26.07.2013 Views

Variational Convergence of Finite Networks

Variational Convergence of Finite Networks

Variational Convergence of Finite Networks

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Interdisciplinary Information Sciences, Vol. 12, No. 1, pp. 57–70 (2006)<br />

<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong><br />

Atsushi KASUE<br />

Department <strong>of</strong> Mathematics, Kanazawa University, Kanazawa 920-1192, Japan<br />

We report some recent results and raise some open questions concerning Gromov-Hausdorff, variational and<br />

spectral convergence <strong>of</strong> finite networks endowed with the resistance metrics and measures, and also Royden’s<br />

compactification <strong>of</strong> infinite networks.<br />

Introduction<br />

Mathematics Subject Classifications (2000): 53C21, 58D17, 58J50.<br />

KEYWORDS: networks, resistance form, resistance metric, convergence <strong>of</strong> Gromov-Hausdorff sense,<br />

-convergence, Royden’s compactification<br />

A Dirichlet problem on harmonic functions in a relatively compact domain <strong>of</strong> a Riemannian manifold is solved by<br />

the direct method <strong>of</strong> calculus <strong>of</strong> variations as follows. We first take a sequence <strong>of</strong> functions with prescribed boundary<br />

values which is minimizing the Dirichlet energy. Then it follows from the Poincaré inequality that the sequence is<br />

bounded in the Sobolev space, and hence by Rellich’s theorem, we have a subsequence which converges to a function<br />

weakly in the Sobolev space and strongly in L 2 space. The lower semi-continuity <strong>of</strong> the Dirichlet energy implies that<br />

the limit function minimizes the energy and satisfies the Laplace equation in a weak sense. Secondly we derive a priori<br />

estimates on the Hölder continuity <strong>of</strong> the solution with respect to the Riemannian distance, using Moser’s iteration<br />

method in which the volume estimates <strong>of</strong> metric balls, the Sobolev inequality, and Poincaré inequality play crucial<br />

roles.<br />

Riemannian manifolds are metric spaces with the Riemannian distances and also Dirichlet spaces with the Dirichlet<br />

energy functionals on the Sobolev spaces. From the former point <strong>of</strong> view, the convergence <strong>of</strong> Riemannian manifolds in<br />

the Gromov-Hausdorff sense has been intensively studied. We have some survey articles, for instances, [8, 28, 29] on<br />

the collapsing phenomena, [22] on the convergence <strong>of</strong> Einstein manifolds, [3, 24] on the Riemannian manifolds with<br />

Ricci curvature bounded below. Also we can refer to the monographs [1, 10, 11] for discussions in details. On the other<br />

hand, from a view point <strong>of</strong> the direct method <strong>of</strong> calculus <strong>of</strong> variations mentioned above, the convergence <strong>of</strong> the<br />

Dirichlet energy functionals on Riemannian manifolds and more generally certain Dirichlet spaces are studied in<br />

[15, 19, 21] (see [16]).<br />

In this paper, we first report some recent results concerning variational convergence <strong>of</strong> finite networks. <strong>Finite</strong><br />

networks are regarded as subspaces <strong>of</strong> metric graphs, by which we mean Riemannian polyhedra <strong>of</strong> dimension one (cf.<br />

[6]). Two metric structures can be introduced on a metric graph, from the geodesic distance and from the effective<br />

resistance which is associated with a canonical energy form <strong>of</strong> the graph, called the resistance form. Accordingly, we<br />

are able to discuss convergence <strong>of</strong> metric graphs with respect to the Gromov-Hausdorff distance. Any compact<br />

geodesic space, for instance, turns out to be the limit <strong>of</strong> compact metric graphs endowed with the Riemannian<br />

distances. On the other hand, a certain class <strong>of</strong> so called fractal sets including the Sierpinski gasket has been studied as<br />

limits <strong>of</strong> finite networks with the resistance metrics by J. Kigami (cf. [20] and the references therein).<br />

An infinite network may be viewed as a limit <strong>of</strong> finite ones and also considered as a combinatorial approximation <strong>of</strong><br />

noncompact Riemannian manifolds. In the final section, we make some observations on Dirichlet finite harmonic<br />

functions and points at infinity <strong>of</strong> an infinite network, raising some open questions.<br />

The details <strong>of</strong> the results mentioned in this report will be taken up in [17] and [18].<br />

1. Metric Graphs, Subspaces and Energy Forms<br />

We are given a finite graph ðV; EÞ with the sets <strong>of</strong> verticies V and edges E and a positive function r on E. Loops and<br />

multiple edges are admitted. Then by regarding edges e as curves <strong>of</strong> length rðeÞ and gluing them at the verticies, we<br />

obtain a compact metric graph, that is, a compact Riemannian polyhedron G <strong>of</strong> dimension one. We suppose that G is<br />

connected unless otherwise are stated. We denote by d R G and sG the Riemannian distance and the canonical Riemannian<br />

measure <strong>of</strong> G, respectively.<br />

An energy form on the space <strong>of</strong> continuous functions on G, CðGÞ, equipped with the uniform norm kk C 0 is defined<br />

as follows. Let D½EGŠ be the space <strong>of</strong> functions u 2 CðGÞ such that on each edge (or one-simplex) e <strong>of</strong> G which is<br />

parametrized by the arc length in the interval ½0; rðeÞŠ, u is absolutely continuous and the derivative u 0 belongs to<br />

Partly supported by the Grant-in-Aid for Scientific Research (B) No. 15340053 <strong>of</strong> the Japan Society for the Promotion <strong>of</strong> Science


58 KASUE<br />

L2ð½0; rðeÞŠÞ. Then a canonical energy form EG on D½EGŠ is given by<br />

Z<br />

Z rðeÞ<br />

Let<br />

EGðu; vÞ ¼<br />

G<br />

u 0 v 0 dsG ¼ X<br />

e2E<br />

0<br />

u 0 ðsÞv 0 ðsÞds; u; v 2 D½EGŠ:<br />

juðxÞ uðyÞj2<br />

RGðx; yÞ ¼sup j u 2 D½EGŠ; EGðu; uÞ 6¼ 0 ; x; y 2 G:<br />

EGðu; uÞ<br />

The number RGðx; yÞ is called the effective resistance between the points x and y <strong>of</strong> G.<br />

The form EG has the following properties.<br />

(i) EG is symmetric and positive semi-definite, and EGðu; uÞ ¼0 if and only if u is constant.<br />

(ii) The form EGðu; vÞþuðoÞvðoÞ on D½EGŠ, where o is a fixed point <strong>of</strong> G, is closed in the sense that it provides a<br />

complete inner product on D½EGŠ.<br />

(iii) For a sequence <strong>of</strong> functions un in D½EGŠ whose energies EGðun; unÞ are uniformly bounded, if un uniformly<br />

converges to a function u in CðGÞ as n !1, then we have<br />

EGðu; vÞ ¼ lim<br />

n!1 EGðun; vÞ<br />

for all v 2 D½EGŠ. As a result, the energy form EG is lower semi-continuous on CðGÞ, that is, for a sequence <strong>of</strong> functions<br />

un 2 CðGÞ which uniformly converges to a function u 2 CðGÞ, we have<br />

EGðu; uÞ lim inf<br />

n!1 EGðun; unÞ ð þ1Þ:<br />

(iv) A function u 2 D½EGŠ satisfies<br />

juðxÞ uðyÞj 2<br />

EGðu; uÞd R Gðx; yÞ; x; y 2 G;<br />

and hence by the definition <strong>of</strong> RG, this estimate is restated as follows:<br />

juðxÞ uðyÞj 2<br />

EGðu; uÞ RGðx; yÞ; u 2 D½EGŠ; x; y 2 G;<br />

RGðx; yÞ d R Gðx; yÞ; x; y 2 G:<br />

In particular, we have<br />

juðxÞj EGðu; uÞ 1=2 RGðx; yÞ 1=2 þjuðyÞj; x; y 2 G: ð1Þ<br />

(v) EG satisfies the Markov property, that is, for u 2 D½EGŠ, u :¼ minfmaxf0; ug; 1g 2D½EGŠ and<br />

EGðu; uÞ EGðu; uÞ:<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

(vi) The embedding <strong>of</strong> ðD½EGŠ; EG þ 2 p<br />

oÞ<br />

into CðXÞ is compact, that is, for a sequence <strong>of</strong> functions in D½EGŠ which<br />

are uniformly bounded and whose energies are also uniformly bounded, there exists a subsequence that converges<br />

uniformly to a continuous function.<br />

(vii) The form EG is strong local, i.e., for u; v 2 D½EGŠ, EGðu; vÞ ¼0 if u is constant on the support <strong>of</strong> v.<br />

Now we are given a closed subset K <strong>of</strong> G. For any u 2 CðKÞ, let Au ¼fv2 D½EGŠ jv ¼ u on Kg and also<br />

