ArticlePDF Available

Spectrum of the Laplacian of an asymmetric fractal graph

Authors:

Abstract

We consider a simple self-similar sequence of graphs that does not satisfy the symmetry conditions that imply the existence of a spectral decimation property for the eigenvalues of the graph Laplacians. We show that, for this particular sequence, a very similar property to spectral decimation exists, and we obtain a complete description of the spectra of the graphs in the sequence.
Spectrum of the Laplacian of an asymmetric
fractal graph
Jonathan Jordan
University of Sheffield
22nd September 2004
Abstract
We consider a simple self-similar sequence of graphs which does not
satisfy the symmetry conditions which imply the existence of a spectral
decimation property for the eigenvalues of the graph Laplacians. We
show that, for this particular sequence, a very similar property to spectral
decimation exists, and obtain a complete description of the spectra of the
graphs in the sequence.
AMS 2000 subject classification: Primary 28A80
1 Introduction and definitions
1.1 Introduction
Many self-similar graphs, and related fractals, display a property known as
spectral decimation, that the spectrum of the Laplacian can be described in
terms of the iteration of a rational function f. Eigenvalues λof the Laplacian
at a given stage of the construction are related to eigenvalues µof the Laplacian
at the following stage of the construction by a relationship
λ=f(µ),(1)
Department of Probability and Statistics, University of Sheffield, Hounsfield Road,
Sheffield S3 7RH, U.K.
1
where fis a rational function on R, unless µis a member of a small exceptional
set,E. This was first observed for the specific case of the Sierpi´nski gasket graph
in [7], and this was given a rigorous mathematical treatment in [4, 10, 11]. In the
case of the Sierpi´nski gasket, using our definition of the Laplacian (see section
1.3), the function f(µ) = µ(5 4µ) and the exceptional set is {1/2,5/4,3/2}.
A generalisation of spectral decimation to a much larger class of self-similar
graphs, including the Vicsek set graph, appears in [6], in which a symmetry
condition is developed which, if satisfied, ensures that spectral decimation ap-
plies to the graph. Each self-similar graph in this class has a function fand
exceptional set Eassociated with it.
In this paper we consider a simple asymmetric self-similar graph which does not
satisfy the symmetry condition of [6]. In section 1.2 we define a sequence of
graphs (Gn)nN, which can be used to define a self-similar graph Gas for the
self-similar graphs in [6].
In section 2, we show that, for this example, a property similar to spectral
decimation exists, in which (1) is replaced by
2(1 λ)2=f(µ) (2)
where f(µ) is a quartic polynomial. Again there is an exceptional set of values of
µwhere the relationship does not necessarily hold. This is proved in Theorems
1 and 2.
Another common spectral property of self-similar graphs and related fractals
is that there are many eigenvalues of the Laplacian with high multiplicity and
Dirichlet-Neumann eigenfunctions, i.e. eigenfunctions which are zero on the
boundary. In [8] it is shown that the eigenvalues with Dirichlet-Neumann eigen-
functions dominate the spectrum in a large class of cases, that of the nested
fractals introduced in [5].
In section 3, we calculate the number of linearly independent eigenfunctions of
the Laplacian which are Dirichlet-Neumann or non-Dirichlet-Neumann, showing
that for large nthe spectrum of the Laplacian of Gnis dominated by eigenvalues
with Dirichlet-Neumann eigenfunctions. Together with the relationship between
the eigenvalues this allows us, in section 4, to describe the spectra of the graphs
in the self-similar sequence of finite graphs used in the construction of our graph.
This is stated in Theorem 6, which gives a complete description of the spectrum,
including the multiplicity of the eigenvalues and which eigenvalues are associated
with Dirichlet-Neumann and non-Dirichlet-Neumann eigenfunctions.
An example of a self-similar graph which does not satisfy the symmetry condi-
tions of [6], and for which spectral decimation appears not to apply, is associated
2
with the pentagasket, as described in [1], in which numerical approximations for
eigenvalues and eigenvectors are obtained, and some theoretical results are ob-
tained, showing how to construct eigenspaces of high multiplicity.
