph 15

Thus, on 1?N, the Laplacian matrix L is equivalent to

L = QLQ>; (7)

which we refer to as the reduced Laplacian. We see that L is symmetric if and only if the graph is balanced, since in that case we have L = L>:

L> = (QLQ>)>= QL>Q>= QLQ>= L:

Theorem 5.1. Let L be the Laplacian matrix, then L given in (7) has the same eigenvalues, except for a single zero eigenvalue, as L. Also, if v is an eigenvector with eigenvalue 6= 0 of L, then v on 1?N is an eigenvector with eigenvalue of L.

Proof. This proof is partially taken from [9].

De ne the matrix

V =

 1

pN1>N Q



; (8)

with Q satisfying (6). Then V V>= IN and V>V = IN hold, which means V>

is the inverse of V , i.e. V> = V 1: Now we see that L is similar to V LV> for the invertible matrix V , thus L and V LV> have the same eigenvalues. Using the fact L1N = 0, it follows V LV> =

 0 0 0 L



with L given in (7) and we see that this is a block matrix. Thus, the eigenvalues of V LV>, which are the same as the eigenvalues of L, are the solutions of   det(IN 1 L) = 0, i.e.

zero and the eigenvalues of L. We conclude that L has the same eigenvalues as L except for a single zero eigenvalue.

Let v be an eigenvector of L with eigenvalue . Then, we have Lv = QLQ>Qv = QLv

= QLv = Q v = v : Thus, v is an eigenvector with eigenvalue of L.

Hence, it follows from Theorems 3.2 and 5.1 that for a connected graph, L is invertible since L only has one 0 eigenvalue.

5.2 Generalized inverse

De nition 5.1. For A 2 Rmn, the generalized inverse of A is de ned as a matrix A+2 Rnm satisfying the so-called Moore-Penrose criteria:

1. AA+A = A 2. A+AA+= A+ 3. (AA+)>= AA+ 4. (A+A)>= A+A

(16)

5 Di erent approach for the undirected graph 16

Theorem 5.2. The generalized inverse of A 2 Rmn is unique.

Proof. Let B; C 2 Rnm both be a generalized inverse of A 2 Rmn, thus they satisfy the Moore-Penrose criteria. Then observe that

AB = (AB)>= B>A> = B>(ACA)> = B>A>C>A>

= (AB)>(AC)>= ABAC = AC:

Analogously we nd BA = CA. It follows,

B = BAB = BAC = CAC = C;

thus the generalized inverse of A 2 Rmn is unique.

As mentioned before, X will be constructed as the unique generalized inverse of L. Thus, it must satisfy the Moore-Penrose criteria. This can also be formulated di erently.

Lemma 5.1. If X satis es the following:

XL = LX =  and X = X = X; (9)

then X satis es the Moore-Penrose criteria.

Proof. Let X be a solution to (9). Since L is symmetric, we have:

1. LXL = L = L 2. XLX = X = X 3. (LX)>= >=  = LX 4. (XL)>= >=  = XL:

Thus it satis es the Moore-Penrose criteria.

5.3 De ning X

Now using the reduced Laplacian L we give an explicit construction for X:

X = Q>L 1Q: (10)

Observe:

 X1N = Q>L 1Q1N = 0, since Q1N = 0;

 Xv = QXv = QQ>L 1Qv = L 1Qv = L 1v for any v 2 RN:

Thus, X acts as the inverse of L, which is equivalent to the inverse of L on 1N

and X is 0 elsewhere.

Corollary 5.1. Let X = [xi;j], then the e ective resistance is given by

rij= (e(i)N e(j)N )>X(e(i)N e(j)N ) = xi;i+ xj;j xi;j xj;i: (11)

(17)

5 Di erent approach for the undirected graph 17

Since L is symmetric for undirected graphs, so is L 1 and thus X is also sym- metric:

X> = (Q>L 1Q)>= Q>(L 1)>Q = Q>L 1Q = X Consequently, we can write

rij = xi;i+ xj;j 2xi;j:

Notice that we always have rij= rji, even if X is not symmetric.

Lemma 5.2. X constructed as in (10) satis es (9) when the graph is undirected and is thus the Moore-Penrose generalized inverse of L. Hence, it is unique.

