Applied Mathematics, 2011, 2, 965-974
doi:10.4236/am.2011.28133 Published Online August 2011 (http://www.SciRP.org/journal/am)
Copyright © 2011 SciRes. AM
Spectral Dependence of the Degree of Localization in a 1D
Disordered System with a Complex Structural Unit
Gleb G. Kozlov
Institute of Physics, Saint-Petersburg State University, Saint-Petersburg, Russia
E-mail: gkozlov@photonics.phys.spbu.ru
Received April 28, 2011; revised May 19, 2011; accepted May 26, 2011
Abstract
We analyze the spectral distribution of localisation in a 1D diagonally disordered chain of fragments each of
which consist of coupled two-level systems. The calculations performed by means of developed pertur-
bation theory for joint statistics of advanced and retarded Green’s functions. We show that this distribution is
rather inhomogeneous and reveals spectral regions of weakly localized states with sharp peaks of the local-
ization degree in the centers of these regions.
m
Keywords: Anderson Localization, Spectral Distribution, Green’s Function, 1D Disordered Chain
1. Introduction
The 1D models were traditionally studied in the theory of
solid state for obtaining qualitative data, which were then
used to analyze more realistic 3D models. For example,
the exactly solvab le Kronig-Penny model, which demon-
strates the most important qualitative properties of the
translationally symmetric systems, such as the band type
of the energy spectrum and possibility of the wave-
vector-based classification of states, provided a basis for
the modern theory of crystalline media. As the second
(and far from the last) example may serve the Ising
model in the theory of phase transitions, in which the
exact analysis of the 1D model has shown a decisive role
of dimentionality of the system for observation of its
critical behavior. These examples, pertaining to the time
when the solid state physics was at the stage of
accumulation of knowledge show that analysis of even
abstract 1D models (which do not correspond to any real
physical system) may give an important qualitative
information.
Note that, at present, there are all grounds to believe
that the 1D models appear to be also important in other
respects. Up-to-date technologies of material production
and experimental techniques make it possible to create
and study objects (quantum superlattices, quantum wires,
J-aggregates, optical waveguiding fibers, etc.) that can be
described quantitatively using such models. Still, the
heuristic significance of the 1D models can be consi-
dered, nowadays, as the most important.
Among models of the physics of disordered systems,
which, at present is also at the stage of accumulation of
information, the 1D models occupy a particularly impor-
tant place. At this stage, a consistent mathematical
analysis of even an abstract model, capable of giving
reliable and non-trivial results, is of interest. As
examples of such an analysis may serve publications
[1-4].
As is known [5,6], an important property of the
homogeneously disordered systems is the appearance of
localized states in their single-particle spectrum. Since
these states play a decisive role in calculations of the
charge transfer in disordered systems, any information
about them is considered to be valuable. For getting
information of this kind, in [7], a consistent perturbation
theory for joint statistics of the advanced and retarded
Green’s functions was proposed. This theory allows one
to calculate distribution of the degree of localization of
states (in the sense of the Anderson criterion [5,6]) as a
function of their energy (called, hereafter, spectral
dependence of the degree of localization). The calcu-
lation presented in [7] refers to the classical case of a
disordered chain with a simplest structural unit, namely,
a two-level system. The goal of this paper is, first, to
generalize the method proposed in [7] to the case of a
disordered chain with a more complicated structural unit
—a segment comprised ofcoupled two-level systems,
and, second, to calculate spectral dependence of the
degree of localization in such a system. The main quanti-
tative result of the paper is derivation of an analytical
m
966 G. G. KOZLOV
formula for the above degree of localization. The for-
mula thus derived shows that distribution of the degree
of localization over the energy spectrum, for the model
with , is essentially inhomogeneou s and is charac-
terized by appearance of energy points in which
the states are virtually delocalized. This result provides
grounds for the following qualitative conclusion: Spec-
tral distribution of the degree of localization, in the 1D
systems, may be essentially inhomogeneous and may
reveal strongly pronounced maxima and minima. This
behavior of the degree of localization qualitatively
differs from that for the case of con sidered in [7 ].
>1m1m
m
=1m
The fact that such a model with can exhibit
unusual behaviour was described in [8] where for a
binary disordered chain of dimers the indications of
delocalization were founded.
=2m
The above model with complex structural unit can also
be used as a simplest one describing the correlated
disorder with parameter playing the role of corre-
lation length. The optical properties of similar random
system were studied in [9] and it was shown that so
called factor of optical amplification for such system
display nonmonotonic dependance on correlation length.
The important role of correlations was pointed out in [10]
where the authors put forward arguments in favor of
appearance of delocalized states in the case of spectral
density of correlation function having the form