A more complicated method, using a rational map on a projective variety rather
than on R, which works for a larger class of self-similar graphs than that in [6]
including some for which spectral decimation does not apply, is described in [9].
However, our graph does not meet all the conditions described in section 1.1.1
of [9].
1.2 The graph
We define a self-similar sequence of finite graphs (Gn)nN, with V(Gn1)
V(Gn).
We start with G0, a single edge between two vertices 1 and 2. To construct
Gn+1, we note that each edge eof Gnconnects two vertices i(e) and j(e),
i(e)V(Gn1) and j(e)V(Gn)\V(Gn1). For each eof Gnwe obtain a
new vertex k(e)V(Gn+1), and the edges of Gn+1 consist of, for each edge e
of Gn, 2 edges connecting i(e) and k(e) and one edge connecting j(e) and k(e).
The following pictures show G1and G2:
} } }
xxxx
x
x
3
The same sequence of graphs can be obtained by using the framework of Def-
inition 5.2 of [6], with the model graph being identical to G1above, but with
conditions on the orientation to deal with the asymmetry.
When 1 m < n, the graph Gncontains 3nmsubgraphs isomorphic to Gm.
We will call these subgraphs m-cells. Using this, we can define a sequence
(˜
Gn)nNsuch that ˜
Gnis isomorphic to Gnand ˜
Gmis a subgraph of ˜
Gnfor
m < n. We then define the infinite graph G=Sn=0 ˜
Gn. This is analogous
to the Definition 5.5 of [6].
We define maps fi:V(Gn1)V(Gn), i = 1,2,3 mapping each vertex of Gn1
to the corresponding vertex in each (n1)-cell. We will label these so that f1
and f2correspond to the two parallel cells.
We note that this graph is similar to that described in [3], although in the
context of that paper the orientation of the cells is not important.
The graph Gnis bipartite with the two parts being V(Gn1) and V(Gn)\
V(Gn1).
1.3 The Laplacian
There are a number of different definitions of the Laplacian of a graph. The
definition of the graph Laplacian used in [6] is the generator matrix of a contin-
uous time random walk on the graph, while in the book [2] a related symmetric
matrix is used. However the eigenvalues of the different definitions differ by at
most a simple transformation.
For convenience in describing the eigenfunctions, we use the following definition
of the Laplacian: the Laplacian LGof a graph (which may have multiple edges
but with no loops) Gis a |V(G)|×|V(G)|matrix with, for a vertex iV(G),
LG(i, i) = 1, and, for i, j V(G) with i6=j,LG(i, j ) = ei,j
δi, where ei,j is
the number of edges linking iand jin Gand δiis the degree of vertex iin G.
This gives the same eigenvalues as the symmetric Laplacian described in [2],
and the eigenfunctions are the “harmonic eigenfunctions” described in [2]. Our
definition of the graph Laplacian differs from that in [6] only in that the sign of
each entry (and hence of the eigenvalues) is reversed.
4
2 The relationship between the eigenvalues
We set f(µ) = 9(µ1)49(µ1)2+ 2, so that (2) becomes
2(1 λ)2= 9(µ1)49(µ1)2+ 2.(3)
We first show how to construct eigenvalues µof LGn+1 from eigenvalues λof
LGnwhen λ /∈ {0,1,2}.
Given λand µ, we set
γ=3(µ1)22
1λ.(4)
Theorem 1. Given an eigenfunction xof LGnwith eigenvalue λ /∈ {0,1,2},
we can
Solve (3) for µto obtain 4 roots
For each possible µ, set γusing (4).
Then define x0by
x0
i=½xiiV(Gn1)
γxiiV(Gn)\V(Gn1)(5)
and, for a vertex k=k(e)V(Gn+1)\V(Gn), we set
x0
k=2xi(e)
3(1 µ)+γxj(e)
3(1 µ).(6)
Then x0is an eigenfunction of LGn+1 with eigenvalue µ.
Proof. To check this, we just calculate LGn+1x0. For iV(Gn1),
(LGn+1 x0)i=xi+1
δ(n)
iX
eE(n)
iµ2xi
3(µ1) +γxj(e)
3(µ1)
=xi+2xi
3(µ1) +γxi(1 λ)
3(µ1)
=xiµ3(µ1) + 2 + 3(µ1)22
3(µ1)
=µxi=µx0
i.
