delocalized vibrations in classical random chains

4
PHYSICAL REVIEW B VOLUME 48, NUMBER 9 1 SEPTEMBER 1993-I Delocalized vibrations in classical random chains Francisco Dominguez-Adame and Enrique Macia Departamento de Fisica de Materiales, Facultad de Ciencias Fisicas, Universidad Complutense, F 280-$0 Madrid, Spain Angel Sanchez Escuela Politecnica Superior, Univer8idad Carlos III de Madrid, Avda. Mediterraneo 20, E-289i8 Ieganes, Madrid, Spain (Received 17 March 1993) Normal modes of one-dimensional disordered chains with two couplings, one of them assigned at random to pairs in an otherwise perfect chain, are investigated. We diagonalize the dynamical matrix to And the normal modes and to study their spatial extent. Multifractal analysis is used to discern clearly the localized or delocalized character of vibrations. In constrast to the general viewpoint that all normal modes in one dimensional random chains are localized, we And a set of extended modes close to a critical frequency, whose number increases with the system size and becomes independent of the defect concentration. In his famous paper on vibrations of glasslike disor- dered chains, ~ Dean stated that when disorder (in any form) exists in a system the lattice modes are localized. Since Dean's paper was published, it has been claimed that unless the chain is ordered, or unless w = 0, all vibra- tional modes are localized in one dimension. Later on, in analogy to previous works concerning related vibra- tional problems, some authors conjectured that electrons in one-dimensional disordered lattices are also localized. That is, localization of all eigenstates by disorder in one- dimensional systems is viewed as an exact statement. In a series of recent papers, however, Wu, and co-workers have proposed some discrete (tight-binding) models that exhibit metal-insulator transitions in spite of their ran- domness. These authors have shown that when defects containing a plane of symmetry" are introduced at ran- dom in an otherwise ordered chain, v K states (K be- ing the total number of states) remain unscattered by the disorder and consequently are extended. In addi- tion, Sanchez and Dominguez-Adame have presented evidence that a large number of states whose localization length is greater than the system size arises in contin- uous (Kronig-Penney) models, in which b functions are regularly spaced and their strength takes two values, one of them in pairs at random (dimer defects). As a con- sequence, in such random systems electronic transport can take place almost ballistically. In view of these re- sults, it is a natural question to ask whether classical vibrational modes in atl random lattices are actually lo- calized. In this paper we attempt to answer this question showing a particular random system that presents a set of delocalized vibrations. We feel that this is a novel re- sult because, whereas electron dynamics is quantum me- chanical and tunneling allows for transitions classically forbidden (roughly speaking, tunneling favors delocaliza- tion), we are dealing with a purely classical problem it is well known that energy levels of the vibrating system can be derived by considering the corresponding classical problem. Quantum features of phonons appear in their statistics, not in their dynamics. n= 1, ... , N, (1) where U is the displacement of the nth atom from its equilibrium position, m is the corresponding mass, and K denotes the strength of the harmonic coupling be- tween atoms. We build our model in the following way: we take all the masses to be the same m = m, and we allow only two values for K, K, and K', with the additional constraint that K' appears only in pairs. As- suming a time harmonic dependence U oc exp(iwt) in Eq. (1), the stationary equation of motion can be cast in the following matrix form: /U. )(K+K( AKK„g)(U ++ i i K K = P„(A) (2) with A = m(u2/K. Consider now a single defect in which three atoms, placed at sites I 1, l, and l+ 1, are coupled by strengths K' between them and by strengths K to the surrounding lattice. The transfer matrix across such a defect is sim- ply given by the matrix product Pj+q(A)P~(A)Pj q(A). It is a matter of simple algebra to demonstrate that at the particular value A = 2K'/K this matrix product equals, appart from a constant phase change without physical relevance, the transfer matrix at any site of the perfect lattice. The meaning of this result is easy to understand: There exists a special frequency w, = /2K'/m for which the reflection coefBcient at the defect vanishes. Since ~ must be below the highest frequency of vibrations in the A standard way to study the acoustic properties of a lattice is to consider a nearest-neighbor harmonic chain. The equation of motion for the N atoms reads d2U„ m„" = K„U„+~+K„jU„g (K„+K„y)U„, dt2 0163-1829/93/48(9)/6054(4)/$06. 00 6054 QC1993 The American Physical Society

Upload: angel

Post on 23-Dec-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Delocalized vibrations in classical random chains