1Sk k
.
Note that, in our opinion, the model of disordered
chain with a complex structural unit, considered below,
is, at any rate, not more abstract than the classical model
with , whereas, in qualitative respect, it is, perhaps,
even closer to real systems than the classical model.
=1m
2. Formulation of the Problem, and the Main
Results
Let us pass to quantitative formulation of the problem.
Consider a 1D chain of two-level atoms consisted of
fragments with the length . We assume that splittings
of all the atoms within a fragment are the same, while the
disorder is provided by randomness of the splitting s from
fragment to fragment. The splittings corresponding to
different fragments will be considered as independent
random quantities with a known distribution fun ction .
Such a system differs from the standard one by that the
structural unit is a fragment of coupled two-level
systems, rather than a single two-level system. Thus, the
Hamiltonian matrix
m
P
m
H
, in the studied model, will have
a usual form
,, ,1,1
=,,
rrrr rrrrr=1,,
H
rr N
 
 

 (1)
where the diagonal elements r
correspond to the
fragments described above. If, for example, , then
=2m
123456
===


,,  and the elements
135


P
are the random independent quantities with
the distribution function
. The off-diagonal ele-
ments equal to unity determine the energy scale. In what
follows, we will imply the thermodynamic limit
.
N
D
t
0
For this model, we consider the following problem.
Let the edge atom is excited at , and we have to
find the probability that this atom remains in the
excited state at . Methematically, it means that
the initial state of the system is described by the wave
function (vector-column) with the components
=0t
,
0=
rrN
, and we have to fi n d

2
=N
Dt
N
(the angle brackets indicate averaging over realizations
of the random splitting ,=
r1,,r
 ). Time depen-
dence of the wave function is given by the formula
=exp 0tiHt
D. It follows herefrom that the
quantity , we are interested in, can be expressed in
terms of eigenvectors
and eigenvalues
,=1,,EN


of matrix (1) as follows

2
22
NN


4
=
=exp
lim
=
N
t
N
Dt
iEE t




(2)
The quantity is known to be crucial in the theory
of Anderson localization [5-7], according to which the
nonzero indicates presence of localized eigenfunctions
(in the sense of Anderson criterion) among those of
Hamiltonian (1). To judge about the degree of locali-
zation of the eigenvectors of (1) in the spectral range
D
D
,UU dU, we have introduced, in [7], the par-
ticipation function
WU, defined by the relationship

4
[,EU

]
d= N
UdU
WU U
(3)
Obviously,
dUU=DW
. It was shown in [7] that,
when all the states, within the interval
,UU dU, are
delocalized, then
=0WU . Otherwise,
WU is
nonzero. One may easily check, that the function
WU
has also the sense of average squared module of the
frequency spectrum of oscillations of the wave function
at the edge site1. In the presence of localization, these
1In the vibration-related problems, when the dynamic matrix has the
form close to (1), the function W(U) is connected with the squared mod-
ule of the frequency spectrum of vibrations of the edge oscillator.
Copyright © 2011 SciRes. AM
G. G. KOZLOV
967
oscillations do not decay, which corresponds to nonzero
value of .

WU
As was already mentioned, the physics of disordered
systems is, at present, at the initial stage of its evolution,
when testing theoretical results in a real experiment
cannot look convincing. For this reason, as an especially
important property of the discrete models described by
the Hamiltonian of type (1), should be considered
possibility of their simple numerical analysis. Fantastic
capabilities of the contemporary computers allow one to
diagonalize matrices of type (1) for a reasonable time
and to observe, in such a numerical experiment, the
quantities (2) and (3) at and more, checking,
in this way, the appropriate theoretical predictions with a
high degree of reliability. The main result of this paper is
the following formulas for the participation function
and the quantity , which are applicable to the
1D random chain with the complex structural unit
described above:
1000N
D

WU




22
2
23
2
23
2
2
=4
4
arctg ,
sin
=2
UM
WU mU
U
mO
U
M
DO
m
 











(4)
Here, the quantity is the parameter of the site
energy distrib u t i on fu nct i o n, whi ch we take in the form
  
=,Px pxpx>0,
 
22
d=1,d=0,d=pxxpxxxpxx xM
 
It follows from Equation (4) that, for the energies defined
by the condition
22
4
arctg=,= 21,
<2,
nn
n
n
Enn
Etg
Emm
Eninteger



 

 

(5)
the participation function
n
WE vanishes to within
the terms , i.e., the states in these energy points
appear to be, in the considered approximation,
delocalized. This result looks strange because it is
considered proven [5] that, in a 1D homogeneous
random system, all the states are localized. In this
connection, we performed analysis of the next term in
expansion of the function
3
WU in powers of
. The
analysis has shown that this term diverges at
(5). Due to this counter-directed behavior of
contributions of different orders, Equation (4), at small
disorder (i.e., at small ), should work well at any
energy except for small regions in the vicinity of . In
these points (where the correction (4) vanishes),
narrow peaks of the participation function
=n
UE
n
E
2
WU ,
corresponding to divergence of the next correction,
should be observed. The numerical experiment described
in the last section of the paper completely confirms this
conclusion.
3. Green’s Function Statistics
It can be easily shown that the mean square of the wave
function module at the edge site