Proof. Using (6) and L = L = L, we nd

XL = XL = Q>L 1QLQ>Q = Q>L 1LQ = Q>Q =  and analogously we nd LX = . Furthermore,

X = Q>L 1Q = Q>L 1QQ>Q = Q>L 1Q = X and again in the same way X = X also holds.

Note L is not unique, since it depends on the choice of Q. However X is independent of the choice of Q as long it satis es (6).

Theorem 5.3. X is independent of the choice of Q.

Proof. Let Q and Q0 both satisfy (6). De ne P := Q0Q>. P is orthogonal:

P>P = (Q0Q>)>Q0Q>= QQ0>Q0Q>= QQ>= IN 1

and analogously we have P P>= IN 1, thus P> is the inverse of P . Write Q0= Q0 = Q0Q>Q = P Q:

Hence,

X0 := Q0>(Q0LQ0>) 1Q0= Q>P>(P QLQ>P>) 1P Q

= Q>(QLQ>) 1Q = X:

Thus X is independent of the choice of Q.

5.4 X: Total e ective resistance

Theorem 5.4. The Kirchho index satis es

Kf = NXN

i=2

Xi

with Xi 's the eigenvalues of X.

(18)

5 Di erent approach for the undirected graph 18

Proof. Take the matrix V as in (8). Then X is similar to V XV> for V , thus X and V XV> have the same eigenvalues. Since X = Q>L 1Q, we can write V XV> as the following: V XV> =

 0 0

0 L 1



: Thus, X has a single 0 eigenvalue and its remaining eigenvalues are the same as L 1. However, the eigenvalues of L 1 are one over the eigenvalues of L and those are the same as the eigenvalues of L, excluding the zero eigenvalue. This implies that X has a single 0 eigenvalue and its remaining eigenvalues are one over the eigenvalues of L. Then using (5) we have

Kf = N XN

i=2

1

Li = N XN i=2

Xi

with Li and Xi the i'th eigenvalue of L respectively X.

5.5 X: Example undirected graph

We will now show the calculation of X corresponding to Example 4.1. Take

Q = p12 p12 0

p1

6 p1

6 p2

6

! , Q>=

0 B@

p1

2 p1

1 6 p2 p1

0 p266

1 CA :

Then we nd

L = QLQ>= p352 p312

12 9

6

!

and L 1= 1 6

 3 p

p 3

3 5

 :

X is given by X = 19 0

@ 5 1 4

1 2 1

4 1 5

1

A, which is the same as L+ and we also have Kf = 3  1 + 13

= 4.

(19)

6 Computing e ective resistance using determinants 19

6 Computing e ective resistance using determinants

As we have seen in Section 4, the e ective resistance between node i and j is given by

rij = L+ii+ L+jj L+ij L+ji:

We will show another expression for rij that is quite simple. This result holds for any weighted Laplacian of an undirected graph.

First de ne L(i) to be the submatrix obtained by deleting the i-th row and i-th column of the Laplacian matrix L of a graph G. The submatrix L(i; j) is obtained by deleting the i-th and j-th rows and i-th and j-th columns of L. The following theorem and proof are from [1].

Theorem 6.1. Let G = (V; E) be an undirected connected graph with n  3 vertices, and 1  i 6= j  n. Let L(i) and L(i; j) be the above de ned submatrices of the Laplacian matrix of G. Then the e ective resistance between node i and j is given by

rij= det L(i; j)

det L(i) : (12)

Proof. Associate to each node vi of the graph G a variable xi and de ne the auxiliary function

f(x1; : : : ; xn) = X

k<l;l2N(k)

(xk xl)2 (13)

where N(k) denotes the set of rst neighbors of vk. The following relation is a result known in the theory of electrical networks [2]. For i 6= j,

rij = max

 1

f(x1; : : : ; xn)

xi= 1; xj= 0; 0  xk 1; k = 1; : : : ; n



: (14) Then using (13) and (14), we derive for k 6= i; j :

@f

@xk = 2 X

l2N(k)

(xk xl) = 0: (15)

Since the vertices of G are numbered arbitrarily, without loss of generality, we may consider the special case in which i = n 1 and j = n. Thus xn 1= 1 and xn = 0 may hold. Let A = (aij) be the adjacency matrix of G and denote the submatrix L(n 1; n) with Ln 2, then we can write the Laplacian matrix as

L =

 Ln 2 B B> D



; where we have

B = 0 B@

a1;n 1 a1;n

... ...