PHYSICAL REVIEW B VOLUME 48, NUMBER 9 1 SEPTEMBER 1993-I

Delocalized vibrations in classical random chains

Francisco Dominguez-Adame and Enrique MaciaDepartamento de Fisica de Materiales, Facultad de Ciencias Fisicas, Universidad Complutense, F 280-$0 Madrid, Spain

Angel SanchezEscuela Politecnica Superior, Univer8idad Carlos III de Madrid, Avda. Mediterraneo 20,

E-289i8 Ieganes, Madrid, Spain(Received 17 March 1993)

Normal modes of one-dimensional disordered chains with two couplings, one of them assignedat random to pairs in an otherwise perfect chain, are investigated. We diagonalize the dynamicalmatrix to And the normal modes and to study their spatial extent. Multifractal analysis is usedto discern clearly the localized or delocalized character of vibrations. In constrast to the generalviewpoint that all normal modes in one dimensional random chains are localized, we And a setof extended modes close to a critical frequency, whose number increases with the system size andbecomes independent of the defect concentration.

In his famous paper on vibrations of glasslike disor-dered chains, ~ Dean stated that when disorder (in anyform) exists in a system the lattice modes are localized.Since Dean's paper was published, it has been claimedthat unless the chain is ordered, or unless w = 0, all vibra-tional modes are localized in one dimension. Later on,in analogy to previous works concerning related vibra-tional problems, some authors conjectured that electronsin one-dimensional disordered lattices are also localized.That is, localization of all eigenstates by disorder in one-dimensional systems is viewed as an exact statement. Ina series of recent papers, however, Wu, and co-workershave proposed some discrete (tight-binding) models thatexhibit metal-insulator transitions in spite of their ran-domness. These authors have shown that when defectscontaining a plane of symmetry" are introduced at ran-dom in an otherwise ordered chain, v K states (K be-ing the total number of states) remain unscattered bythe disorder and consequently are extended. In addi-tion, Sanchez and Dominguez-Adame have presentedevidence that a large number of states whose localizationlength is greater than the system size arises in contin-uous (Kronig-Penney) models, in which b functions areregularly spaced and their strength takes two values, oneof them in pairs at random (dimer defects). As a con-sequence, in such random systems electronic transportcan take place almost ballistically. In view of these re-sults, it is a natural question to ask whether classicalvibrational modes in atl random lattices are actually lo-calized. In this paper we attempt to answer this questionshowing a particular random system that presents a setof delocalized vibrations. We feel that this is a novel re-sult because, whereas electron dynamics is quantum me-chanical and tunneling allows for transitions classicallyforbidden (roughly speaking, tunneling favors delocaliza-tion), we are dealing with a purely classical problem itis well known that energy levels of the vibrating systemcan be derived by considering the corresponding classicalproblem. Quantum features of phonons appear in theirstatistics, not in their dynamics.

n= 1, . . . , N, (1)

where U is the displacement of the nth atom from itsequilibrium position, m is the corresponding mass, andK denotes the strength of the harmonic coupling be-tween atoms. We build our model in the following way:we take all the masses to be the same m = m, andwe allow only two values for K, K, and K', with theadditional constraint that K' appears only in pairs. As-suming a time harmonic dependence U oc exp(iwt) inEq. (1), the stationary equation of motion can be cast inthe following matrix form:

/U. )(K+K(—AKK„g)(U++i

—i

K K

= P„(A) (2)

with A = m(u2/K.Consider now a single defect in which three atoms,

placed at sites I —1, l, and l+ 1, are coupled by strengthsK' between them and by strengths K to the surroundinglattice. The transfer matrix across such a defect is sim-ply given by the matrix product Pj+q(A)P~(A)Pj q(A). Itis a matter of simple algebra to demonstrate that at theparticular value A = 2K'/K this matrix product equals,appart from a constant phase change without physicalrelevance, the transfer matrix at any site of the perfectlattice. The meaning of this result is easy to understand:There exists a special frequency w, = /2K'/m for whichthe reflection coefBcient at the defect vanishes. Since ~must be below the highest frequency of vibrations in the

A standard way to study the acoustic properties of alattice is to consider a nearest-neighbor harmonic chain.The equation of motion for the N atoms reads

d2U„m„" = K„U„+~+K„jU„g—(K„+K„y)U„,dt2

0163-1829/93/48(9)/6054(4)/$06. 00 6054 QC1993 The American Physical Society

Page 2: Delocalized vibrations in classical random chains

48 DELOCALIZED VIBRATIONS IN CLASSICAL RANDOM CHAINS 6055

perfect lattice tu „=2/It/m, the additional conditionK' ( 2K is required. This result is related to that foundfor electrons in tight-binding models containing symmet-rical defects, where the location in the energy band ofthe unscattered states is determined by the vanishing ofthe reflection coe%cient at a single defect. " It is worthmentioning that the special frequency u, is nothing butthe &equency of the longitudinal mode of a single atomof mass m attached to two springs of constant K' withAxed ends.