2
Nt (which is
connected with quantity (2) we are interested in by the
relation

2
=N
D ) can be calculated in the
following way:
 

2
12 1
2
0
1,2
1122
1
=ddexp
lim 4
NV
tUUUt
UiV UiV



 
2
iU


(6)
where
is the edge Green’s function (EGF) for
Hamiltonian (1) :

2
N
E
 
(7)
To calculate the mean value of the product of fwo
Green’s functions entering Equation (6) at different
complex energies 111
=UiV
and 222
,
one has to know their joint distribution function (statis-
=UiV
tics)
112 2

112 2112
dddd
2

is the
probability that Re
,
iiii
d
 
 and Im
,
iii
di

 
=1,2i, ). We will usually write
down the function
as a function of two complex
variables =
ii
zi,=1,2
i
i

:

12
zz
112 2=

.
We derive equation for this function using the method of
[7,11].
Consider the chain described in Introduction, whose
structural unit is a fragment consisting of two-level
systems. Let be the number of the edge two-level
system of the edge fragment and
m
N
N
—the EGF of such
a chain. If we add to this chain one more fragment
consisted of sites with the splitting
m
, then the EGF
, corresponding to the edge site of this new chain can
be expressed through
:

1
=R


(8)
The explicit form of the function

1
R

=1m
entering
this equation, for an arbitrary , will be given below,
and now, we remind that, for the case ,
m
1=1
|=1
m
R

 [5,7,11]. Relationship (8)
Copyright © 2011 SciRes. AM
968
G. G. KOZLOV
allows us to express statistics of the edge Green’s
functions of the chain with the added fragment
(
112 2
,,,
x
yxy
) through the function
112 2


1
2
i
i


N
of the initial chain as follows

1122
11 2
=ddd d
xyxy
xReR
xReR
  




  


2 1122
11
1111
11
11
2222
22
dP
i yImR
i yImR

 






 


 

1
2
(9)
It is clear that, in the thermodynamic limit ,
should coincide with
. By calculating the integrals
with
-functions in (9), we obtain the following
equation for the stead y-state joint d istribution fun ction o f
EGF:
 
 

12
12
12
dd
ddd
,
RR
P
RR



 
 
 
 
22
12
12
12
,=

(10)
In this equation, symbol denotes transformation
inverse to (8), and the joint statistics
R
is written as a
function of two complex arguments.
The fact that the Green’s functions entering Equation
(6) have vanishingly small imaginary parts of the energy
arguments can be used to reduce the problem to analysis
of the equation much simpler than (10) [7]. The average
product of the two Green’s functions, entering Equation
(6), can be written in the form of the sum of four terms:
 
 
1122
1122112 212121221
12121212
=dddd
UiV UiV
x
yxyxyxyxxyy ixyxy
xxyyixy iyx


 


 
(11)
It was shown, in [7], that it suffices to calculate the
term 12
x
y and to multiply the result by 4. Using
Equation (10), we can write the quantity 12
x
y, we are
interested in, in the form

  


12112212 1 122
22
12
12
12
11112 22 2
11 22
12 1 122
=dddd
dd
=d dd
,, ,
ddddd
xyxyxyxyx y xy
RR
P
uxyvxyuxyvxy
xyx y xy

 

 
 
 
(12)
where the real functions and
,uxy
,vxy
and
the Jacobians entering Equations (10) and (12) are
determined by the relations

,,uxyivxyRxiy


(13)

222
d==
d
=
ii
iii
ii
ii
ii
ii
iii
uu
Ru
xy
vv xx
xy
xiy



 
 
 






 


i
i
v
(14)
( =1,2)i. In Equation (14), we used the Cauchy-Riman
relations. In integral (12), we change the variables:

 
11111
11
2222
22
=,=
=,=
x uxyyvxy
1
22
,
x
uxyyvxy


 
 


(15)
Since Jacobians of these transforms coincide with (14),
for 12
x
y we have:
 

121 1221122
1112 22
=ddddd
x
yxyxyPxyx
xxyyxy

 

y
(16)
with the form of
111
x
xy
 and
222
y
xy
 being deter-
mined by the function

1
R
 (8):
 
11
1112 2
12
=,=
2
x
Re RxiyyIm Rxiy


 