an 2;n 1 an 2;n 1

CA and D = PN

i=1an 1;i an 1;n

an;n 1 PN

i=1an;i

! :

(20)

6 Computing e ective resistance using determinants 20

Then from (15) the following can be derived for all k 6= n 1; n:

X

l2N(k)

(xk xl) = X

l2N(k)nfn 1;ng

(xk xl) + (xk xn 1)1(k)n 1+ (xk xn)1(k)n

= 0; (16)

with

1(k)n 1=

(1 if vn 1 is adjacent to vk, 0 otherwise:

and 1(k)n similar. Now since xn 1= 1 and xn= 0, we have X

l2N(k)nfn;n 1g

(xk xl) + xk1n 1(k) + xk1(k)n = 1  1(k)n 1

Thus Ln 2x = b; (17)

where x = (x1; : : : ; xn 2)> and b = (b1; : : : ; bn 2)>, with bk= 1 if the vertices vn 1 and vk are adjacent, otherwise bk = 0, for all k = 1; : : : ; n 2.

Since xn 1= 1; xn= 0 holds and L is symmetric, the following relations hold:

x>B(xn 1; xn)>= (xn 1; xn)B>x =

n 2X

i=1

an 1;ixi= X

k2N(n 1)

xk

and (xn 1; xn)D(xn 1; xn)>= dn 1;

with dn 1denoting the degree of the vertex vn 1. Now rewrite (13) as follows f(x1; : : : ; xn) = (x>; xn 1; xn)L(x; xn 1; xn)

= x>Ln 2x + (xn 1; xn)B>x + x>B(xn 1; xn)>

+ (xn 1; xn)D(xn 1; xn)>

= x>b 2 X

k2N(n 1)

xk+ dn 1

= dn 1

X

k2N(n 1)

xk;

(18)

where the last equation holds because x>b =

n 2X

i=2

an 1;ixi= X

k2N(n 1)

xk:

Since G is an undirected, connected graph, its Laplacian matrix L is positive semi-de nite. Thus, the submatrix Ln 2 is a positive de nite matrix, which means its inverse (Ln 2) 1 exists. Denote the (i; j)-th entry of (Ln 2) 1 with tij, then x = (Ln 2) 1b implies

xk= X

l2N(n 1)

tkl: (19)

We will use the following lemma.

(21)

6 Computing e ective resistance using determinants 21

Lemma 6.1. The following equation holds:

det L(i; j) = det L(j; i):

Proof. We have

L(i; j) = (L(j; i))>and detL(i; j) = det(L(i; j))>: Thus, it follows

detL(i; j) = detL(j; i):

Let Ln 2(k; l) be the submatrix obtained by removing from Ln 2the k-th row and the l-th row. Then using Cramer's rule and Lemma 6.1 we obtain

tkl= ( 1)k+ldetLn 2(l; k)

detLn 2 = ( 1)k+ldetLn 2(k; l)

detLn 2 : (20) Next we can rewrite (18) using (19) and (20) into

f(x1; : : : ; xn) = dn 1

X

k2N(n 1)

X

l2N(n 1)

( 1)k+ldetLn 2(k; l)

detLn 2 : (21) Now look at detL(n) by expanding it with respect to its last column, which is the (n 1)-th column of L:

detL(n) = dn 1detLn 2 X

k2N(n 1)

( 1)k+n 1detL(n; k; n 1); (22)

where L(n; k; n 1) is the submatrix obtained by removing the k-th row and the (n 1)-th column from L(n). Furthermore, we expand detL(n; k; n 1) with respect to its last row,

detL(n; k; n 1) = X

l2N(n 1)

( 1)l+n 1detLn 2(k; l):

Substituting this in (22) yields detL(n) = dn 1detLn 2 X

k2 (n 1)

( 1)k+n 1 X

l2N(n 1)

( 1)l+n 1detLn 2(k; l)

= dn 1detLn 2

X

k2N(n 1)

X

l2N(n 1)

( 1)k+ldetLn 2(k; l); (23)

which results in the following expression when substituted back in (21) f(x1; : : : ; xn) = detL(n)

detLn 2: Finally, substituting this in (14), gives us (12).