F K ¼fu 2 CðKÞ jAu 6¼;g. Then in view <strong>of</strong> the properties (iii) and (iv), we see that for any u 2 F K, there exists<br />

uniquely a minimizer HK;u on Au, which is characterized as a function h 2 Au that satisfies EGðh; vÞ ¼0 for all<br />

v 2 D½EGŠ vanishing on K. Define a form EK on CðKÞ by<br />

D½EKŠ¼F K; EKðu; uÞ ¼EGðHK;u; HK;uÞ; u 2 D½EKŠ: Then E K satisfies the same properties <strong>of</strong> (i) through (vi) as EG; however (vii) does not hold true in general. We note that<br />

given a closed subset L <strong>of</strong> K and a function u 2 D½E K Š, E K ðu; vÞ ¼0 for all v 2 D½E K Š with supp v K n L if and only<br />

if EGðHK;u; vÞ ¼0 for all v 2 D½EGŠ with supp v G n L. In addition, let<br />

juðxÞ uðyÞj2<br />

RKðx; yÞ ¼sup j u 2 D½EKŠ; EKðu; uÞ 6¼ 0 ; x; y 2 K:<br />

EKðu; uÞ<br />

Then RK is nothing but the restriction <strong>of</strong> the effective resistance RG <strong>of</strong> G to the set K, i.e.,<br />

R K ðx; yÞ ¼RGðx; yÞ; x; y 2 K:<br />

Now the Markov property (v) implies that the maximum principle holds true for HK;u, that is,<br />

min<br />

K u HK;u max<br />

K u:<br />

Here we recall another important consequence from the Markov property that for u; v 2 D½EKŠ, uv 2 D½EKŠ and we<br />

have


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 59<br />

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

EKðuv; uvÞ<br />

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

kukC0 EKðv; vÞ þkvkC0 EKðu; uÞ<br />

(cf. [9], Theorem 1.4.2). In particular, D½EKŠ is a subalgebra <strong>of</strong> CðKÞ containing the unit element 1 and separating<br />

points <strong>of</strong> K. As a result, D½EKŠ is dense in CðKÞ, since K is compact.<br />

Now we consider the case where K consists <strong>of</strong> N points x1; ...; xN <strong>of</strong> G. We denote by ui ði ¼ 1; ...; NÞ the function<br />

on K defined by uiðxjÞ ¼ ij. Let cij ¼ EKðui; ujÞ ði; j ¼ 1; ...; NÞ. Since EKðui; 1Þ ¼0, we get PN j¼1 cij ¼ 0 ði ¼<br />

1; ...; NÞ, and hence we can deduce that<br />

EKðu; vÞ ¼ 1 X<br />

2<br />

N<br />

cijðuðxiÞ uðxjÞÞðvðxiÞ vðxjÞÞ; u; v 2 D½EKŠ: i;j¼1<br />

Note that cij 0 if i 6¼ j, which follows from the Markov property <strong>of</strong> E K (cf. [20], p. 43).<br />

2. Green Functions, Effective Resistance, and Laplace Operators<br />

2.1 Let G be a compact connected metric graph corresponding to a finite network ðV; E; rÞ. For points a; x 2 G, in view<br />

<strong>of</strong> (1), we have uniquely a function gaðx; Þ 2 D½EGŠ satisfying EGðgaðx; Þ; vÞ ¼vðxÞ vðaÞ for all v 2 D½EGŠ, and<br />

gaðx; aÞ ¼0. Observe that gaðx; yÞ ¼gaðy; xÞ, since gaðx; yÞ ¼EGðgaðx; Þ; gaðy; ÞÞ ¼ EGðgaðy; Þ; gaðx; ÞÞ ¼ gaðy; xÞ.<br />

The functions ga are called the Green functions <strong>of</strong> the form EG. Given two points x; y 2 G, let HxyðzÞ ¼gyðz; xÞ=gyðx; xÞ<br />

ðz 2 GÞ. Then the effective resistance RG is expressed as<br />

RGðx; yÞ ¼EGðHxy; HxyÞ 1 ¼ gyðx; xÞð¼ gxðy; yÞÞ; x; y 2 G:<br />

In fact, we can show the following result, which will be verified after Theorem 2.2.<br />

Theorem 2.1. The following mutually equivalent identities hold:<br />

gzðx; yÞ ¼ 1<br />

2 fRGðx; zÞþRGðz; yÞ RGðx; yÞg;<br />

RGðx; yÞ ¼gxðy; yÞ ¼gyðx; xÞ ¼gxðy; zÞþgyðx; zÞ<br />

for all x; y; z 2 G.<br />

Since gzðx; yÞ 0 by the maximum principle, the first identity implies that RG : G G ! R satisfies the triangle<br />

inequality and hence we have the following<br />

Corollary 2.2. The effective resistance RG is a metric on G, called the resistance metric <strong>of</strong> G, which induces the same<br />

topology as the Riemannian distance.<br />

Remarks (i) Let G and ðV; E; rÞ be as above. Then the resistance metric <strong>of</strong> G coincides with its Riemannian distance if<br />

and only if G is simply connected. Let ðV0 ; E0Þ be a subnetwork <strong>of</strong> ðV; EÞ, i.e., a pair <strong>of</strong> sets V0 V and E0 E such that<br />

x; y 2 V0 if fx; yg 2E0 . Let G0 be the metric graph associated with the subnetwork ðV0 ; E0 ; rjE0Þ and assume that G0 is<br />

connected. Then we have<br />

RGðx; yÞ RG 0ðx; yÞ; x; y 2 G0 :<br />

This is known as Rayleigh’s monotonicity principle.<br />

(ii) The identities in Theorem 2.1 are presented for finite networks in [27].<br />

2.2 Let be a Radon measure on G with support K and consider the Laplace operator L associated with the closed<br />

form EK on L2ðK; Þ. The domain D½L Š <strong>of</strong> L consists <strong>of</strong> functions u 2 D½EKŠ such that the functional on D½EKŠ defined by v ! EKðu; vÞ is continuous with respect to the norm <strong>of</strong> L2ðK; Þ, and for any u 2 D½L Š, L u is the unique<br />

function (in the closure <strong>of</strong> D½EKŠ in L2ðK; Þ) satisfying<br />

Z<br />

EKðv; uÞ ¼ v L ud ; v 2 D½EKŠ: Let<br />

g ðx; yÞ ¼ 1<br />

Z<br />

gzðx; yÞ d ðzÞ;<br />

ðKÞ K<br />

x; y 2 K:<br />

Then by the definition <strong>of</strong> the Green functions, we have<br />

uðxÞ ¼ 1<br />

Z<br />

ud<br />

ðKÞ K<br />

Z<br />

þ g ðx; yÞL uðyÞ d ðyÞ;<br />

K<br />

u 2 D½L Š; x 2 K;<br />

and hence<br />

K


60 KASUE<br />

uðxÞ ¼ 1<br />

Z<br />

ud<br />

ðKÞ K<br />

þ EKðg ðx; Þ; uÞ; u 2 D½EKŠ; x 2 K:<br />

Applying this to the functions gzðx; Þ, we can verify the following identity:<br />

gzðx; yÞ ¼g ðz; zÞ g ðz; xÞ g ðz; yÞþg ðx; yÞ; x; y; z 2 K: ð2Þ<br />

Since RGðx; yÞ ¼gxðy; yÞ, direct computations show the following<br />

Theorem 2.3. The effective resistance on K is given by<br />

RGðx; yÞ ¼g ðx; xÞ 2g ðx; yÞþg ðy; yÞ; x; y 2 K:<br />

Now by taking a Radon measure such that supp ¼ G and using (2), we can derive the first identity <strong>of</strong><br />

Theorem 2.1, from which the remaining one follows.<br />

Let<br />

g ðx; yÞ :¼ g ðx; yÞ C ; x; y 2 K;<br />

and<br />

Z<br />

G uðxÞ ¼<br />

K<br />

g ðx; yÞuðyÞ d ðyÞ; u 2 L 2 ðK; Þ;<br />

where C is a positive constant given by<br />

C ¼ 1<br />

Z<br />

1<br />

g ðx; yÞ d ðyÞ ¼<br />

ðKÞ K<br />

2 ðKÞ2 Z Z<br />

RGðy; zÞ d ðyÞd ðzÞ:<br />

K K<br />

Then G turns out to be the Green operator <strong>of</strong> L , that is, it satisfies<br />

I ¼ H þ L G on L2ðK; Þ;<br />

I ¼ H þ G L on D½L Š;<br />

H G ¼ G H ¼ 0;<br />

where we put<br />

H u ¼ 1<br />

Z<br />

ud :<br />

ðKÞ K<br />

Let N ¼ #K þ1and i ði ¼ 0; 1; ...; N 1Þ the i-th eigenvalue <strong>of</strong> L , where 0 ¼ 0. We take a complete<br />

orthonormal system <strong>of</strong> eigenfunctions i with eigenvalue i in L2ðK; Þ. Then the Green kernel g is given by<br />

g ðx; yÞ ¼<br />

XN 1<br />

i¼1<br />

1<br />

i<br />

iðxÞ iðyÞ:<br />

Since RGðx; yÞ ¼g ðx; xÞ 2g ðx; yÞþg ðy; yÞ, as a result <strong>of</strong> Theorem 2.3, we obtain<br />

Corollary 2.4. It holds:<br />

In particular, one has<br />

RGðx; yÞ ¼<br />

Z<br />

K<br />

Z<br />

K<br />

XN 1<br />

i¼1<br />

1<br />

ð iðxÞ iðyÞÞ 2 ; x; y 2 K:<br />

i<br />

XN 1<br />

1<br />

RGðx; yÞd ðxÞd ðyÞ ¼2 ðKÞ :<br />

i¼1 i<br />

From the last identity <strong>of</strong> the corollary, it is easy to see that<br />

i<br />

where DðK; RGÞ ¼supfRGðx; yÞ jx; y 2 Kg.<br />

2i<br />

DðK; RGÞ<br />

;<br />

ðKÞ<br />

i ¼ 1; 2; ...; ð3Þ<br />

2.3 Now we consider the case where a closed subset K <strong>of</strong> a compact metric graph G consists <strong>of</strong> a finite number <strong>of</strong><br />

points fx1; ...; xNg ðN < þ1Þ. We define N functions ui ði ¼ 1; ...; NÞ on K by uiðxjÞ ¼ ij. Then ui 2 D½EKŠ. Let<br />

cij ¼ EKðui; ujÞ. For every fixed k, 1 k N, the definition <strong>of</strong> the Green functions reads