 
(17)
As shown in [7], to calculate 12
x
y at vanishingly
small imaginary parts of the energy arguments (i.e., at
), we may replace, in (16),
0
i
V
by its limiting
value 0
, which corresponds to pure real energies
ii
U
, whereas in (17), one has to take into
account imaginary parts i
V of the energy arguments
(taking advantage, when possible, of their smallness).
The reasoning similar to that presented in [7] shows that
the function
( =1,2)i
0
can be represented in the form:

011221122 =
1,2 1,2
12 1 2
12
=
U
UU
xyxyxyxy
x
xyy


(18)
where the function
12
,
UU
12
x
x
controls statistics of
the real Greens functions and meets the equation that is
much simpler than (10):
 
 
121 2
1212 12
12
12
12
,=d ,
dd
dd
UUUU UU
UU
x
xP RxRx
RxRx
xx







(19)
Copyright © 2011 SciRes. AM
G. G. KOZLOV
969
Thus, in Equations (16) and (17), we may set 1,2 ,
since Equation (18) for 0
=0y
contains the relevant
-functions. Since transformation (8) for real
and
gives real result, we may write, for
1
x
:

1
1
1
=U1
x
Rx
(20)
An essentially different situation takes place for the
quantity 2 from Equation (17), which, with allowance
for the above remarks, has the form
y

1
2
=UiV
yImR x

1
2
R
2
22 . The imaginary part of this
expression, in the general case, tends to zero, because the
transform
 is real at . The exceptions are
peculiarities of the transform
20V
1
2
R
 , i.e., the values of
its argument 0
at which 0
2
U.2 At this
stage of the calculation, we need the explicit form of the
transformation


1
Rx
1
R
x [12]:
 
 
1sinsin 1
=sin1 sin
mmx
abx
Rx cgx
mmx







 (21)
where

2
4
=arctg







The peculiarity 000
=zcg i

 of the
function 
1
Rz
corresponds to zero denominator in
(21), and, in the limit Im , we are interested in,
its imaginary part tends to zero. Bearing this in mind and
using the same reasoning as in Footnote, one may make
sure that Equation (17), for , can be rewritten in the
form:
0
2
y

1
22
02
20
2
2
2
2
=lim
=
VV
UU UUU
U
U
yImRx
ag bcc
x
g
g
UU







(22)
From Equations (20) and (21), we obtain the following
expression for 1
x
:
1
0
1=,=
UU
VUU
abx
xU
cgx
By replacing, in (16),

112 21212
12 12
UU UU
x
yxyxxyy

(Equation (18)) and by substituting in to it Equations (22)
and (23) for and
2
y1
x
, we obtain

2
12 12 2
2222 11
2
11
2
=dd ,U
UU U
UUUU UU
UU
U
c
xyxPx g
agbc abx
cg
g
 
x

 






(24)
To further simplify this expression, we introduce a
new variable:

111
111
==
UU
UUU
ab
zRx cg




x
x
Then
 
1
1
d
=,d=
d
U
U
Rz
d
x
Rzxz
z (25)
Then, Equation (24) can be transformed into the form
 

1
12
22222
2
12 122
d
=dd d
,
U
UUUUU
UU UUU
Rz
xyzzP z
cagbc
Rz
gg









(26)
It follows from Equation (19) that


 
2
12
12 12
2
12
,
lim
=d ,
lim
dd
dd
UU
a
UU UU
a
UU
aza
PRzR
RzRa
a
za







a
(27)
From (21), one can easily obtain
 

2
=,dd=
UU UUUU
UU
UU UU
acz bcag
Rz Rzz
gz bgz b

(28)
1
U

(23) and calculate the limits entering Equation (27)


2
22
2
2222
2
2
=,
lim
d=
lim d
U
U
aU
UUUU
aU
c
Ra
g
Rabc ag
a
ag
2
U
 

 

(29)
2This can be clarified by the following example. Let, e.g.,

1=1[ ]R
, where, in the general case, and=UiV
=
x
iy
. For real and
, this transformation is real
and has a singularity at
(==0)Vy
0=
x
U0V. If , then an imaginary part,
located in this singularity (at small V), arises: Im
 
2
1
2
=
R
By comparing (27) and (26), we eventually have
xVxU
V


xU

. In our case, the trans-
formation
1
R
 is more complicated, but its imaginary part arises
like in this simple example.