(22)

6 Computing e ective resistance using determinants 22

According to Kirchho 's theorem, the number of spanning trees is equal to any cofactor of the Laplacian matrix of G. Thus

det L(i) = t(G);

where t(G) is the number of spanning trees of G. Now, according to [7], det L(i; j) is equal to the number of trees of G0, where G0 is the graph where vertex vi and vertex vj is contracted to one vertex v. Thus,

detL(i; j) = t(G0) and we can write

rij= t(G0) t(G):

This can be interpreted as follows. The e ective resistance between vertex i and j is equal to the number of trees of G containing the edge (i; j) divided by the total number of spanning trees of G. This can be seen as how much the edge (i; j) contributes to the number of spanning trees of G.

Remark: Since any cofactor of the Laplacian is equal to the number of spanning trees, we have detL(i) = detL(j) for every i; j 2 V . Using Lemma 6.1, we can write

rij= det L(i; j)

det L(i) = det (L(j; i))>

det L(j) =det L(j; i) det L(j) = rji:

6.1 Example undirected graph using determinants

Recall Example 4.1:

2

1 3 with the Laplacian matrix L =

0

@ 1 1 0

1 2 1

0 1 1

1 A :

Using the corresponding Laplacian pseudoinverse of this graph 1

9 0

@ 5 1 4

1 2 1

4 1 5

1 A ;

yields

r12= L+11+ L+22 2L+12= 5 9 +2

9 +2 9 = 1;

r13= L+11+ L+33 2L+13= 5 9 +5

9 +8 9 = 2;

r23= L+22+ L+33 2L+23= 2 9 +5

9 +2 9 = 1:

(23)

6 Computing e ective resistance using determinants 23

These are the same as using (12):

r12= detL(1; 2) detL(1) = 1;

r13= detL(1; 3) detL(1) = 2;

r23= detL(2; 3) detL(2) = 1:

(24)

7 L+: Directed graph 24

7 L

+

: Directed graph

It is not always possible to compute L+ for directed graphs as in Section 4 for undirected graphs. The matrix U such that L = UDU 1 can not always be constructed.

Example 7.1. Consider 2

1 3 with L =

0

@ 1 1 0

0 1 1

0 0 0

1 A :

The eigenvalues of L are 1= 0 with multiplicity two and 2 = 1 with multi- plicity one. The eigenvectors are

v1=

1 3

p3;1 3

p3;1 3

p3



and v2= (1; 0; 0) :

This means that L is not diagonalizable; we can't construct U such that L = UDU 1. Furthermore, even if we were to restrict the graphs, in [6] it has been argued that the robustness if de ned through L+could decline when an edge is added, which contradicts with our concept of robustness. Thus, this approach is not suitable for directed graphs.

(25)

8 X: Directed graph 25

8 X: Directed graph

Here as well, for directed graphs we can't always compute X in the same manner as in Section 5. From Theorem 5.1 it follows that in the case of a directed graph, the algebraic multiplicity of the eigenvalue zero is equal to the number of strongly connected closed components, which is not always equal to one.

Therefore, the inverse of L does not necessarily need to exist.

Therefore in [10], the authors propose a di erent way to construct X when the graph is directed. This approach will lead to a matrix X that is unique and symmetric. Furthermore, it is equal to the Laplacian pseudoinverse when calculated for undirected graphs and in Section 8.2 it is shown that the square root of the e ective resistance is a metric on the nodes of any connected directed graph. The symmetry of X will be used to prove this. If we would take X as the Moore-Penrose inverse neither the e ective resistance nor the square root of it is a metric.

In section Conclusion and discussion, we will discuss the e ective resistance function satisfying symmetry.

The approach used in [10] considers the following Lyapunov equation:

A + A + B = 0; (24)

with A; B square matrices, A the conjugate transpose of A and  unknown.

Theorem 8.1. If A has only eigenvalues with a negative real part, then

 = Z 1

0 eAtBeAtdt (25)

is the unique solution to (24).

Proof. This proof is taken from [3].

Let  be given as in (25). Then we have A + A =

Z 1

0

d dt

heAtBeAti

dt = eAtBeAt 1

0 = B;

which means that (25) is a solution for (24). To prove the uniqueness of (25), we de ne a linear map : Cnn ! Cnn, with () = A + A. Now for all B 2 Cnn, the equation () = B has a solution. Thus, the dimension of the image of is n2. Note that the dimension of the domain of is also n2. Therefore, ker( ) = 0 holds. We conclude () = B has a unique solution for each B.