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 61<br />

X N<br />

‘¼1<br />

gxk ðxi; x‘Þc‘j ¼ ij; i 6¼ k; j 6¼ k; ð4Þ<br />

that is, defining ðN 1Þ ðN 1Þ matrices Ck and Gk respectively by Ck ¼ðcijÞ and Gk ¼ðgxkðxi; i 6¼ k; j 6¼ k, we have<br />

xjÞÞ, where<br />

GkCk ¼ IN 1; k ¼ 1; ...; N;<br />

where IN 1 stands for the unit matrix. By (4), we get<br />

XN XN k¼1 i;j¼1<br />

gxk ðxi; xjÞcij ¼ NðN 1Þ: ð5Þ<br />

Since gxkðxi; xjÞ ¼ 1<br />

2 ðRGðxi; xkÞþRGðxk; xjÞ RGðxi; xjÞÞ and PN j¼1 cij ¼ 0, we thus obtain the following:<br />

1 X<br />

2<br />

N<br />

RGðxi; xjÞEK ðui; ujÞ ¼N 1 ð6Þ<br />

When K is the set <strong>of</strong> all vertecies V, cij ¼ 1=rði; jÞ ði 6¼ jÞ, and hence the identity reads<br />

1<br />

2<br />

i;j¼1<br />

X N<br />

i;j¼1;i 6¼ j<br />

RGðxi; xjÞ<br />

rði; jÞ<br />

¼ N 1:<br />

This is a classical result due to R. M. Foster (cf. [7, 27]).<br />

P c N<br />

Now taking a canonical measure on K defined by ¼ i¼1<br />

the following<br />

xi and using the identities <strong>of</strong> Corollary 2.4, we have<br />

Corollary 2.5. Under the conditions as above, one has<br />

and<br />

max<br />

1 k n 1<br />

2k<br />

ðn 1Þ c<br />

kð Þ<br />

2<br />

c<br />

n 1ð Þ<br />

max<br />

xi;x j2K RGðxi; xjÞ<br />

RGðxi; xjÞ; xi; xj 2 K; xi 6¼ xj:<br />

2<br />

c<br />

1ð Þ<br />

Remark The second inequality in this corollary is proved in [2].<br />

2.4 Let V be a finite set, V ¼fx1; ...; xNg, and consider a nonnegative quadratic form E on the space ‘ðVÞ <strong>of</strong> functions<br />

on V such that Eðu; uÞ ¼0 if and only if u is constant. Let REðx; yÞ ¼maxfjuðxÞ uðyÞj2 =Eðu; uÞ ju 2 ‘ðVÞ; Eðu; uÞ ><br />

0g ðx; y 2 VÞ. Then the arguments and the results as in the previous subsections 2.1 through 2.3 are valid, except the<br />

nonnegativity <strong>of</strong> the Green functions gzðx; yÞ <strong>of</strong> E; this relies on the maximum principle under the Markov property. In<br />

fact, if we assume that the form satisfies the Markov property, then we see that cij ¼ Eðui; ujÞ 0 for all i; j ¼ 1; ...; N<br />

with i 6¼ j, where fu1; ...; uNg is a canonical basis <strong>of</strong> ‘ðVÞ described in subsection 2.3. In this case, by letting E ¼<br />

ffx; yg V j cij < 0g and rði; jÞ ¼ 1=cij, we obtain a network ðV; E; rÞ. In particular, Theorem 2.1 holds true without<br />

the Markov property, and hence we have the following<br />

Theorem 2.6. Let V be a finite set. Let E ð ¼ 1; 2Þ be nonnegative quadratic forms on ‘ðVÞ such that E ðu; uÞ ¼0 if<br />

and only if u is constant. Then E1 ¼ E2 if and only if RE1 ¼ RE2 .<br />

Given a subset K <strong>of</strong> V, we denote as before by EK a form on ‘ðKÞ defined by EKðu; uÞ ¼inffEðv; vÞ jv 2 ‘ðVÞ;<br />

vjK ¼ ug. Then we have RE ðx; yÞ ¼REðx; yÞ for all x; y 2 K. As a result <strong>of</strong> Theorem 2.6, we have the following<br />

K<br />

Corollary 2.7. Let K be a subset <strong>of</strong> a finite set V. Let E and F be respectively quadratic forms on ‘ðVÞ and ‘ðKÞ as<br />

above. Then F ¼ EK if and only if RF ðx; yÞ ¼REðx; yÞ for all x; y 2 K.<br />

Theorem 2.6 and Corollary 2.7 are respectively proved in [20], Theorem 2.1.12 and Corollary 2.1.13, under the<br />

condition <strong>of</strong> the Markov property, and they affermatively answer an question raised there.<br />

2.5 In this part, we recall some definitions and results on resistance forms and resistance metrics from [20], Chap. 2.<br />

Let X be a set and ðE; D½EŠÞ a pair <strong>of</strong> a linear subspace D½EŠ <strong>of</strong> the space ‘ðXÞ <strong>of</strong> all functions on X and a nonnegative<br />

quadratic form E on D½EŠ. We call such a pair ðE; D½EŠÞ a resistance form on X if it satisfies the following conditions<br />

(RF-i) through (RF-v):<br />

(RF-i) Eðu; uÞ ¼0 if and only if u is constant on X.<br />

(RF-ii) Let be an equivalence relation on D½EŠ defined by u v if and only if u v is constant on X. Then<br />

ðD½EŠ= ; EÞ is a Hilbert space.


62 KASUE<br />

(RF-iii) For any finite subset V <strong>of</strong> X and for any u 2 ‘ðVÞ, there exists v 2 D½EŠ such that vjV ¼ u.<br />

(RF-iv) For any x; y 2 X, supfjuðxÞ uðyÞj2 =Eðu; uÞ ju 2 D½EŠ; Eðu; uÞ > 0g is finite. The supremum is denoted by<br />

REðx; yÞ, which will be called the effective resistance between the points x and y.<br />

(RF-v) If u 2 D½EŠ, then u 2 D½EŠ and Eðu; uÞ Eðu; uÞ, where u is defined by uðxÞ ¼1 if uðxÞ 1, uðxÞ ¼uðxÞ if<br />

0 < uðxÞ < 1, and uðxÞ ¼0 if uðxÞ 0.<br />

We use RF ðXÞ to denote the collection <strong>of</strong> resistance forms on X. Given ðE; D½EŠÞ 2 RF ðXÞ and a finite subset V <strong>of</strong><br />

X, let<br />

EVðu; uÞ ¼inffEðv; vÞ jv 2 D½EŠ; vjV ¼ ug; u 2 ‘ðVÞ:<br />

Then ðEV ;‘ðVÞÞ 2 RF ðVÞ and RE ¼ RE on V. This implies that if ðE; D½EŠÞ 2 RF ðXÞ, then RE induces a metric on X,<br />

V pffiffiffiffiffi called the resistance metric associated with the resistance form E on X. Obviously RE is also a distance on X; indeed,<br />

so is it without the Markov property (RF-v), if we assume instead <strong>of</strong> property (RF-iii) that REðx; yÞ > 0 for all x; y 2 X<br />

with x 6¼ y.<br />

A function R : X X !½0; þ1Þ is by definition a resistance metric on X if, for any finite subset V X, there exists<br />

a resistance form EV on ‘ðVÞ such that Rðx; yÞ ¼REV ðx; yÞ for all x; y 2 V, where REV is the effective resistance with<br />

respect to the form EV. The collection <strong>of</strong> resistance metrics on X is denoted by RMðXÞ. Each ðE; D½EŠÞ 2 RF ðXÞ is<br />

associated with RE 2 RMðXÞ.<br />

pffiffiffiffiffi This correspondence is in fact bijective.<br />

We note that, when ðX; REÞ<br />

is separable, E can be thought to be a limit <strong>of</strong> finite dimensional forms; in fact, given an<br />

increasing sequence fVmg <strong>of</strong> finite subsets <strong>of</strong> X such that [mVm is dense in X, a function u on X belongs to D½EŠ if and<br />

only if limm!1 EVmðujVm ; ujVmÞ < þ1 and in this case, Eðu; uÞ ¼limm!1 EVmðujVm ; ujVmÞ, where E is the induced<br />

Vm<br />

form on ‘ðVmÞ as above.<br />

Given ðE; D½EŠÞ 2 RF ðXÞ, ifX is the completion <strong>of</strong> X with respect to the resistance metric, then any u 2 D½EŠ is<br />

naturally extended to a continuous function on X. Using this extention, E is regarded as the collection <strong>of</strong> functions on X.<br />

Then ðE; D½EŠÞ is a resistance form on X and the resistance metric associated with ðE; D½EŠÞ on X is the natural<br />

extension <strong>of</strong> the resistance metric associated with ðE; D½EŠÞ on X.<br />