2
12 12
=
lim UU
a,d
x
yaxa

xx (30)
Copyright © 2011 SciRes. AM
970 G. G. KOZLOV
4. Calculating the <x1y2> Contribution
Let us change variables in Equation (6):
21 1
,=UUUU
 . Then, with allowance for the
remark given after Equation (11) and Equation (30), we
may write the following expression for the quantity :
D


2
12
20,
1,2
2,
,
=
=edd
lim
=e ,dd
lim
N
it
Vt
it UU
at
Dt
ixyU
iaxaxx


 


d
U
(31)
Note that, as was shown in [7], the participation
function
WU can be obtained from (31) by omitting
in it the integration over
U
 
2,
,
=e ,d
lim it UU
at
i
WUa xaxx
d
 
(32)
Following methodology of [7], we can represent the
distribution function of the site energies

P
in the
form
  

1
1
=,0,d
d0
nn
Pppxpxx
pxx xMM
 




1,
0
(33)
In the case of an ordered chain, and, as a
result, . The perturbative approach to
solution of the equation similar to Equation (19),
proposed in [7], represents expantion in powers of
0

==DWU
,
with the first nonzero correction being of the order of
. It was also shown in [7] that, for calculation of the
quantities and , it suffices to have only the
singular in
2
DWU

part of solution of the equation for joint
statistics of the Green’s functions (Equation (19)), with
the singularity being of the pole type. Therefore, if we
denote this singular part by the symbol “sing,” then we
may present it in the form
 

23
1212
=
UU
singx xFxxO

(34)
Using Equation (30), we obtain the following formula
for the sought function
WU:
 

22 3
=,d
lim U
a
WUa FxaxxO

 
(35)
In the next section, we describe the perturbation theory
for solving Equation (19) and derive explicit expression
for the function
12U
F
xx .
5. Perturbative Approach to Equation (19)
Assuming that the parameter is small, we can
present the sought function
12
12UU

121 2
12 =0
=,
n
UU n
n
xxQxx
(36)
Let us expand the function
 
12
12
12
12 1212
dd
,dd
UU
UU UU
RxRx
RxRxxx






under the integral sign, in the right-hand side of Equation
(19), into power series in
. Then, Equation (19) yields:

 
 
12
=0
12
12
,=0
12
12
12
=0
,
=,
!
dd
dd
n
n
n
nl n
nlU U
n
nl
UU
Qxx
MQRx Rx
n
RxRx
xx






(37)
By equating the coefficients of the same powers of
in the left- and right-hand sides of (37), we obtain the
recurrent relation for the functions :
n
Q

 
0012 012
12
12
12
12
:,
dd=0
dd
UU
UU
QxxQRx Rx
Rx Rx
xx

(38)
Since the first moment of the function

P
is zero,
we have
112
=0Qxx .
 
 
 
 
2212 212
12
12
12
12
22
201
12
112 2
12
=0
:,
dd
dd
=2 ,
dddd
UU
UU
UU
UU
QxxQRx Rx
Rx Rx
xx
2
M
QRxRx
RxxRxx








(39)
Now, we define the linear operator m
H
that acts
upon an arbitrary function

f
x as follows:
 
dR
d
mU
UU
H
fxfR x
x


(40)
Here,
U
Rx is the transformation inverse to (21). In
[12], there has been solved the spectral problem for the
operator
1
=1
2
H
fxfUxx(), and it was
shown that, at
=1m
<2U, its eigenvalues n
and
eigenfunctions
n
U
x
are given by the expression s:
 
2
2
=,
4
=,=1
4
n
nn
U
UU UU
U
n
nn
Rx
x
Lx LxGx
Rx
UiU
Ui U










(41)
x
x
in the form of
a power series in
Copyright © 2011 SciRes. AM
G. G. KOZLOV
971
where

22
4
=,=,=
2
U
UU U
U
Rx Ui UUi U
Gx RR
Rx

4
2
and Lorentzian is defined as:

U
Lx

2
2
41111
==
22
1
UUU
U
Lx ixR
x
Uxx R



 

(42)
Remind [12] that an arbitrary function

f
x may be
presented in the form of the series
 

=
=,where=
n
nUn n
nU
fxd
f
xAx A
Gx


x (43)
One can easily make sure that the functions
n
U
x
m
are also the eigenfunctions for the operator
H
(40),
and the eigenvalues are given by m-th power of
eigenvalues (41):
 
=
mn mn
UnU
H
x

m
x. The func-
tional operator
H
enters Equations (38) and (39).
Taking into account its properties described above, we
can immediately write the expression for :

012
Qxx


01212
12
=UU
QxxLxLx
(44)
To solve the functional Equation (39), let us write the
sought function in the form of expansion
over eigenfunctions of operator (40):
212
Qxx
 
2121 2
12
|| ||0
=nl
nl UU
nl
QxxC xx


(45)
By substituting this series into the left-hand side of
Equation (39) and expanding its right-hand side using
(43), we obtain, for the coefficients (45), the
following formulas: nl
C
 
2
212
2=0
12
1
=2
1
nln l
mm
nl
M
CJUJ
UU U

 



(46)
where

n
J
U
are given by
 







1
d
=d=
UU U
nn
U
Un
n
UU
LR xR x
J
U
Gx
Lz zJU
GR z

x

(47)
To expand the righ-hand side of (39), we used
Equation (44) for the function .