5
For jV(Gn)\V(Gn1),
(LGn+1 x0)j=γxj+1
δ(n)
jX
eE(n)
jµ2xi(e)
3(µ1) +γxj
3(µ1)
=γxj+γxj
3(µ1) +2xj(1 λ)
3(µ1)
=xjµ3(µ1)γ+γ+ 2(1 λ)
3(µ1)
Using (3) and (4),
2(1 λ) = 9(µ1)49(µ1)2+ 2
1λ
=(3(µ1)21)(3(µ1)22)
1λ
= (3(µ1)21)γ
and so
(LGn+1 x0)j=xjγµ3(µ1)+1+3(µ1)21
3(µ1)
=µγxj
=µx0
j
and finally, for kjV(Gn)\V(Gn1), which satisfies k=k(e) for some edge
eof Gn, we have
(LGn+1 x0)k=x0
k2
3x0
i(e)1
3x0
j(e)
=µ1
1µ1µ2
3xi+γ
3xj
=µµ2xi
3(1 µ)+γxj
3(1 µ)
=µx0
k,
so x0is indeed an eigenfunction of LGn+1 with eigenvalue µ.
Theorem 2. If µ /∈ {1,1 + p2/3,1p2/3,1 + p1/3,1p1/3}and λand
µsatisfy (3), then µis an eigenvalue of Gn+1 if and only if λis an eigenvalue
of Gn, with the same multiplicity.
Proof. If we have an eigenfunction x0of LGn+1 with eigenvalue µ6= 1, then, for
each edge eof Gn+1,
x0
k(e)=2x0
i(e)
3(1 µ)+x0
j(e)
3(1 µ)
6
so that, for each iV(Gn1),
x0
i(1 µ) = 1
δ(n)
iX
eE(n)
i
Ã2x0
i
3(1 µ)+x0
j(e)
3(1 µ)!
=2x0
i
3(1 µ)+1
δ(n)
iX
eE(n)
i
x0
j(e)
3(1 µ)
giving
x0
i(3(1 µ)22) = 1
δ(n)
iX
eE(n)
i
x0
j(e).(7)
Similarly for jV(Gn)\V(Gn1),
x0
j(1 µ) = 1
δ(n)
jX
eE(n)
j
Ãx0
j
3(1 µ)+2x0
i(e)
3(1 µ)!
=x0
j
3(1 µ)+1
δ(n)
jX
eE(n)
j
2x0
i(e)
3(1 µ)
giving
x0
j(3(1 µ)21) = 1
δ(n)
jX
eE(n)
j
x0
i(e).(8)
By the conditions of the theorem (1 µ)26=2
3. Then (7) implies that, for any
λ,
x0
i(1 λ) = 1
δ(n)
iX
eE(n)
i
1λ
3(1 µ)22x0
j(e)(9)
while (8) gives, if λ6= 1,
1λ
3(1 µ)22x0
j
(3(1 µ)22)(3(1 µ)21)
2(1 λ)=1
δ(n)
jX
eE(n)
j
x0
i(e).(10)
Then, if we set xi=x0
ifor iV(Gn1) and
xj=x0
j
1λ
3(1 µ)22,
(9) and (10) become
xi(1 λ) = 1
δ(n)
iX
eE(n)
i
xj(e)
7
and
xj
(3(1 µ)22)(3(1 µ)21)
2(1 λ)=1
δ(n)
jX
eE(n)
j
xi(e)
which imply that xis an eigenfunction of LGnwith eigenvalue λif
(1 λ) = (3(1 µ)22)(3(1 µ)21)
2(1 λ),
which is equivalent to the quartic (3).
This eigenfunction can be degenerate only if 1λ= 0, i.e. if either (1µ)2= 1/3
or (1 µ)2= 2/3.
The set {1,1 + p2/3,1p2/3,1 + p1/3,1p1/3}of values of µwhere
Theorem 2 does not apply plays a similar role to that of the exceptional set in
[6].