Bearing in mind the above result, we now proceedto study the localization properties of vibrating systemswhen several of such defects are located at random alongthe chain. For comparison, we have also studied ordinaryrandom chains, that is, without requiring the constraintthat K appears in pairs. In addition, we have obtainedresults both realizationwise and on average. With noloss of generality we set K' = 1.5 K, so the special fre-quency for which the reflection coefficient at a single de-fect vanishes is u, = /3'/m. For brevity let us denoteA, = mu, /K = 3. Another important parameter in ourmodel is p, defined as the ratio between the number ofcouplings K' and the total number of couplings, whichwe set in the range 0.01 up to 0.8.

The localized or delocalized character of the vibrationshas been elucidated by means of multifractal analysis, amethod succesfully used in characterizing electronic wavefunctions in disordered samples (see Ref. 9 and referencestherein). The amplitude distribution of normal modescan be characterized by the scaling with the system size ofmoments associated to the measure defined in the systemby us. We use the standard definition of those moments

2.500 9.. .

~ ~

2.699 2... .

I I I I

2.999 6. . .

0I I I

1000 8000 3000 4000 5000

,i) 0.0005

0.0004-

0.0003

FIG. 1. Squared atomic displacements for three values ofthe frequency, indicated in each plot, in a system of sizeN = 5000 with p = 0.2. States become more delocalizedin approaching the critical value A = 3.

Notice that the second moment p2(N) coincides with theinverse participation ratio (IPR), as introduced, for in-

stance, in Ref. 10. The multi&actal dimension Dq is de-

fined via p~(K) N (& ) ~, for q g I. For localized vi-

brations one finds that Dq vanishes for all q, whereas Dqequals unity (the space dimension) for vibrations spread-ing uniformly. In these two extreme cases trivial multi-fractal spectra are obtained.

We first describe our studies realizationwise. To findeigenfrequencies and normal modes we directly diago-nalize the tridiagonal, symmetrical matrix arising fromEq. (2) with rigid boundaries (U0 ——

Univ+i——0).ii The

system size was as large as N = 5000. We systematicallyfound that vibrations are more delocalized in approach-ing the critical value A . Figure 1 shows the squaredatomic displacements for three difFerent frequencies closeto the critical value A for a system of N = 5000 andp = 0.2. A simple inspection of normal modes, however,does not sufFice to discuss the localized or delocalizedcharacter of vibrations. Usually the IPR works fine toclearly discern localized and extended states. Delocal-ized states are expected to present small IPR, of orderof N, while localized states have larger IPR values. Atypical situation is presented in Fig. 2(a), where the IPRfor a chain with the same parameters as in Fig. 1 is plot-

2.9 ;.3.0 3., ~

'i

1~'

1O

10 pr

. p I'„(,,"

(-!

1 0' ~ ~,

a

FIG. 2. Inverse participation ratio, obtained by direct diag-onalization of the secular matrix, for systems of size N=5000with p = 0.2 as a function of A = mu /K, where K' strengthsare set (a) at random to pairs and (b) at random. The insetshows the inverse participation ratio close to A on a linearscale.

Page 3: Delocalized vibrations in classical random chains

6056 DOMINGUEZ-ADAME, MACIA, AND SANCHEZ

ted vs A, in the range 0 & A ( 4. For comparison theIPR of an ordinary random chain with the same lengthand the same value of p is shown in Fig. 2(b). One canobserve a deep minimum of the IPR around the criti-cal value A, while this minimum is completely absentin ordinary random chains. Interestingly, the values ofthe IPR at that frequency are similar to those at A 0,which so far were believed to be the only ones that couldexhibit delocalization properties. Moreover, the inset ofFig. 2(a) shows a plateau close to A, revealing the exis-tence of a set of states with an IPR almost equal to theminimum value. This result indicates that in our modelvibrations become extended for frequencies close to w .We have estimated that the number of these states isabout the square root of the system size, although thisresult cannot be stated rigorously due to the uncertaintyin selecting these states. It is important to mention herethat the same results are obtained for larger values of theconcentration p. In particular, the value of the IPR at A

only depends on the system size but not on p. Thereforeit seems that the exact number of defects is inmaterialregarding the existence of extended vibrations.