012
Qxx
As was said above, we are interested only in the part
of singular in 21
212
Qxx
=UU
. To extract this
part, we have to retain in Equation (45) only the terms
with [7], because only for these terms the
denominator
=nl

1
1mm
nl
UU

=0
2
, in Equation (46),
vanishes at
. The calculations identical to those
performed in [7] lead to the following expression for the
function
12U
F
xx entering Equation (34):
 
 
22
2
2=0
0
12
=4
Un
n
nn
UU
Fx U JU
xx
n

2
12 4
iM
xm
(48)
Here are the explicit expressions for integrals (47):

=1,
n
nU
G R




0
11
==
,
n
UU
JU
JUR JUn



>0
(49)
These expressions are obtained by integrating (47)
over residues. When calculating the derivatives entering
(48), the quantity
can be considered so small that the
arrangement of the poles of the integrants with respect to
the real axis does not depend on
.
Using properties of the function
U
Gx
(41) pre-
sented in [12], we can obtain the following re lationship



11
1
nm
z
11
==
n
mn U
n
nUU
UU
Rz
GzR z
GR




(50)
which shows that
,0 =0
n
JU at and that, in
the general case, the expansion of 0n

11
R

UU
GR U
in
powers of
starts from the first power of
and may
be written in the form:
2 3
1
=,=
UUU U
GRJUK TO
11U
R
 
 
(51)
By substituting this expression into (49), we see that,
in Equation (48), only the terms containing
1
J
U
survive, for which the second derivative of their module
squared is nonzero at =0
. Thus, Equation (48) for the
function
12U
F
xx

may be transformed to the form:
 
2
2
2
12
11 11
12 12
=4
2
()
UU
UU UU
iM
FxxUK
m
x
xxx
 


(52)
The direct algebraic calculation with the use of
explicit expressions (41) and (21) for the function
U
Gx and transformation , respectively,
shows that

1
U
Rx

2
12
22
sin
=,=4
44
mmm
U
UUUU
Rm
KRRK
UU





(53)
Finally, using the expressions for the moments and
limiting values of the function

n
U
x
, presented in [7]
 
2
22 4
d=4,=
lim
22
nn
UU
a
in U
xxxUa a
n


(54)
Copyright © 2011 SciRes. AM
972 G. G. KOZLOV
with the aid of Equation (35), we obtain, for the
participation function and the quantity ,
formula (4).

WU D
6. Delocalization Points
As was mentioned in Introduction, the appearance of the
delocalization points (5) predicted in Equation (4) looks
curious, taking into account the known assertion that all
states in a 1D random system are localized [5]. The fact
that the character of the states of Hamiltonian (1)
(localized/delocalized), calculated in the second per-
turbation order, appears to be, in these energy points, the
same as in the totally ordered system, indicates that, for
studying the states with energies (5), one has to analyze
statistics of Green’s functions to within the terms of the
order higher than . Complete analysis of this kind is
rather cumbersome and lies outside the scope of this
paper. Still, with the aid of reasoning presented below, it
is exactly in the vicinity of the points n (5) where the
behavior of the participation function can be
predicted.
2
E

WU
Assume that the odd moments of the site energy
distribution function

P
(33) are zeros ()
and consider the fourth-order correction 4 in ex-
pansion (36) of the joint statistics of Green’s functions.
The functional equation for this correction is derived in
the same way as Equation (39) for and has the
form:
21
=0
n
M
Q
2
Q
 
 
 

  
 
4412 412
12
12
12
12
21
21
21 2
212 1
4
2
24
01
412
2=0
12
12
12
=0
:,
dd
dd
d
=,
2d
d,
d4!
dd
dd
UU
UU
U
UU
U
UU
UU
QxxQRx Rx
Rx Rx
xx
Rx
MQRxRxx
Rx MQRxRx
x
RxRx
xx











2

(55)
The right-hand side of this equation is a sum of
contributions, so that, if dependence of any of them on
energy arguments 1,2
U has a peculiarity, then such a
peculiarity will be displayed by the function
412
Qxx.
We will now show that some terms of series (45), for
function
212
Qxx
U, diverge at the values of their energy
arguments equal to (5).
1,2 n
Using Equations (46), (49), and (51), one can make
sure that, among the coefficients (46), nonzero are
E
nl
C
only 111, 11,11,1
=,=CC CC

0, 2
,C
, ,
1,01,00,10, 1
=,=CCCC


2,0
C
. Let us present expression, e.g., for .
Equation (46) yields:
1,0
C

21
1,0 11
=1
U
m
MT
CU
(56)
where 1 is determined by Equation (51). It can be
easily shown that (41) and
U
T