We note that the eigenvalues λand 2 λproduce the same values of µ, with
the same eigenfunctions. This is related to the bipartite nature of the graph
- in fact if xis an eigenfunction with eigenvalue λthen, following [2], we can
obtain an eigenfunction with eigenvalue 2 λby simply changing the sign of
xon V(Gn)\V(Gn1). These two eigenfunctions will then produce the same
new eigenfunction using the above construction.
We now consider the cases when Theorems 1 and 2 do not apply, i.e. when
λ∈ {0,1,2}or µ∈ {1,1 + p2/3,1p2/3,1 + p1/3,1p1/3}.
We note that if µ= 1 and λand µsatisfy (3), then λ= 0 or 2.
When λ= 1 and xi6= 0 for some iV(Gn1), we use the same method, but
with γ= 0 and the quartic (3) replaced by
(µ1)2= 2/3.(11)
We cannot use this method if xi= 0 for all iV(Gn1), because the con-
structed eigenfunction would be zero everywhere.
However, in the case where λ= 1 and xi= 0 for all iV(Gn1), we can
construct an eigenfunction x0by setting x0
i= 0 for iV(Gn1), and x0
i=xi
for iV(Gn)\V(Gn1). This gives eigenvalues µwith
(µ1)2= 1/3,(12)
using similar methods to those above.
8
3 Dirichlet-Neumann and non-Dirichlet-Neumann
eigenfunctions
The graph Gnhas vnvertices and enedges where v0= 2, e0= 1 and en= 3en1,
vn=vn1+en1. Hence en= 3nand vn=1
2(3n+ 3).
The following lemma provides a means of constructing Dirichlet-Neumann eigen-
functions, which are zero on the two boundary vertices 1 and 2.
Lemma 3. Let G0be a connected graph with mvertices, including distinguished
endpoints 1and 2, and let Gbe the graph formed by defining G1and G2to be two
identical copies of G0and connecting them in parallel by identifying their end-
points. Then the Laplacian of Ghas m2linearly independent eigenfunctions
which are zero on the endpoints.
The associated eigenvalues are the eigenvalues of the Laplacian LGrestricted to
the set {2j: 2 jm1}of vertices in G2.
Proof. We label the vertices of G01,2, . . . , m. Then we label the vertices in G
so that, for j3, vertex jin G0corresponds to vertex 2j3 in G1and vertex
2j2 in G2.
Now consider the Laplacian LG. For 2 j, k m1 we have
LG(1,2j1) = LG(1,2j)
LG(2,2j1) = LG(2,2j)
LG(2j1,2k1) = LG(2j, 2k)
LG(2j1,2k) = LG(2j, 2k1) = 0
and consider functions xsatisfying
x(1) = x(2) = 0
x(2j1) = x(2j) for 2 jm1.
Now
(LGx)(1) =
m1
X
j=2
(LG(1,2j1)x(2j1) + LG(1,2j)x(2j)) = 0
9
and similarly (LGx)(2) = 0, while
(LGx)(2j1) =
m1
X
k=2 LG(2j1,2k1)x(2k1)
=
m1
X
k=2 LG(2j, 2k)x(2k) = (LGx)(2j).
So the Laplacian LGpreserves vectors of this form, which form a vector space
of dimension m2, and it acts on them like the Laplacian restricted to the
interior vertices of G2. As the Laplacian is symmetric, there are m2 linearly
independent Dirichlet-Neumann eigenfunctions of the Laplacian.
If Gncontains a subgraph Gof this form, where the vertices of Gother than
the endpoints have no edges linking them to Gn\G, then we can take one of
the eigenfunctions xon Gconstructed by the above lemma and extend it to an
eigenfunction ˜xon Gnby setting
˜x(v) = x(v) for vV(G)
˜x(v) = 0 otherwise
If the endpoints 1 and 2 are both not in the interior of the subgraph G, then
this ˜xwill be Dirichlet-Neumann.
Proposition 4. The graph Gnhas at least 1
2(3n+3)2n1linearly independent
Dirichlet-Neumann eigenvalues.