2.6 Before proceeding to the next section, we consider a wider class <strong>of</strong> quadratic forms for our purpose. Let ðE; D½EŠÞ<br />

be a nonnegative quadratic form on a set X and<br />

juðxÞ uðyÞj2<br />

REðx; yÞ ¼sup j u 2 D½EŠ; Eðu; uÞ > 0 ; x; y 2 X:<br />

Eðu; uÞ<br />

pffiffiffiffiffi We suppose that 0 < REðx; yÞ < þ1 for all x; y 2 X with x 6¼ y. Then RE<br />

pffiffiffiffiffi induces a distance on X, and the domain<br />

D½EŠ is a subspace <strong>of</strong> CðX; REÞ.<br />

Moreover we suppose that for a point o 2 X, a quadratic form defined by Eðu; vÞþ<br />

uðoÞvðoÞ provides a complete inner product on D½EŠ, that is, ðD½EŠ; E þ 2 oÞ is a Hilbert space (this holds for any o 2 X<br />

because <strong>of</strong> the assumption above).<br />

pffiffiffiffiffi Now we suppose that the metric space ðX; REÞ<br />

is separable. Then we are able to verify the following assertions:<br />

pffiffiffiffiffi (i) When the form E is regarded as a functional on the space <strong>of</strong> continuous functions CðXÞ on X ¼ðX; REÞ<br />

by<br />

letting EðuÞ ¼Eðu; uÞ if u 2 D½EŠ and EðuÞ ¼þ1otherwise, pffiffiffiffiffi the functional E : CðXÞ !½0; þ1Š is lower semicontinuous<br />

with respect to the uniform topology <strong>of</strong> CðX; REÞ.<br />

(ii) Given a closed subset K <strong>of</strong> X, let<br />

D½EKŠ¼fu 2 CðKÞ ju ¼ vjK 9v 2 D½EŠg;<br />

EKðu; uÞ ¼inffEðv; vÞ jv 2 D½EŠ; vjK ¼ ug:<br />

Then ðEK ; D½EKŠÞ is a quadratic form on CðKÞ with the same properties as E on X. In fact, there exists a unique<br />

minimizer HK;u and the correspondence u 2 D½EKŠ!HK;u 2 D½EŠ is linear and injective. The minimizer is<br />

characterized as a function h 2 D½EŠ which coincides with u on K and satisfies Eðh; vÞ ¼0 for all v 2 D½EŠ vanishing on<br />

K. In addition, we see that supfjuðxÞ uðyÞj2 =EKðu; uÞ ju 2 D½EKŠ; EKðu; uÞ 6¼ 0g ¼Rðx; yÞ for all x; y 2 K.<br />

(iii) If 1 2 D½EŠ and Eð1; 1Þ ¼0, then Theorem 2.1 is valid for this situation. If, in addition, E satisfies the Markov<br />

property, then so does EK , and hence D½EKŠ\L1 is a subalgebra <strong>of</strong> CðKÞ containing the unit elememt 1 and separating<br />

points <strong>of</strong> K;thus in this case, ðE; D½EŠÞ is a resistance form on X.<br />

(iv) Let be a -finite Borel measure on X and K the support <strong>of</strong> . Let H be the L2-closure <strong>of</strong> D½EKŠ\L2ðK; Þ and<br />

L the self-adjoint operator associated with the closed form EK in H with domain D½EKŠ\L2ðK; Þ. Given a positive<br />

number , we use R ; to denote the resolvent operator ðL þ IÞ 1 <strong>of</strong> L . Suppose that<br />

Z<br />

ðKÞ < þ1; REðo; xÞ d ðxÞ < þ1<br />

K<br />

for some o 2 X (and hence any o). Then D½E K Š L 2 ðK; Þ and the embedding is compact with respect to the norm<br />

ðE K ðu; uÞþ R u 2 d Þ 1=2 <strong>of</strong> D½E K Š. If, in addition, 1 2 D½EŠ and Eð1; 1Þ ¼0, then Theorem 2.3, Corollary 2.4 and<br />

Corollary 2.5 hold true.


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 63<br />

P c N<br />

Example Let V be a set <strong>of</strong> N points, V ¼fx1; ...; xNg, and ¼ i¼1 xi . Let fe0; e1; ...; eN 1g be an orthonormal<br />

basis <strong>of</strong> L2 c ðV; Þð¼ ‘ðVÞÞ such that e0ðxÞ ¼1= ffiffiffiffi p<br />

N for all x 2 V. Given k, 0 < k < N, and positive numbers i<br />

ði ¼ 1; ...; kÞ such that 1 k, we define a quadratic form ðE; D½EŠÞ on V by D½EŠ ¼fu ¼ Pk i¼1 uiei j u 2 ‘ðVÞg<br />

and Eðu; vÞ ¼ Pk i¼1 iuivi, u; v 2 D½EŠ. Then the effective resistance RE is given by<br />

REðx; yÞ ¼ Xk 1<br />

ðeiðxÞ eiðyÞÞ 2 ; x; y 2 V:<br />

i¼1<br />

i<br />

We observe that if, for any pair <strong>of</strong> points<br />

p<br />

x;<br />

ffiffiffiffiffi<br />

y 2 V, there exists an ei, 1 i k, such that eiðxÞ 6¼ eiðyÞ, that is,<br />

fe1; ...; ekg separates the points <strong>of</strong> V, then RE is a distance on V. We also note that given a sequence <strong>of</strong> positive<br />

numbers "n tending to 0 as n !1, a sequence <strong>of</strong> the forms En defined by Enðu; uÞ ¼ Pk i¼1 iu2 P 1 N 1<br />

i þ " n j¼kþ1 u2j converges to ðE; D½EŠÞ as n !1.<br />

Consider the case where N ¼ 3, k ¼ 1, 1 ¼ 1 and e1 ¼ð2ð 2 þ þ 1ÞÞ 1=2ð1; ; 1 Þ. Then the Green functions<br />

<strong>of</strong> E are all nonnegative if and only if 2 < < 1=2; in this case, RE satisfies the triangle inequality.<br />

3. Gromov-Hausdorff and <strong>Variational</strong> <strong>Convergence</strong><br />

3.1 Let us recall first the definition <strong>of</strong> the Gromov-Hausdorff convergence <strong>of</strong> metric spaces. Let X and Y be metric<br />

spaces and ">0. A (not necessarily continuous) map f : X ! Y is called an "-Hausdorff approximation if<br />

sup jdYð f ðx1Þ; f ðx2ÞÞ dXðx1; x2Þj 0, there exists a positive integer n0 such that for any n > n0, there is a map f from the ball BrðpnÞ around<br />

pn with radius r in Xn to X satisfying the following properties:<br />

(i) f ðpnÞ ¼p;<br />

(ii) supfjdXð f ðx1Þ; f ðx2ÞÞ dXnðx1; x2Þj j x1; x2 2 BrðpnÞg


64 KASUE<br />

if limn!1 supXn ju fn unj ¼0. Let F : CðYÞ !½0; þ1Š and F n : CðXnÞ !½0; þ1Š be lower semi-continuous<br />

functionals on CðYÞ and CðXnÞ respectively. We say that F n -converges to F if the following conditions are satisfied:<br />

(i) if a sequence <strong>of</strong> functions un 2 CðXnÞ uniformly converges to a function u 2 CðYÞ, then we have<br />

F ðuÞ lim inf<br />

n!1 F nðunÞ ð þ1Þ;<br />

(ii) for any u 2 CðYÞ, there exists a sequence <strong>of</strong> functions un 2 CðXnÞ such that un uniformly converges to u and<br />

lim sup F nðunÞ<br />

n!1<br />

F ðuÞ ð þ1Þ:<br />

The following is a basic fact on this variational convergence.<br />

Theorem 3.3. Let Y and fXng be respectively a compact, separable, Hausdorff space and a sequence <strong>of</strong> such spaces.<br />

Given a sequence <strong>of</strong> maps fn : Xn ! Y and a sequence <strong>of</strong> lower semi-continuous functionals F n : CðXnÞ !½0; þ1Š,<br />

there exsists a subsequence, F m, and a lower semi-continuous functional F : CðYÞ !½0; þ1Š such that F m<br />

converges to F as m !1.<br />

-<br />

Pro<strong>of</strong>. Using the idea <strong>of</strong> De Giorgi’s -convergence (cf. e.g. [5],), we introduce a functional on CðYÞ as follows: Let<br />

B ¼fOig be a countable basis <strong>of</strong> CðYÞ such that Oi is totally bounded. Given Oi and a positive integer k, let<br />

Oi;k;n ¼fv 2 CðXnÞ jsup ju<br />

Xn<br />

fn vj < 1=k for some u 2 Oig;<br />

and<br />

Ei;k;n ¼ inffF nðvÞ jv 2 Oi;k;ng ð þ1Þ:<br />

Then passing to a subsequence, fXmg, we may assume that for any Oi and every k, Ei;k;m tends to an extended number<br />

Ei;k 2½0; þ1Š as m !1, and thus we are able to obtain a lower semi-continuous functional F : CðYÞ !½0; þ1Š<br />

defined by<br />

F ðuÞ ¼supfEi;k j u 2 Oi; k > 0g; u 2 CðYÞ;<br />

to which F m -converges as m !1.<br />

Let F n : CðXnÞ !½0; þ1Š and F : CðYÞ !½0; þ1Š be as in Theorem 3.3. If F n is induced from a quadratic form<br />

En on a subspace D½EŠ <strong>of</strong> CðXnÞ, that is, F ðuÞ ¼Eðu; uÞ for u 2 D½EŠ and F ðuÞ ¼þ1for u 2 CðXnÞnD½EŠ, then the -<br />

limit F : CðYÞ !½0; 1Þ is also induced from a quadratic form E on a subspace D½EŠ <strong>of</strong> CðYÞ. For the functional<br />

induced from a quadratic frm E : D½EŠ D½EŠ !R, we do not distinguish between the functional and the form E. If all<br />