1=1
m
n
E
0
En
T
(51).
It is this fact that gives rise to the above divergence of
2
x
41
Qx at , which, in turn, leads to
divergence of the correction to the participation
function
1=n
UE
4
WU. The terms of expansion (45)
proportional to and also diverge at
(5), with no compensation for the div ergence.
21
Qx
0,
C
2
x
11,0 0,1
,CC
1,2 =Un
E
Spectral behavior of other terms of the expansion at
1,2 (5) is essentially different. For instance, the
coefficient defined by the re l a t i onship
=n
UE
1,1
C
 
12
1,12 1112
=1
UU
mm
KK
CM UU

(57)
tends to zero at 1,2 (5) because the product
12
=n
UE
UU
K
K tends to zero faster than

1112
1mm
UU

.
Appearance of divergence is also possible in the points
other than Equation (5), where the corresponding
peculiarities of the function
n
E
WU are, however, not so
noticeable, because the function calculated in
the second order in

WU
is nonzero.
The divergence described above should look as a
strong deviation of the participation function
WU, at
=n
UE
n
E
(5), from that described by Equation (4). The
size of the spectral regions, in the vicinity of the points
(5), where the correction begins to exceed
the correction
4
2
(4), calculated in this paper, will
decrease with decreasing
. Since the correction 2
of the function
WU turns to zero at (5), the
diverging corrections of higher orders should give rise to
narrow (at small
=n
UE
) peaks in these spectral points.
Quantitative description of the amplitude and shape of
these peaks lies outside the scope of this paper.
7. Numerical Experiment
All the results obtained above refer to the case of an
infinite 1D chain, and, therefore, in computer testing of
these results, one has to employ matrices (1) of so large
dimension that quantities (2 ) and (3) calculated for them,
cease to depend on it. When choosing the matrix
dimension for the numerical experiment, one may be
governed by visual sense of Equation (4) which we
consider below.
Copyright © 2011 SciRes. AM
G. G. KOZLOV
973
The qualitative picture of the excitation dynamics
studied in this pap er is that this excitation, being initially
located at the edge site N, remains localized near this site
at . In this case, the appropriate wave function
proves to be essentially nonzero only at some finite
number of sites in the vicinity of the edge site. The
t
L
quantity

2
=N
Dt calculated in this paper
and the normalization condition of the wave function
provide opportunity to evaluate the number as
L1D
and to introduce the following natural definition for the
mean localization radius L:
2
2
12
=m
LDM
(58)
It is evident that, in the numerical experiments, the
matrix dimension should substantially exceed
NL .
The participation function calculated in this
paper, allows one to judge about spectral dependence of
the localization radius. For, instance, if the function
, for some energy U, is by a factor of k
smaller than its mean value (equal to

WU

WU
0
DV , where
0 is the width of the matrix (1) spectrum), then, we
may say that the localization radius of the states
with the energy U is by a factor of larger than
4V

lUk
L. It means that the function , obtained
numerically, may strongly deviate from Equation (4)
near the points of delocalization (5), because
the localization radius of the states with the energies
close to n (5) substantially exceeds the mean one and
may become greater than the matrix dimension used
in the numerical experiment.
WU
=n
E
U
EN
The above reasoning shows the reasons why the
smallest matrix dimension th at can be used for the testing
decreases with increasing degree of disorder . On the
other hand, evidently, it is possible to neglect the terms
of the order higher than second, in Equation (4), only
when , which is the case only at small
1D
. For
this reason, the degree of disorder and the matrix
dimension , in the numerical testing, should, at least,
meet the following cond ition:
N
2
2
11
2
M
Nm

(59)
Our numerous computer experiments with matrices (1)
of different dimension and with the different degree
of diagonal disorder support the above qualitative
conclusions. To obtain statistics of site energies (33), in
the numerical testing, we used the function
N
2
=0.50.5,=1px xxM 12
Figure 1 shows spectral dependences of the function
obtained numerically using Equation (3)