Proof. The model graph contains parallel edges, so Gncontains a subgraph
consisting of two copies of Gn1with their boundary points identified as in
Lemma 3. This gives vn12 eigenfunctions. For each eigenfunction xobtained
thus, we have x(f1(v)) = x(f2(v)) and x(f3(v)) = 0 for each vV(Gn1).
Furthermore, given a Dirichlet-Neumann eigenfunction xof Gn1we can obtain
3 Dirichlet-Neumann eigenfunctions x1, x2, x3of Gnby extending them from
(n1)-cells to the whole graph i.e.
xi(fj(v)) = δij x(v) for all vV(Gn1), i, j = 1,2,3.
However, we only obtain 2 linearly independent eigenfunctions, because the
linear combination x1x2is of the form obtained using Lemma 3.
Hence, if lnis the number of Dirichlet-Neumann eigenvalues constructed by
these methods, it satisfies
ln= 2ln+1
2(3n1+ 3) 2
and l2= 1, which gives the result.
10
Let the total number of Dirichlet-Neumann eigenfunctions of Gnbe ln+ˆ
ln, so
that ˆ
lnis the number which are not constructed by the methods of Proposition
4
For this graph, we can also describe a set of non-Dirichlet-Neumann eigenfunc-
tions.
Proposition 5. The graph Gnhas 2n+ 1 linearly independent eigenfunctions
which are not Dirichlet-Neumann.
Proof. We consider the set of eigenfunctions xwhich are zero on the central
vertex of V(G1), i.e. they satisfy x(3) = 0 and, for each vV(Gn1),
x(f1(v)) = x(f2(v)) = 1
2x(f3(v)), a (vn11)-dimensional subspace. This
is preserved by the Laplacian of Gn, so we can find vn11 = 1
2(3n1+ 1)
linearly independent eigenfunctions satisfying these properties.
To exclude those which are Dirichlet-Neumann, this is equivalent to the con-
dition that x(1) = x(2) = 0. So a Dirichlet-Neumann eigenfunction satisfying
the above conditions reduces to a Dirichlet-Neumann eigenfunction on each
(n1)-cell. Hence there are 1
2(3n1+ 1) ln1ˆ
ln1= 2n1ˆ
ln1non-
Dirichlet-Neumann eigenfunctions satisfying the above conditions.
Now, given any eigenfunction xof Gn1, we can extend it to an eigenfunction
x0of Gnby setting x0(1) = 2x(1), x0(2) = x(1), x0(3) = 3x(2), x0(f1(v)) =
x0(f2(v)) = x0(f3(v)) = x(v) for v3, from the structure of the graph. This
means that each of the eigenfunctions constructed in Proposition 5 for Gn1
can be extended to a non-Dirichlet-Neumann eigenfunction of Gn, which will be
linearly independent of those already found (because x0(3) 6= 0). Inductively,
this also applies to those constructed for Gnm, m > 1.
The graphs are bipartite, so eigenfunctions (which are zero nowhere) for eigen-
values 0 and 2 exist as described in [2].
Hence the total number of linearly independent non-Dirichlet-Neumann eigen-
functions is at least 2n+ 1 Pn1
m=0 ˆ
ln. But we know that there are exactly
1
2(3n+ 3) linearly independent eigenfunctions. Hence
1
2(3n+ 3) 2n+ 1
n1
X
m=0
ˆ
lm+1
2(3n+ 3) 2n1 + ˆ
ln
and so ˆ
lnPn1
m=0 ˆ
lm. But ˆ
lm= 0 for m2, and hence for all m.
Hence the eigenfunctions constructed are all that exist, and there are 2n+ 1 lin-
early independent non-Dirichlet-Neumann ones, the remainder being Dirichlet-
Neumann.
11
4 The spectra of the graphs
We can now use the relationships between eigenvalues and the information on
Dirichlet-Neumann and non-Dirichlet-Neumann eigenfunctions to obtain a com-
plete description of the spectra of the graphs Gn.
Theorem 6. Set α(1)
1= 1,α(2)
1= 1 q2
3and α(2)
2= 1 + q2
3. We extend this
to define {α(n)
i; 1 i2n1}to be the 2n1values µsatisfying the quartic (3),
with λ=α(n1)
jfor some j.