F n satisfy the property that F nðuÞ F ðuÞ, u 2 CðXnÞ, then so does the -limit F , where for a continuous function<br />

u 2 CðXnÞ, we set u ¼ maxf0; minfu; 1gg.<br />

3.3 Now we state some results. Let fðXn; RnÞg be a sequence <strong>of</strong> compact metric spaces <strong>of</strong> resistance forms En. We<br />

assume that the following conditions are satisfied:<br />

(i) There exist a compact, separable Hausdorff space Y, a sequence <strong>of</strong> maps fn : Xn ! Y and also a sequence <strong>of</strong> maps<br />

hn : Y ! Xn such that fn hn uniformly converges to the identity map <strong>of</strong> Y as n !1.<br />

(ii) The functional En -converges (via fn) to a functional E : CðYÞ !½0; þ1Š as n !1.<br />

(iii) For any sequence <strong>of</strong> functions un 2 D½EnŠ such that supn maxXn junj < þ1 and supn Enðun; unÞ < þ1, there<br />

exists a subsequence fumg which uniformly converges to a function u 2 CðYÞ as m !1.<br />

Under these conditions, our main result is stated in<br />

Theorem 3.4.<br />

(iv) Let<br />

The following assertions hold:<br />

juðxÞ uðyÞj2<br />

REðx; yÞ ¼sup j u 2 D½EŠ; Eðu; uÞ 6¼ 0 ;<br />

Eðu; uÞ<br />

x; y 2 Y:<br />

Then RE : Y Y !½0; þ1Š induces a continuous pseudo-distance on Y (admitting þ1 in its values), and one has<br />

0 REðx; yÞ ¼ lim<br />

n!1 RnðhnðxÞ; hnðyÞÞ þ1; x; y 2 Y:<br />

(v) On Y, an equivalence relation b is introduced as follows: x b y if and only if REðx; yÞ < þ1. Then Y is<br />

decomposed into a finite number <strong>of</strong> the equivalence classes Y ð ¼ 1; ...; pÞ; each class Y is open and closed in Y.<br />

Moreover for each and large n, the inverse image <strong>of</strong> Y by fn, Xn; ¼ f 1<br />

n ðY Þ, is open and closed in Xn and one has<br />

lim<br />

n!1 supfjRnðx; yÞ REð fnðxÞ; fnðyÞÞj j x; y 2 Xn; g¼0;<br />

lim REð fnðXn; Þ; yÞ ¼0:<br />

n!1 sup<br />

y2Y


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 65<br />

(vi) Let be the characteristic function <strong>of</strong> the subspace Y ð1 pÞ. Then 2 D½EŠ and Eð ; uÞ ¼0 for all<br />

u 2 D½EŠ, and if Eðu; uÞ ¼0, then u is a linear combination <strong>of</strong> the characteristic functions ð1 pÞ. Moreover let<br />

D½E Š¼fu 2 D½EŠ jsupp u Y g and E ðu; vÞ ¼Eðu; vÞ ðu; v 2 D½E ŠÞ. Then one has D½EŠ ¼ Pp ¼1 D½E Š and<br />

Eðu; vÞ ¼ Pp ¼1 E ð u; vÞ<br />

(vii) Another equivalence relation 0 is defined on Y by x 0 y if and only if REðx; yÞ ¼0. Let Y ¼ Y= 0 and<br />

Y ¼ Y = 0 ð ¼ 1; ...; pÞ be respectively the quotient spaces <strong>of</strong> Y and Y . Then for each , RE provides Y a distance<br />

R which induces the same topology as the original one, and D½E Š is included in the pull-back <strong>of</strong> CðY Þ by the<br />

canonical projection <strong>of</strong> Y onto Y . Thus the form E can be assumed to be defined on CðY Þ; ðE ; D½E ŠÞ becomes a<br />

resistance form on Y and R is the associated resistance metric, that is,<br />

juðx Þ uðy Þj2<br />

R ðx ; y Þ¼sup j u 2 D½E Š; E ðu; uÞ 6¼ 0 ;<br />

E ðu; uÞ<br />

x ; y 2 Y :<br />

Moreover a sequence <strong>of</strong> the compact metric spaces ðXn; ; RnÞ converges to ðY ; R Þ as n !1 in the Gromov-<br />

Hausdorff sense via the approximating maps fn : Xn; ! Y , and the form EXn; on CðXn;<br />

n !1.<br />

Þ -converges to E as<br />

We remark that the compactness condition (iii) as above is indispensable in Theorem 3.4, and some results follows<br />

from the theorem.<br />

Corollary 3.5. Let fKng be a sequence <strong>of</strong> closed subspaces <strong>of</strong> compact metric graphs Gn and suppose that the metric<br />

space Kn with the induced Riemannian distance dR Gn converges to a compact metric space ðX; dXÞ in the Gromov-<br />

Hausdorff sense via approximating maps fn : Kn ! X. Then the resistance metric RGn on Kn converges to a continuous<br />

pseodo-distance R on X with respect to the Gromov-Hausdorff distance via the same approximating maps if and only if<br />

the resistance form EKn has<br />

-converges to a lower semi-continuous functional E on CðXÞ as n !1. In these cases, one<br />

0 Rðx; yÞ dXðx; yÞ; x; y 2 X;<br />

and R is given by<br />

Rðx; yÞ ¼sup<br />

juðxÞ uðyÞj2<br />

Eðu; uÞ<br />

j u 2 D½EŠ; Eðu; uÞ 6¼ 0 ; x; y 2 X:<br />

Corollary 3.6. Let fðXn; RnÞg be a sequence <strong>of</strong> compact metric spaces <strong>of</strong> resistance forms En which converges to a<br />

compact metric space ðY; RYÞ in the Gromov-Hausdorff sense via approximating maps fn : Xn ! Y. Then the resistance<br />

form E -converges to a resistance form E on CðXÞ which is associated with the limit distance RY.<br />

Kn<br />

In these corollaries, the definition <strong>of</strong> a resistance metric allows us to apply Ascoli-Arzelà’s theorem to the sequences<br />

and see that the compactness condition (iii) holds true.<br />

In the case where a sequence <strong>of</strong> networks ðV; En; rnÞ with the same set <strong>of</strong> verticies V is considered, letting fn as in<br />

Theorem 3.4 be the identity map <strong>of</strong> V, we see tha the compactness condition (iii) is always satisfied. Therefore we can<br />

apply our theorem to this case (cf. [4]).<br />

4. Spectral Embeddings and Spectral <strong>Convergence</strong><br />

4.1 We consider a nonnegative quadratic form ðE; D½EŠÞ on a set X satisfying the properties<br />

pffiffiffiffiffi (RF-i), (RF-ii), (RF-iv) and<br />

that REðx; yÞ > 0 for all x; y 2 X with x 6¼ y. We suppose that the metric space ðX; RE<br />

pffiffiffiffiffi Þ is separable. Let be a Borel<br />

measure on X ¼ðX; REÞ<br />

with support K such that<br />

Z<br />

ðXÞð¼ ðKÞÞ < þ1; REðx0; xÞ d ðxÞ < þ1<br />

for some x0 2 X (and hence for any x0). Denote the cardinality <strong>of</strong> the set K by N ¼ #Kð þ1Þ. Letf i j 0 ¼ 0 <<br />

1 2 ; 0 i N 1g the set <strong>of</strong> eigenvalues <strong>of</strong> the self-adjoint operator L associated with the induced form<br />

E K on L 2 ðK; Þ and ¼f i j 0 i N 1g an orthonormal complete system <strong>of</strong> eigenfunctions. Recall that the<br />

effective resistance RE is expressed on K as follows:<br />

REðx; yÞ ¼<br />

XN 1<br />

i¼1<br />

iðxÞ<br />

pffiffiffiffi<br />

i<br />

X<br />

iðyÞ<br />

pffiffiffiffi<br />

i<br />

In what follows, we understand i= ffiffiffiffi p<br />

i ¼ 0 for i N if N is finite.<br />

Let ‘ 2 ¼fðaiÞ j P1 i¼1 a2i < þ1g, and define a map I <strong>of</strong> K into ‘2 by<br />

2<br />

; x; y 2 K:


66 KASUE<br />

I ðxÞ ¼<br />

iðxÞ<br />

p ffiffiffiffi ; x 2 K:<br />

i<br />

ffiffiffiffiffi<br />

RE<br />

Then I<br />

p<br />

gives rise to an isometric embedding <strong>of</strong> ðK;<br />

2 Þ into ‘ . In fact, it satisfies the following properties:<br />

RGðx; yÞ ¼kI ðxÞ I ðyÞk 2 ‘ 2;<br />

g ðx; yÞ ¼ðI ðxÞ; I ðyÞÞ ‘ 2; x; y 2 K:<br />

Now we introduce another embedding <strong>of</strong> K into ‘ 2 as follows:<br />

J ðxÞ ¼ðe i=2<br />

iðxÞÞ; x 2 K:<br />

This embedding has the following properties:<br />

kJ ðxÞ J ðyÞk 2 XN 1<br />

‘ 2 ¼<br />

i¼1<br />

e i ð iðxÞ iðyÞÞ 2<br />

¼ p ð1; x; xÞ 2p ð1; x; yÞþp ð1; y; yÞ; x; y 2 K;<br />

where p ðt; x; yÞ denotes the kernel function <strong>of</strong> the semigroup expð tL Þ generated by L . Let<br />