Let us now comment on the average results. We havenumerically evaluated the positive integer moments asdefined in Eq. (3) with q = 2, . . . , 6. Atomic displace-ments were recursively computed from the transfer ma-trix equation (2), with the initial conditions Uo = 0 andU~ ——1, and the system size was as large as N = 10 .6

When computing averages, they were taken over a num-ber of realizations up to 5000 to exclude errors due topoor statistics. Convergence of moments is reached aftervery few averages for frequencies close to u . This factis easily understood in view of our previous results: Insuch a frequency range almost all states are unscatteredby defects and consequently no strong fluctuations are ex-pected. Conversely, the number of averages to obtain ac-curate results increases as one considers more distant fre-quencies. Our data indicate that 500 averages are enoughto investigate the main features of the scaling of momentswith system size. Numerical calculations point out thatmoments scale very accurately as pz(X) X (~ ) for

A in all systems we have studied, as illustratedin Fig. 3. Hence the generalized dimension Dq at thiscritical frequency is, within the numerical uncertainty,exactly one, i.e., the space dimension. This means thatvibrations spread homogeneously over the whole chain,supporting our claim that those states are completely ex-tended. Close to A, we find that p~(N) follows a powerlaw for small N but tends to a constant value for largeN, as plotted in Fig. 3 for A = 2.90. The critical sizefor which deviation from power fit occurs increases in

q:2—5- q—

q=4—10- q=5

q=6

—20-

—25-

—30 I I I I I I I

2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0log N

FIG. 3. Scaling of moments with system size for two dif-ferent frequencies: A = A, (solid line) and A = 2.90 (dashedline). The concentration of defects is p = 0.2 in both cases.The logarithms are to base 10.

approaching A, suggesting that vibrations become moreand more delocalized. This is a crucial point because itsupports that these are main features of our model irre-spective of the particular realization of the disorder.

In summary, we have studied a one-dimensional dis-ordered chain which, in contrast to the generally ac-cepted viewpoint, presents a set of delocalized states closeto a critical frequency. The number of such extendedstates roughly grows with the squared root of the systemsize. We have shown that correlation in coupling betweenatoms —foreign springs appear in pairs —leads to a cjta8-8ical resonance effect which allows for such extended vi-brations. In our studies, not reported here, we also foundthat if correlation is introduced in the masses rather thatin the couplings —foreign masses appear in pairs —thesame resonance effect occurs yielding delocalized vibra-tions. On the other side, it is clear that the main in-fluence of these states concern transport properties ofphonons. Accordingly, when computing the contributionof different modes to the thermal conductivity, as de-fined in Ref. 12 we have observed a drastic enhancementaround the critical frequency. Work currently in progressregarding these topics will be reported elsewhere.

We are indebted to Sergey A. Gredeskul for helpful dis-cussions. We also thank the Comision Interministerial deCiencia y Tecnologia of Spain for financial support underProject No. MAT90-0544 and the Universidad Carlos IIIde Madrid for computer facilities.

P. Dean, Proc. Phys. Soc. 84, 727 (1964).E. H. Lieb and D. C. Mattis, Mathematical Physics in OneDimension (Academic, New York, 1966).J. M. Ziman, Models of Disorder (Cambridge UniversityPress, London, 1979).

D. H. Dunlap, H.-L. Wu, and P. W. Phillips, Phys. Rev.Lett. 65, 88 (1990).H.-L. Wu and P. Phillips, Phys. Rev. Lett. 66, 1366 (1991).P. Phillips and H.-L. Wu, Science 252, 1805 (1991).H.-L. Wu, W. Go@, and P. Phillips, Phys. Rev. B 45, 1623

Page 4: Delocalized vibrations in classical random chains

48 DELOCALIZED VIBRATIONS IN CLASSICAL RANDOM CHAINS 6057

(1992).A. Sanchez and F. Dominguez-Adame (unpublished).M. Schreiber and H. Grussbach, Mod. Phys. Lett. B 6, 851(1992).J. Canisius and J. L. van Hemmen, J. Phys. C 18, 4873(1985).

W. H. Press, B. P. Flannery, S. A. Teukolsky, and W. T.Vetterling, Numerical Recipes in 0, 1st ed. (CambridgeUniversity Press, Cambridge, 1989).H. Matsuda and K. Ishii, Prog. Theor. Phys. Suppl. 45, 56(1970).