Similarly set β(1)
1= 1,β(2)
1= 1 q1
3and β(2)
2= 1 + q1
3. We extend this to
define {β(n)
i; 1 i2n1}to be the 2n1values µsatisfying the quartic (3),
with λ=β(n1)
jfor some j.
Then
(a) If nm, then LGnhas a non-Dirichlet-Neumann eigenfunction with
eigenvalue α(m)
i,1i2m1. Together with eigenfunctions with eigen-
values 0and 2, this describes the non-Dirichlet-Neumann spectrum of LGn.
(b) If nm+ 1, then LGnhas 1
2(3nm1) linearly independent Dirichlet-
Neumann eigenvalues with eigenvalue β(m)
i, for 1i2m1.
Proof. We note that G1has a non-Dirichlet-Neumann eigenfunction with eigen-
value 1. As described in the proof of Proposition 5, this can then be extended
to give a non-Dirichlet-Neumann eigenfunction xwith eigenvalue 1 for each Gn,
n1.
Because this eigenfunction is non-Dirichlet-Neumann, it has xi6= 0 for at least
some iV(Gn1). Hence the construction of eigenfunctions of Gn+1 with
eigenvalues µsatisfying (1 µ)2=2
3produces non-degenerate eigenfunctions,
which are also non-Dirichlet-Neumann. So Gn+1 has non-Dirichlet-Neumann
eigenfunctions with eigenvalues 1 ±q2
3.
We have already shown that Gnhas a non-Dirichlet-Neumann eigenfunction
with eigenvalue α(2)
i, 1 i2n1, for n2. Now, if Gn1has a non-Dirichlet-
Neumann eigenfunction with eigenvalue α(m1)
ifor each 1 i2m2, then the
construction of eigenfunctions gives us a non-Dirichlet-Neumann eigenfunction
of Gnwith eigenvalue α(m)
ifor each 1 i2m1. Using this inductively, we
12
find that Gnhas a non-Dirichlet-Neumann eigenfunction with eigenvalue α(m)
i,
1i2m1, for nm.
Along with the eigenfunctions with eigenvalues 0 and 2, this gives us all 2n+ 1
non-Dirichlet-Neumann eigenfunctions from Proposition 5, and hence completes
the proof of (a).
We now consider the Dirichlet-Neumann eigenfunctions. We note that G2has
a Dirichlet-Neumann eigenfunction with eigenvalue 1. Such an eigenfunction
is zero on V(G1), so our main construction produces a degenerate eigenfunc-
tion. However the alternative construction with (1 µ)2=1
3does produce two
eigenfunctions of G3. Hence, as there are 1
2(3n+ 3) 2n1 Dirichlet-Neumann
eigenfunctions of Gn, we can use the constructions to obtain 3n+ 3 2n+1 2
Dirichlet-Neumann eigenfunctions, with eigenvalues other than 1, of Gn+1.
We now show that, for n2, Gnhas 1
2(3n11) linearly independent Dirichlet-
Neumann eigenfunctions with eigenvalue 1. This is the case for n= 2. For each
such eigenfunction of Gn, we can construct 3 of Gn+1 using the methods in the
proof of Proposition 4.
Assuming that Gnhas 1
2(3n+3)2n1 linearly independent Dirichlet-Neumann
eigenfunctions of which 1
2(3n11) have eigenvalue 1, we have 1
2(3n3) linearly
independent Dirichlet-Neumann eigenfunctions of Gn+1 with eigenvalue 1, and
3n+ 3 2n+1 2 with other eigenvalues. However, we know from Proposition 4
that there are 1
2(3n+1 +3)2n+1 1 in total, and 1
2(3n3)+3n+ 3 2n+1 2 =
1
2(3n+1 + 3) 2n+1 2.
The one unexplained eigenfunction must also have eigenvalue 1, because if it had
eigenvalue λ6= 1 we would also have an unexplained eigenfunction with eigen-
value 2 λ. Hence we have 1
2(3n1) linearly independent Dirichlet-Neumann
eigenfunctions of Gn+1 with eigenvalue 1, giving the result by induction.