Lemma 8.1. Furthermore, if B is positive de nite, then  is also positive de nite.

Proof. Let z 2 Cn and z the conjugate transpose of z, then observe zz = z

Z 1

0 eAtBeAtdtz = Z 1

0 zeAtBeAtzdt (26)

= Z 1

0 (eAtz)B eAtz

dt > 0: (27)

(26)

8 X: Directed graph 26

The last inequality follows since B is positive de nite.

Now by taking A = L> and B = IN 1, (24) can be rewritten to

L + L> = IN 1: (28)

It follows from Lemma 8.1 that  is positive de nite. We will see that  is also symmetric.

Lemma 8.2.  satisfying (28) is symmetric.

Proof. Let  satisfy (28). Then from Theorem 8.1 it follows that

 = Z 1

0 e Lte L>tdt:

Observe the following:

>= Z 1

0

e Lte L>t>

dt = Z 1

0 e Lte L>tdt = :

Thus,  is symmetric.

Construct X as

X = 2Q>Q; (29)

with Q satisfying (6) and  the unique solution to the Lyapunov equation (28).

Note that since  is symmetric, X is also symmetric for any graph. Now if L is symmetric, we have

 =1

2L 1: (30)

Corollary 8.1. If L is symmetric, X in (29) is equal to the Laplacian pseu- doinverse L+.

Proof. Straightforward.

Corollary 8.2.  satis es

 = Z 1

0 Qe Lte L>tQ>dt:

Proof. Let  satis es (25) with A = L> and B = IN 1. Then we can write

(27)

8 X: Directed graph 27

the following:

 = Z 1

0 e Lte L>tdt

= Z 1

0

X1 k=0

( 1)k(QLQ>)ktk k!

X1 k=0

( 1)k(QL>Q>)ktk

k! dt

= Z 1

0 QX1

k=0

( 1)kLktk

k! Q>QX1

k=0

( 1)k(L>)ktk k! Q>dt

= Z 1

0 QX1

k=0

( 1)kLktk k! X1

k=0

( 1)k(L>)ktk k! Q>dt

= Z 1

0 Qe Lte L>tQ> 1 NQ

X1 k=0

( 1)kLktk

k! 1N1>Ne L>tQ>dt

= Z 1

0 Qe Lte L>tQ> 1

NQ1N1>Ne L>tQ>dt

= Z 1

0 Qe Lte L>tQ>dt:

The following lemma shows us that what kind of eigenvalues X has.

Lemma 8.3. X satisfying (29) has a single 0 eigenvalue and its remaining eigenvalues are twice those of .

Proof. Take V = h

p1

N1N Q> i

with Q as in Section 5. Then V is an or- thogonal matrix. X is similar to V>XV =

 0 0>

0 2



. Hence, X has a single 0 eigenvalue and its remaining eigenvalues are twice those of .

Now in [10] they de ne the e ective resistance between nodes in the graph as the following.

De nition 8.1. Let G be a connected graph with N nodes and Laplacian matrix L. Then the e ective resistance between nodes i and j in G is de ned as

rij = (e(i)N e(j)N )>X(eN(i) e(j)N ) = xi;i+ xj;j 2xi;j; (31) where

X = 2Q>Q; L + L>= IN 1; L = QLQ>; (32) and Q satisfying (6).

Remark: This expression for the e ective resistance comes from (3). However X de ned as in (32) does not satisfy all the properties mentioned in Theorem 3.1. X acts as the zero map on spf1Ng, but it does not act as the inverse of the Laplacian matrix on 1?N; XL =  does not hold.

By taking (31) as the de nition for the e ective resistance between two nodes, the de nition of the Kirchho index for undirected graphs can be generalized

(28)

8 X: Directed graph 28

by summing over all distinct e ective resistances of pairs of vertices of a graph, directed or undirected.

De nition 8.2. The total e ective resistance, Kf, of a graph is the sum over all distinct e ective resistances of pairs of vertices:

Kf= XN i=1

XN j=i+1

rij (33)

Theorem 8.2. The Kirchho index satis es

Kf = N XN i=1

Xi ;

with Xi 's the eigenvalues of the matrix X satisfying (32).