S ðx; yÞ ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

p<br />

p ð1; x; xÞ 2p ð1; x; yÞþp ð1; y; yÞ;<br />

x; y 2 K:<br />

Then S : K K !½0; þ1Þ is a distance on K and J is an isometric embedding <strong>of</strong> the metric space ðK; S Þ into ‘ 2 .It<br />

is easy to see that<br />

0 S ðx; yÞ REðx; yÞ 1=2 ; x; y 2 K:<br />

Given a function u 2 D½EKŠ, we simply denote by eu the minimizer HK;u in the set fv 2 D½EXŠ jvjK ¼ ug (see Section<br />

1). Correspondingly we denote by eI , eJ , ep and eS respectively the continuous extentions <strong>of</strong> the embeddings I , J ,<br />

the kernel function p and the distance S . We note that the extended maps eI and eJ may not be injective on X and<br />

hence eS may be degenerate somewhere on X.<br />

4.2 Now consider the same situation as in Theorem 3.4 and let fðXn; RnÞg, ðY; RY; EÞ, fn : Xn ! Y and hn : Y ! Xn be as<br />

before. Let us take a sequence <strong>of</strong> Radon measures n on Xn with supp n ¼ Xn such that the push-forward measure<br />

fn n weakly converges to a Radon measure on Y as n !1. Let L be the Laplacian in L2ðK; Þ, where<br />

K ¼ supp , associated with the induced form EK on CðKÞ and ep the continuous extension <strong>of</strong> the kernel function p <strong>of</strong><br />

the semigroup generated by L .<br />

To state a result on spectral convergence, we need some definitions. We define a linear map Tn : D½EKŠ!L2ðXn; nÞ<br />

by TnðuÞ ¼ fn ~uð¼ fn HK;uÞ; u 2 D½EKŠ. A sequence <strong>of</strong> functions un 2 L2ðXn; nÞ is said to L2 strongly (resp. L2 weakly) converge to a function u 2 L2ðK; Þ if there exists a sequence <strong>of</strong> functions vi in D½EKŠ such that<br />

limi!1 lim supn!1 kTnðviÞ unkL2ðXn; nÞ ¼ 0 (resp. if, for every v 2 L2ðK; Þ and any sequence <strong>of</strong> vn 2 L2ðXn; nÞ<br />

which L2-strongly converges to v, limn!1ðun; vnÞL2ðXn; nÞ ¼ðu; vÞL2ðK; Þ) (cf. [21]).<br />

Theorem 4.1. Under the conditions above, the following assertions hold:<br />

(i) Let fung be a sequence <strong>of</strong> functions un 2 L2ðXn; nÞ with supn kunkL2ðXn; nÞ < þ1 which L2-weakly converges to a<br />

function u 2 L2ðK; Þ. Then for any >0, wn ¼ R n; un uniformly converges to w ¼ðR ; uÞ and Enðwn; wnÞ tends<br />

to EKðw; wÞ as n !1, where R n; and R ; are respectively the resolvents <strong>of</strong> the Laplacians L and L .<br />

n<br />

(ii) The maps fn and hn approximate the kernel functions p and p in such a way that<br />

n<br />

lim<br />

sup<br />

n!1 t>0;x;y2Xn<br />

e ðtþ1=tÞ jp n ðt; x; yÞ ep ðt; fnðxÞ; fnðyÞÞj ¼ 0<br />

lim<br />

n!1 sup e<br />

t>0;a;b2Y<br />

ðtþ1=tÞ jp ðt; hnðaÞ; hnðbÞÞ n ep ðt; a; bÞj ¼ 0<br />

lim eS ða; fn hnðaÞÞ ¼ 0:<br />

n!1 sup<br />

a2Y<br />

(iii) For each i, the i-th eigenvalue <strong>of</strong> L converges to that <strong>of</strong> L n<br />

then the i-th eigenvalue diverges to infinity as n !1.<br />

in the sense that if N ¼ #K < þ1 and i N,<br />

Example Let ðV; E; rÞ be a finite network and E the associated form on ‘ðVÞ. Given a sequence <strong>of</strong> positive functions<br />

on V, we have a sequence <strong>of</strong> Morkovian forms E on ‘ðVÞ defined by<br />

n<br />

n<br />

E ðu; vÞ ¼Eðu= n n; v= nÞ; u; v 2 ‘ðVÞ:<br />

Suppose that n converges to a nonnegative function 1 on V. Let n (resp. 1) be measures on V given by<br />

nðuÞ ¼ P<br />

x2V uðxÞ nðxÞ2 (resp. 1ðuÞ ¼ P<br />

x2K uðxÞ 1ðxÞ2Þ, where K stands for the support <strong>of</strong> 1. Then as n !1,<br />

the Laplacian L <strong>of</strong> E in L n 2ðV; nÞ converges to the Laplacian L <strong>of</strong> the induced form E 1 K in L2ðK; 1Þ in the sense


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 67<br />

<strong>of</strong> Theorem 4.1. Moreover the form E n -converges to a Markovian form ðE1; D½E1ŠÞ on ‘ðVÞ defined by D½E1Š ¼<br />

fHK;u j u 2 ‘ðKÞg and E1ðHK;u; HK;uÞ ¼EðHK;u; HK;uÞ.<br />

4.3 Given two positive constants D and M, let BðD; MÞ ¼fðaiÞ 2‘ 2 j P1 i¼1ð1 þ i=DMÞa2i 4D2 compact subspace <strong>of</strong> ‘<br />

Mg. Then it is a<br />

2 .<br />

Now we consider a nonnegative quadratic form ðE; D½EŠÞ on a set X such that<br />

pffiffiffiffiffi it satisfies the properties (RF-i), (RF-ii)<br />

and (RF-iv), REðx; yÞ > 0 for all x; y 2 X with x 6¼ y, and the metric space ðX; REÞ<br />

is compact. Let us denote by SD;M pffiffiffiffiffi a<br />

set <strong>of</strong> such triplets ðX; E; REÞ<br />

endowed with Radon measures such that<br />

pffiffiffiffiffi diamðX; REÞ<br />

D < þ1; supp ¼ X; ðXÞ M:<br />

Then in view <strong>of</strong> (3), we see that the image J ðXÞ <strong>of</strong> the isometric embedding J<br />

in the compact subspace BðD; MÞ <strong>of</strong> ‘<br />

described in the last section is included<br />

2 .<br />

Recall here that the set <strong>of</strong> closed subspaces <strong>of</strong> a compact metric space is indeed compact with respect to the<br />

Hausdorff distance on the set (cf. e.g., [1]). It suggests us that the set SD;M is precompact in a sense from a view-point <strong>of</strong><br />

Laplacians (cf. [15, 19]). In fact, we have the following.<br />

Theorem 4.2. Let fðXn; En; nÞg be a sequence <strong>of</strong> SD;M. Then there exist a subsequence fðXm; Em; mÞg, a compact<br />

metric space ð ^X; ^SÞ, a nonnegative quadratic form ð^E; D½^EŠÞ on ^X satisfying (RF-i), (RF-ii) and (RF-iv), and a Radon<br />

measure on ^X such that<br />

(i) the sequence <strong>of</strong> compact metric spaces ðXm; S Þ converges to ð ^X; ^SÞ in the Gromov-Hausdorff sense via<br />

m<br />

approximating maps fm : Xm ! ^X and hm : ^X ! Xm;<br />

(ii) the sequenec <strong>of</strong> the forms Em -converges to ^E;<br />

(iii) letting ^Rðx; yÞ ¼supfjuðxÞ uðyÞj2 =^Eðu; uÞ ju 2 D½^EŠ; ^Eðu; uÞ 6¼ 0g, one has<br />

^Sðx; yÞ ^Rðx; yÞ 1=2<br />

lim inf<br />

m!1 RmðhmðxÞ; hmðyÞÞ 1=2 D; x; y 2 ^X:<br />

(iv) the image measure fm m weakly converges to ;<br />

(v) by using a linear map Tm : D½EKŠ!L2ðXm; mÞ defined as before, where K ¼ supp<br />

assertions as in Theorem 4.1.<br />

, one has the same<br />

Remarks (i) In view <strong>of</strong> the assertion (iii) above, the identity map <strong>of</strong> ð ^X; ^R 1=2Þ onto ð ^X; ^SÞ is continuous, but not<br />

homeomorphic in general. In fact, ð ^X; ^R 1=2Þ may be noncompact. See Example below.<br />

(ii) When each form Em is Markovian, so is the limit form ^E, and hence in this case, ^E is a resitance form on the set ^X<br />

and D½^EŠ is dense in Cð ^X; ^SÞ.<br />

(iii) Given a closed subset B <strong>of</strong> ð ^X; ^SÞ, we have the induced form ð^E B ; D½^E BŠÞ in CðBÞ; in particular, for any<br />

w 2 D½^E BŠ, there exists a unique function HB;u 2 D½^EŠ such that ^EðHB;u; vÞ ¼0 for all v 2 D½^EŠ vanishing on B.<br />

Example Let n be a subgraph <strong>of</strong> Zd generated by the set <strong>of</strong> vertices Vn ¼fðx1; ...; xdÞ jjxij n; i ¼ 1; ...; dg. Then it<br />

is known that the effective resistance <strong>of</strong> n satisfies<br />

c2 log n maxfRnðx; yÞ jx; y 2 Vng C2 log n<br />

if d ¼ 2, and<br />

0 < cd minfRnðx; yÞ jx; y 2 Vn; x 6¼ yg maxfRnðx; yÞ jx; y 2 Vng Cd<br />

if d 3, where cd and Cd are positive constants depending only on d (cf. [2]).<br />