We know that, for n2, Gnhas 1
2(3n11) linearly independent Dirichlet-
Neumann eigenfunctions with eigenvalue 1. We now use our constructions m
times to show that, for nm+ 1, Gnhas 1
2(3nm1) linearly independent
Dirichlet-Neumann eigenfunctions with eigenvalue β(m)
i, for 1 i2m1. This
completes the proof of (b).
13
5 Reversing the orientation
We remark that very similar results can be obtained if we reverse the orientation
of the model graphs in the definitions of Section 1.2. Because of the asymmetry
this gives a different self-similar sequence of graphs. Eigenvalues λand µof
Laplacians successive members of the sequence are related by the same equation
(3) when λ /∈ {0,1,2}, but the two equations (11) and (12) for µwhen λ= 1
are reversed. This has the effect that, in Theorem 6, the roles of α(n)
iand β(n)
i
are reversed.
References
[1] B. Adams, S. A. Smith, R. S. Strichartz, and A. Teplyaev. The spectrum
of the Laplacian on the pentagasket. In Trends in Mathematics: Fractals
in Graz 2001, pages 1–24. Birkhauser, 2003.
[2] F. R. K. Chung. Spectral Graph Theory. Number 92 in CBMS Regional
Conference Series. AMS, Providence, Rhode Island, 1997.
[3] C. D. Essoh and J. Bellisard. Resistance and fluctuation of a fractal network
of random resistors: a non-linear law of large numbers. J. Phys. A, 22:4537–
4548, 1989.
[4] M. Fukushima and T. Shima. On a spectral analysis for the Sierpi´nski
gasket. Potential Anal., 1:1–35, 1992.
[5] T. Lindstrøm. Brownian motion on nested fractals. Mem. Amer. Math.
Soc., 420, 1990.
[6] L. Malozemov and A. Teplyaev. Self-similarity, operators and dynamics.
Mathematical Physics Analysis and Geometry, 6:201–218, 2003.
[7] R. Rammal and G. Toulouse. Random walks on fractal structures and
percolation clusters. J. Physique Letters, 44:L13–L22, 1983.
[8] C. Sabot. Pure point spectrum for the Laplacian on unbounded nested
fractals. J. Funct. Anal., 173:497–524, 2000.
[9] C. Sabot. Spectral properties of self-similar lattices and iteration of rational
maps. em. Soc. Math. Fr., 92:104pp, 2003.
[10] T. Shima. On eigenvalue problems for the random walks on the Sierpi´nski
pre-gaskets. Japan. J. Indust. Appl. Math., 8:127–141, 1991.
[11] T. Shima. On eigenvalue problems for Laplacians on p.c.f. self-similar sets.
Japan. J. Indust. Appl. Math., 13:1–23, 1996.
14
... We investigate two examples of the spectral theory of the Laplacians of fractal graphs in the context of the general theory developed by Sabot. One example is the pentagasket, where the related problem of the spectral theory of the Laplacian on the fractal itself is investigated in [1], and the other is related to the variant on spectral decimation found in [5]. Although the graphs defined in [5] do not quite fit the definitions in Section 1.1.1 of [11], we will see that much of the theory does apply. ...
... One example is the pentagasket, where the related problem of the spectral theory of the Laplacian on the fractal itself is investigated in [1], and the other is related to the variant on spectral decimation found in [5]. Although the graphs defined in [5] do not quite fit the definitions in Section 1.1.1 of [11], we will see that much of the theory does apply. ...
... For example, for the pentagasket N 0 = N = 5, the equivalence relation R is given by (1, 3)R(2, 5), (2, 4)R(3, 1), (3,5)R(4, 2), (4, 1)R(5, 3) and (5, 2)R (1,4), the α i are all equal, and the function β is simply β(j) = j. The first few graphs in the resulting sequence are shown in Figure 1. Figure 1: The first few graphs, Γ (0) , Γ (1) , Γ (2) and Γ (3) , in the sequence associated with the pentagasket. ...