Proof. Similarly to the proof of Theorem 4.1, the Kirchho index can be written as

Kf =XN

i=1

XN j=i+1

rij= NXN

i=1

(xi;i) 1>NX1N

= N XN

i=1

(xi;i) 1>N(2Q>Q)1N = N XN i=1

xi;i

= N Tr(X) = NXN

i=1

Xi ;

with Xi 's the eigenvalues of X.

Then from Lemma 8.3 it follows

Kf = 2N

N 1X

i=1

i;

with i's the eigenvalues of .

8.1 E ective resistance for the directed graph is well-de ned

To con rm that the concept of e ective resistance for directed graphs given in De nition 8.1 is indeed well-de ned, observe the following two lemmas from [10].

Lemma 8.4. The value of the e ective resistance between two nodes in a con- nected directed graph is independent of the choice of Q.

Proof. Let Q an Q0be two matrices satisfying (6) and let the e ective resistance between nodes i and j be rij, r0ij, computed using Q and Q0 respectively. Fur- thermore, de ne W = Q0Q>. Then we have W Q = Q0Q>Q = Q0 = Q0 and

(29)

8 X: Directed graph 29

W W>= Q0Q>(Q0Q>)> = Q0Q>QQ0>= Q0Q0>= Q0Q0>= IN 1. W>W is computed analogously. Now using (7) and the above we can write

L0= Q0LQ0>= W QL(W Q)>= W QLQ>W> = W LW>:

Substituting this expression in (28) for L0, we see  = W>0W . And thus we nd X = 2Q>Q = 2Q>W>0W Q = 2Q0>0Q0= X0, hence rj;k= rj;k0 . From this lemma it follows that the e ective resistance is independent of the choice of Q as long as it satis es (6). The e ective resistance is also independent of the labelling of the nodes, which will be proven in the following lemma.

Lemma 8.5. The value of the e ective resistance between two nodes in a con- nected directed graph is independent of the labelling of the nodes.

Proof. Let L and L0 be two Laplacian matrices associated with the same graph, but with di erent labellings of the nodes. Then L0 can be found from L by permuting its rows and columns. That is, there exists a matrix P such that L0 = P LP>. Note that since P is a permutation matrix, P has exactly one 1 in every row and column, and every other entry is 0. Consequently, we have P 1= P>; P 1N = 1N and 1>NP = 1>N. Furthermore, we have

P>= (QP Q>)>= QP>Q> = QP 1Q>= P 1; P Q = QP Q>Q = QP  = QP (IN 1

N1N1>N)

= QP QP 1

N1N1>N = QP

and analogously Q>P = P Q>. Now using (7) and the properties above, it follows

L0= QL0Q>= QP LP>Q>= P QLQ>P> = P LP>:

Then substituting this expression in (28) for L0 yields  = P>0P . Hence, X = 2Q>Q = 2Q>P>0P Q = 2(P Q)>0QP

= 2P>Q>0QP = P>X0P:

Thus, if P permutes node i to node k and node j to node l, r0kl= rij holds.

8.2 The square root of the e ective resistance is a metric

The following theorem is taken from [10].

Theorem 8.3. The square root of the e ective resistance is a metric on the nodes of any connected directed graph. That is, we have for all nodes i; j; k

rij 0;

rij = 0 , i = j;

rij= rji; and prik+ prkj prij:

Referenties

GERELATEERDE DOCUMENTEN

You might have noticed that this page is slightly scaled to accommodate its content to the slide’s

If they have a common factor, divide both by their greatest common divisor.. Pete Agoras Some

The spectral excess theorem states that, in a regular graph Γ, the average excess, which is the mean of the numbers of vertices at maximum distance from a vertex, is bounded above

Simulations: Monte Carlo simulations were performed to compare a simple staircase method, PSI method and a random staircase method. A stochastic psychophysical model was applied

This statistic was surprising as there is continuously an increase in the number of opportunities available to BEE (black economic empowerment) candidates. 5 students that

In het eerste halfjaar van 1978 nam het aantal verkeersdoden onder bestuurders van personenauto's t.o.V. Het aantal overleden passagiers van per- sonenauto's was

A suitable homogeneous population was determined as entailing teachers who are already in the field, but have one to three years of teaching experience after

If a plant R can be arbitrarily pole assigned by real memoryless output feedback in the sense of Definition 2.3.1, then in particular does there exist a regular feedback law (2.29)