Now we consider the case where d 3 and a sequence <strong>of</strong> measures n on Vn defined by nðuÞ ¼ P<br />

uðxÞ ðxÞ2<br />

x2Vn<br />

ðu 2 ‘ðVnÞÞ, where is a positive function on Zd such that P<br />

x2Zd ðxÞ2 < þ1. Then we have a compact metric space<br />

ðZd [ f1g; ^S; Þ to which the sequence <strong>of</strong> Gn with the measure n convegres in the sense <strong>of</strong> Theorem 4.1. However the<br />

effective resistance <strong>of</strong> Zd [ f1g provides it the discrete topology.<br />

5. Royden’s Compactification <strong>of</strong> an Infinite Network<br />

In the last example, the one-point compactification <strong>of</strong> the lattice Zd turns out to be the Royden compactification <strong>of</strong> Zd if d 3. Relevant to this example, we consider Royden’s compactification <strong>of</strong> a locally finite, connected and infinite<br />

network. We refer the reader to e.g., Soardi [26] and the references therein for some basic properties and results on the<br />

compactification <strong>of</strong> networks, and Sario–Nakai [25] for the case <strong>of</strong> a Riemannian manifold. See also Sal<strong>of</strong>f-Coste [23]<br />

and the references therein for some related topics and open questions.<br />

5.1 Let ¼ðV; E; rÞ be a locally finite, connected and infinite networks and ðEV ; D½EVŠÞ its resistance form. In what<br />

follows, we write ðE ; D½VŠÞ instead <strong>of</strong> ðEV ; D½EVŠÞ and denote by BD½VŠ the space <strong>of</strong> bounded functions in D½VŠ which<br />

is endowed with the norm: kukBD ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p<br />

E ðu; uÞ þ supx2V juðxÞj. Associated to the Banach algebra BD½VŠ with unit<br />

element 1, we have a compactification <strong>of</strong> V, called the Royden compactification <strong>of</strong> the network, that is described as<br />

follows: there exists a unique (up to homeomorphisms) compact Hausdorff space < such that < contains V as an<br />

open dense subset and every function in BD½VŠ can be continuously extended to < and BD½VŠ separates points in < .


68 KASUE<br />

The compact set @< ¼< n V is called the Royden boundary <strong>of</strong> the network and a distinguished part <strong>of</strong> the Royden<br />

boundary, called the harmonic boundary <strong>of</strong> the network, is defined by<br />

V ¼fx 2 @< j f ðxÞ ¼0 for all f 2 BD0½VŠg;<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

where D0½VŠ is the closure <strong>of</strong> the space <strong>of</strong> finitely supported functions in ðD½VŠ; E þ 2 p<br />

oÞ.<br />

Here we recall the fact that<br />

for any u 2 Cð Þ, there is a unique harmonic function h on V such that for all p 2 , limx2V!p hðxÞ ¼uðpÞ, and if<br />

u 2 D½VŠ, then h is a unique energy minimizer among functions in D½VŠ with the same boundary value as u (cf. [26],<br />

Chap. VI).<br />

Now as in Section 4, let us consider a measure ¼ P<br />

x2V ðxÞ2 P<br />

x on V such that ðxÞ > 0 for all x 2 V, ðVÞ ¼<br />

x2V ðxÞ2 < þ1, and R<br />

V R ðo; yÞd ðyÞ < þ1 for some o 2 V, where R denotes the effective resistance <strong>of</strong> the<br />

network. Let ¼f i j i ¼ 0; 1; 2; ...g be an orthonormal complete system <strong>of</strong> eigenfunctions <strong>of</strong> the Laplace operator<br />

L acting on L2ðV; Þ and J : V ! ‘ 2 the isometric embedding <strong>of</strong> the metric space ðV; S Þ into ‘ 2 as in Section 4. We<br />

denote by V the completion <strong>of</strong> ðV; S Þ. ThenVis compact if the effective resistance R is bounded in V, as seen in<br />

Section 4. Moreover in this case, setting<br />

D½V Š¼fu 2 CðV ÞjujV 2 D½VŠg; EVðu; uÞ ¼EVðujV; ujVÞ; u 2 D½V Š;<br />

we have a resistance form ðE ; D½V ŠÞ on V . The topology induced from the resistance metric may be strictly finer<br />

than that <strong>of</strong> the distance S , as is noted in the last example.<br />

Now we state the following<br />

Theorem 5.1. Let ¼ðV; E; rÞ be a locally finite, connected and infinite network. Suppose that the effective<br />

resistance <strong>of</strong> the network is uniformly bounded in V. Then the following assertions hold:<br />

(i) The spaces BD½VŠ and D½VŠ coincide and further the norms are equivalent.<br />

(ii) Given a measure on V as above, the metric space ðV ; S Þ is homeomorphic to the Royden compactification <strong>of</strong><br />

the network.<br />

(iii) ¼ @< .<br />

As a result <strong>of</strong> the last assertion (iii), we see that if in addition any harmonic function <strong>of</strong> bounded energy on V must be<br />

constant, then the Royden boundary <strong>of</strong> V consists <strong>of</strong> a single point.<br />

It would be interesting to give sufficient conditions on a network under which the resistance metric has bounded<br />

diameter (cf. Corollary 2.5).<br />

Let ¼ðV; EÞ be a locally finite, connected graph with all the resistances equal to 1. Assume that supx2V degðxÞ <<br />

þ1 and there exists a positive constant such that<br />

X<br />

uðxÞ 2<br />

EVðu; uÞ; u 2 D0½VŠ:<br />

x2V<br />

Then we have V ¼ @< , since is transient and BD0½VŠ ¼ff 2 BD½VŠ j f ðxÞ ¼0 for all x 2 Vg (cf. [26], p. 144).<br />

It would be asked if such a graph <strong>of</strong> bounded degree and a positive spectral gap is <strong>of</strong> bounded diameter with respect to<br />

the resistance metric when the graph has only one end, that is, it is connected at infinity.<br />

5.2 Let ðX; dÞ, ðX0 ; d0Þ be two metric spaces. A map : X ! X0 is a quasi-isometry from X to X0 if there are constants<br />

C1; ...; C5 such that:<br />

(i) For all x0 2 X0 , there exists x 2 X such that d0ðx0 ; ðxÞÞ C1;<br />

(ii) For all x; y 2 X, C2dðx; yÞ C3 d0ð ðxÞ; ðyÞÞ C4dðx; yÞþC5.<br />

A quasi-isometry is called in [13] and [14] a rough isometry.<br />

In what follows, we consider locally finite, connected graphs<br />

distances d<br />

¼ðV; EÞ <strong>of</strong> bounded degree with the geodesic<br />

R .<br />

Theorem 5.2. Let ¼ðV; E; dRÞ, 0 0 0 R ¼ðV ; E ; d 0Þ be two graphs <strong>of</strong> bounded degree, and suppose that there is a<br />

quasi-isometry : V ! V0 . Then the following assertions hold:<br />

(i) Relative to the resistance metrics <strong>of</strong> and 0 , the map induces a quasi-isometry (with different constants). In<br />

particular, the diameters on the resistance metrics are finite if so is either <strong>of</strong> them.<br />

(ii) A quasi-isometry : V ! V0 is continuously extended to a map from the Royden boundary <strong>of</strong> , < , to that <strong>of</strong><br />

0 , < 0, in such a way that sends homeomorphically the Royden boundary and the harmonic boundary <strong>of</strong> to those<br />

<strong>of</strong> 0 , respectively, that is, ð@< Þ¼@< 0 and ð Þ¼ 0.<br />

Let ¼ðV; E; rÞ be a locally finite, connected and infinite network. Given a point p, we have a positive regular Borel<br />

measure p, called the harmonic measure on the Royden boundary supported in the harmonic boundary (cf. [26],<br />

Chap. VI). In the same situation as in Theorem 5.3, it will be interesting to see how the image measure p <strong>of</strong> the<br />

harmonic measure on , p, can be compared with that on 0, p0, where p 2 V and p0 2 V0 . We would like also to<br />

know how we could characterize a map : V ! V0 which is continuously extended to a map : < !< 0 in such a<br />

way that induces a homeomorphism between the boundaries.<br />

Now we say that a complete, connected Riemannian manifold M is <strong>of</strong> bounded geometry if the Ricci curvature <strong>of</strong> M


<strong>Variational</strong> <strong>Convergence</strong> <strong>of</strong> <strong>Finite</strong> <strong>Networks</strong> 69<br />

is bounded below and the injectivity radius injðMÞ <strong>of</strong> M is bounded below by a positive constant. For such a<br />

Riemannian manifold M, we take a maximal set W <strong>of</strong> "-separated points, where " is a fixed positive number less than<br />

injðMÞ=2. We say that there is an edge with endpoints x; y 2 W if 0 < dR Mðx; yÞ 3", and the set <strong>of</strong> all such edges will<br />

be denoted by F. Then N ¼ðW; FÞ is a locally finite, connected graph <strong>of</strong> bounded degree and the inclusion <strong>of</strong> W into<br />

M induces a quasi-isometry between ðW; dR NÞ and ðM; dR MÞ. Let ¼ðV; EÞ be a locally finite, connected graph <strong>of</strong> finite degree and assume that there exists a quasi-isometry<br />

from V into M. Given a continuous function f on M, let us define a function f on V by<br />