Article
Full-text available
We consider the spectra of the Laplacians of two sequences of fractal graphs in the context of the general theory introduced by C. Sabot [Mém. Soc. Math. Fr., Nouv. Sér. 92, vi, 104 p. (2003; Zbl 1036.82013)]. For the sequence of graphs associated with the pentagasket, we give a description of the eigenvalues in terms of the iteration of a map from (ℂ 2 ) 3 to itself. For the sequence of graphs introduced in a previous paper by the author, we show that the results found therein can be related to Sabot’s theory.
Article
Full-text available
We establish the pure point spectrum of the Laplacians on two point self-similar fractal graphs. All eigenvalues have infinite multiplicity and a countable system of orthonormal eigenfunctions with compact support is complete in the corresponding Hilbert space.
Article
Full-text available
The authors study rigorously the resistance and fluctuation of resistance of a large deterministic fractal lattice in the limit of an infinite number of resistors. They give estimates on corrections to the effective medium approximation of the total resistance. They prove scaling laws for the relative fluctuation, and prove that the normalised relative fluctuation converges in distribution to the standard normal variable. This is a kind of non-linear law of large numbers.
Article
The spectrum of the fully symmetric Laplacian on the fractal pentagasket is studied by theoretical and experimental methods. We show how to construct derived eigenspaces of high multiplicity for both Dirichlet and Neumann spectrum starting from primitive Neumann eigenspaces. We prove that both spectra may be parceled into groups of five dimensional spaces which decompose in a prescribed way under the D 5 symmetry group. We show that D 5 invariant eigenfunctions possess additional local symmetries that in particular force them to be constant along the Cantor set bordering the inner deleted pentagon. Numerical approximations for eigenfunctions and eigenvalues obtained using the finite element method are reported. We formulate several conjectures based on this data. More data can be found at http://www.mathlab.cornell.edu/~sas60/
Article
We prove for the class of nested fractals introduced by T. Lindstrøm (1990, Memoirs Amer. Math. Soc.420) that the integrated density of states is completely created by the so-called Neuman–Dirichlet eigenvalues. The corresponding eigenfunctions lead to eigenfunctions with compact support on the unbounded set and we prove that for a large class of blow-ups the set of Neuman–Dirichlet eigenfunctions is complete, leading to a pure point spectrum with compactly supported eigenfunctions. This generalizes previous results of H. Teplyaev (1998, J. Funct. Anal.159, 537–567) on the Sierpinski Gasket. Our methods are elementary and use only symmetry arguments via the representations of the symmetry group of the set.
Article
A complete description of the eigenvalues of the Laplacian on the finite Sierpinski gasket is presented. We then demonstrate highly oscillatory behaviours of the distribution function of the eigenvalues, the integrated density of states (for the infinite gasket) and the spectrum of the Laplacian on the infinite gasket. The method has two ingredients: the decimation method in calculating eigenvalues due to Rammal and Toulouse and a simple description of the Dirichlet form associated with the Laplacian.
Article
We formulate and study a strong harmonic structure under which eigenvalues of the Laplacian on a p.c.f. self-similar set are completely determined according to the dynamical system generated by a rational function. We then show that, with some additional assumptions, the eigenvalue counting function ρ(λ) behaves so wildly that ρ(λ) does not vary regularly, and the ratior(l)/lds /2\rho (\lambda )/\lambda ^{d_s /2} is bounded but non-convergent as λϖ∞, whered s is the spectral dimension of the p.c.f. self-similar set.
Article
We work with increasing finite setsV m called pre-gaskets approximating the finite Sierpinski gasket located inR N−1 (N ≥ 3). The eigenvalues of the discrete Laplacian onV m under the Dirichlet and Neumann boundary conditions are completely determined using the decimation method due to Rammal.
Article
We construct a large class of infinite self-similar (fractal, hierarchical or substitution) graphs and show, under a certain strong symmetry assumption, that the spectrum of the Laplacian can be described in terms of iterations of an associated rational function (so-called 'spectral decimation'). We prove that the spectrum consists of the Julia set of the rational function and a (possibly empty) set of isolated eigenvalues which accumulate to the Julia set. In order to obtain our results, we start with investigation of abstract spectral self-similarity of operators.