WUdU
(a)
(b)
(c)
Figure 1. Distribution of the degree of localization of the
states in 1D disordered chains with a complex structural
unit. Noisy plots are obtained by computer simulation, and
smooth curves, using Equation (4). Dimension of the
random matrices, in all cases, is 4000. The values of other
parameters are: (a) , (b) , (c) m=0.5, =2m=0.5, =3
m
= 0.25,=3 . In all case s, =dU 150.
(noisy plots) and the results of calculations using
Equation (4) (smooth curves). The calculations were
performed at =0.5
, =150dU , (Figure
1(a)) and (Figure 1(b)). The matrix dimension
was , and averaging over 100 realization was
made. One can see from Figures 1(a) and (b) that the
numerical and theoretical dependences well agree with
each other, with the points of delocalization (5) distinctly
seen in both figures. The narrow peaks of the numerical
plot in these points correspond to qualitative predictions
made in the previous section.
=2m
=3m
4000=N
Figure 1(c) shows the results of similar calculations
for the disordered chain with. In this case,
=3m
=1152L, and at , condition (59) is satis-
fied relatively weak. One can see from Figure 1(c) that,
in the spectral regions near the points of delocalization,
where the localization radius of the states exceeds
= 4000N
L
(58) and becomes comparable with , the discrepancy
between the results of numerical experiments and theo-
N
Copyright © 2011 SciRes. AM
G. G. KOZLOV
Copyright © 2011 SciRes. AM
974
retical curve is more noticeable than far away from these
points. The calculations were performed with no fitting.
8. Conclusions
In this paper, we have analyzed spectral dependence of
the degree of localization of the states in a 1D disordered
chain with a complex structural unit in the form of a
segment consisted of two-level atoms. It is shown
that distribution of the degree of localization, for such a
model, qualitatively differs fro m that for the chain with a
simple structural unit and is essentially inhomogeneous.
This distribution is characterized by appearance of
spectral regions in which the states are, to a considerable
extent, delocalized, with exception of central points of
these regions, where the degree of localization exhibits
sharp peaks. The calculations are performed using the
developed perturbative approach for the joint statistics of
the advanced and retarded Green’s functions.
m
The work was supported by the Analytical Depart-
mental Targeted Program “Development of Scientific
Potential of Higher School” of 2009-2010 (project no.
2.1.1/1792).
9. References
[1] V. L. Berezinsky, “Kinetics of a Quantum Particle in a
One-Dimensional Random Potential,” Journal of Ex-
perimental and Theoretical Physics, Vol. 65, 1973, pp.
1251-1266.
[2] A. A. Gogolin, V. I. Mel’nikov, E. I. Rashba, “Conduc-
tivity in a Disordered One-Dimensional System Induced
by Electron-Phonon Interaction,” Journal of Experimen-
tal and Theoretical Physics, Vol. 69, 1975, pp. 327-349.
[3] A. A. Gogolin, V. I. Mel’nikov, E. I. Rashba, “Effect of
Dispersionless Phonons on the Kinetics of Electrons in a
One-Dimensional Conductors,” Journal of Experimental
and Theoretical Physics, Vol. 72, 1977, pp. 629-645.
[4] Y. A. Bychkov, “Absence of Sound in One-Dimensional
Disordered Crystals,” Journal of Experimental and Theo-
retical Physics, Vol. 67, 1974, pp. 2323-2325.
[5] I. M. Lifshits, S. A. Gredeskul and L. A. Pastur, “Intro-
duction to the Theory of Disordered Systems,” English
Translation, Wiley, New York, 1988.
[6] P. W. Anderson, “Absence of Diffusion in Certain Ran-
dom Lattices,” Physical Review, Vol. 109, No. 5, 1958,
pp. 1492-1505. doi:10.1103/PhysRev.109.1492
[7] G. G. Kozlov, “Computation of Localisation Degree in
the Sence of the Anderson Criterion for a One-Dimen-
sional Diagonally Disordered System,” Theoretical and
Mathematical Physics, Vol. 162, No. 2, 2010, pp. 238-
253. doi:10.1007/s11232-010-0019-1
[8] D. H. Danlap, H.-L. Wu and P. W. Phillips, “Absence of
Localization in a Random-Dimer Model,” Physical Re-
view Letters, Vol. 65, No. 1, 1990, pp. 88-91.
doi:10.1103/PhysRevLett.65.88
[9] V. A. Malyshev, A. Rodriguez and F. Dominguez-Adame,
“Linear Optical Properties of One-Dimensional Frenkel
Exciton Systems with Intersite Energy Correlations,”
Physical Review B, Vol. 60, No. 20, 1999, pp. 14140-
14146. doi:10.1103/PhysRevB.60.14140
[10] F. A. B. F. de Moura and Marcelo L. Lyra, “Delocaliza-
tion in the 1D Anderson Model with Long- Range Corre-
lated Disorder,” Physical Review Letters, Vol. 81, No. 17,
1998, pp. 3735-3738. doi:10.1103/PhysRevLett.81.3735
[11] F. J. Dyson, “The Dynamics of a Disordered Linear
Chain,” Physical Review, Vol. 92, No. 6, 1953, pp. 1331-
1338. doi:10.1103/PhysRev.92.1331
[12] G. G. Kozlov, “Spectrum and Eigen Functions of the
Operator
2
1
U
HfxfUx x and Strange At-
tractor’s Density for the Mapping
1=1
nn
x
Ux
,”
arXiv: 0803. 1920. [math-ph], 2008.