Z<br />

1<br />

f ðqÞ ¼<br />

fdvM; q 2 V;<br />

VolðBð ðqÞ; 4"ÞÞ Bð ðqÞ;4"Þ<br />

and note that<br />

E ð f ; f Þ CEMð f ; f Þ<br />

if f has finite energy:<br />

Z<br />

EMð f ; f Þ¼<br />

M<br />

jdfj 2 dvM < þ1;<br />

where C is a positive constant independent <strong>of</strong> f (cf. [14]). Using this and a priori estimates on harmonic functions, we<br />

can prove the following<br />

Theorem 5.3. Let ¼ðV; EÞ and M be respectively a locally finite, connected and infinite graph <strong>of</strong> bounded degree<br />

and a complete, connected Riemannian manifold <strong>of</strong> bounded geometry. Suppose that ðV; dRÞ is quasi-isometric to<br />

M ¼ðM; dR MÞ and let : V ! M be a quasi-isometry. Then the following assertions hold:<br />

(i) For any u 2 Cð Þ, there is a unique harmonic function H on M such that for all p 2<br />

lim H ðxÞ ¼uðpÞ:<br />

x2V!p<br />

, one has<br />

Moreover if u continuously extends to a function in D½VŠ, then the harmonic function H has bounded energy and,<br />

letting h be a unique harmonic function in D½VŠ with the same boundary value u as H , one has<br />

C 1 EVðh; hÞ EMðH; HÞ CEVðh; hÞ;<br />

where C is a positive constant independent <strong>of</strong> u.<br />

(ii) For every harmonic function H <strong>of</strong> finite energy on M, H is extended to a continuous function on with values<br />

in R [ f 1g.<br />

Theorem 5.3 implies that possesses no nonconstant harmonic functions <strong>of</strong> finite energy if and only if so does M.In<br />

fact, Holopainen and Soardi show in [12] that this holds true for p-harmonic finctions <strong>of</strong> finite p-Dirichlet energy,<br />

where 1 < p < þ1. For a function u on a locally finite, connected network<br />

energy <strong>of</strong> u by<br />

¼ðV; E; rÞ, we define a p-Dirichlet<br />

E ðpÞ ðuÞ ¼ 1 X<br />

cðx; yÞ<br />

2 x y<br />

p 1 juðxÞ uðyÞj p where we set cðx; yÞ ¼1=rðx; yÞ. Setting<br />

;<br />

R ðpÞ juðxÞ uðyÞjp<br />

ðx; yÞ ¼sup<br />

E ðpÞ j u 2 ‘ðVÞ; E<br />

ðuÞ<br />

ðpÞ ( )<br />

ðuÞ > 0 ; x; y 2 V;<br />

we might expect an analogue <strong>of</strong> Theorem 5.1 relative to p-Dirichlet energy.<br />

Finally we consider a sequence <strong>of</strong> finite networks n ¼ðVn; En; rnÞ which satisfies the conditions in Theorem 3.4.<br />

Applying Theorem 3.3 to a sequence <strong>of</strong> the functionals E ðpÞ<br />

n , we have lower semi-continuous functionals EðpÞ on the<br />

space <strong>of</strong> continuous functions <strong>of</strong> the limit space Y in Theorem 3.4, which are -limits <strong>of</strong> fE ðpÞ<br />

g. We are interested in the<br />

n<br />

limit functionals; it may occur that EðpÞ is trivial.<br />

REFERENCES<br />

[1] Burago, D., Burago, Y., and Ivanov, S., A Course in Metric Geometry, GSM 33, Amer. Math. Soc., Providence, R.I., 2001.<br />

[2] Chandra, A. K., Raghavan, P., Ruzzo, W. L., Smolensky, R., and Tiwari, P., ‘‘The electrical resistance <strong>of</strong> a graph captures its<br />

commute and cover times,’’ Comput. Complex., 6: 312–340 (1996/1997).<br />

[3] Cheeger, J., Degeneration <strong>of</strong> Einstein metrics and metrics with special holonomy, Surveys in Differential Geometry, VIII, 29–<br />

73, International Press, 2003.<br />

[4] Colin de Verdière, Y., Pan, Y., and Ycart, B., ‘‘Singular limits <strong>of</strong> Schrödinger operators and Markov processes,’’ J. Operator<br />

Theory, 41: 151–173 (1999).<br />

[5] Dal Maso, G., An Introduction to –<strong>Convergence</strong>, Progress in Nonlinear Differential Equations and Their Applications 8,<br />

Birkäuser, Boston, 1993.


70 KASUE<br />

[6] Eells, J., and Fuglede, B., Harmonic Maps between Riemannian Polyhedra, Cambridge University Press, Cambridge, 2001.<br />

[7] Foster, R. M., The average impedance <strong>of</strong> an electrical network, Contributions to Applied Mechanics (Reissner Anniversary<br />

Volume), Edwords Bros., Ann Arbor, Mich. 333–340.<br />

[8] Fukaya, K., Hausdorff convergence <strong>of</strong> Riemannian manifolds and its applications, in Recent Topics in Differential and<br />

Analytic Geometry, 143–238, Adv. Stud. in Pure Math. 18, North-Holland, Kinokuniya, Amsterdam, Tokyo, 1990.<br />

[9] Fukushima, M., Oshima, Y., and Takeda, M., Dirichlet Forms and Symmetric Markov Processes, Walter de Gruyter, Berlin-<br />

New York, 1994.<br />

[10] Gromov, M., Structures métriques pour les variétés riemanniennes, rédigé par J. LaFontaine et P. Pansu, Cedic Fernand-<br />

Nathan, Paris, 1981.<br />

[11] Gromov, M., Metric Structures for Riemannian and non-Riemannian spaces, (with appendices by M. Katz, P. Pansu, and S.<br />

Semmes, ed. by J. LaFontaine and P. Pansu, English translation by S. M. Bates), Progress in Mathematics 152, Birkhäuser,<br />

Boston, 1999.<br />

[12] Holopainen, I., and Soardi, P. M., ‘‘p-harmonic functions on graphs and manifolds,’’ Manuscripta Math., 94: 95–110 (1997).<br />

[13] Kanai, M., ‘‘Rough isometries and the combinatorial arroximations <strong>of</strong> geometries <strong>of</strong> noncompact Riemannian manifolds,’’ J.<br />

Math. Soc. Jpn., 37: 391–413 (1985).<br />

[14] Kanai, M., ‘‘Rough isometries and the parabolicity <strong>of</strong> Riemannian manifolds,’’ J. Math. Soc. Jpn., 38: 227–238 (1986).<br />

[15] Kasue, A., <strong>Convergence</strong> <strong>of</strong> Riemannian manifolds and Laplace operators, I, Ann. l’Institut Fourier, 52 (2002); — II, to appear<br />

in Potential Analysis.<br />

[16] Kasue, A., ‘‘<strong>Convergence</strong> <strong>of</strong> metric measure spaces and energy forms,’’ Amer. Math. Soc. Transl., 215: 79–96 (2005).<br />

[17] Kasue, A., ‘‘<strong>Convergence</strong> <strong>of</strong> metric graphs and energy forms,’’ in preparation.<br />

[18] Kasue, A., ‘‘Dirichlet finite harmonic functions and points at infinity <strong>of</strong> a network,’’ in preparation.<br />

[19] Kasue, A., and Kumura, H., ‘‘Spectral convergence <strong>of</strong> Riemannian manifolds,’’ Tōhoku Math. J., 46: 147–179 (1994); — II,<br />

Tōhoku Math. J., 48: 71–120 (1996).<br />

[20] Kigami, J., Analysis on Fractals, Cambridge University Press, Cambridge, 2001.<br />

[21] Kuwae, K., and Shioya, T., ‘‘<strong>Convergence</strong> <strong>of</strong> spectral structures: a functional analytic theory and its appications to spectral<br />

geometry,’’ Comm. Anal. Geom., 11: 599–673 (2003).<br />

[22] Nakajima, H., ‘‘<strong>Convergence</strong> theorem <strong>of</strong> Einstein metrics and ALE spaces (Japanese),’’ Sūgaku, 44: 133–146 (1992).<br />

[23] Sal<strong>of</strong>f-Coste, L., Analysis on Riemannian co-compact covers, Surveys in Differential Geometry IX, 351–384, International<br />

Press, 2004.<br />

[24] Sakai, T., Riemannian manifolds with bounded Ricci curvature and their limits (Japanese), ‘‘Riemmanian manifolds and their<br />

limits,’’ Sūgaku Memoire, Math. Soc. <strong>of</strong> Japan, 2004.<br />

[25] Sario, L., and M., Nakai, Classification theory <strong>of</strong> Riemann surfaces, Springer Verlag, Berlin-Hidelberg-New York, 1970.<br />

[26] Soardi, P. M., Potential theory on infinite networks, Lec. Notes in Math. 1590, Springer, 1994.<br />

[27] Tetali, P., ‘‘Random walks and effective resistance <strong>of</strong> networks,’’ J. Theoret. Probability 4: 101–109 (1991).<br />

[28] Yamaguchi, T., ‘‘The development <strong>of</strong> convergence theory for Riemannian manifolds (Japanese),’’ Sūgaku, 47: 46–61 (1995).<br />

[29] Yamaguchi, T., ‘‘Collapsing Riemannian 4-manifolds (Japanese),’’ Sūgaku, 52: 172–186 (2000).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!