Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

VOLUME 82, NUMBER 7 PHYSICAL REVIEW LETTERS 15 FEBRUARY 1999

Theory of Sound Attenuation in Glasses: The Role of Thermal Vibrations


Jaroslav Fabian1,2 and Philip B. Allen 2
1
Department of Physics, University of Maryland at College Park, College Park, Maryland 20742-4111
2
Department of Physics and Astronomy, State University of New York at Stony Brook, Stony Brook, New York 11794-3800
(Received 13 October 1998)
Sound attenuation and internal friction coefficients are calculated for a realistic model of amorphous
silicon. It is found that, contrary to previous views, thermal vibrations can induce sound attenuation at
ultrasonic and hypersonic frequencies that is of the same order or even larger than in crystals. The rea-
son is the internal strain induced anomalously large Grüneisen parameters of the low-frequency resonant
modes. [S0031-9007(99)08429-X]

PACS numbers: 62.65. + k, 62.80. + f, 63.50. + x

Sound attenuation in glasses is poorly understood. This til now, however, there has been no numerical study to
is because many competing factors lead to sound-wave test this argument.
damping. Most important are thermally activated struc- In this paper we examine the role that thermal vibrations
tural relaxation, hypothetical tunneling states, topological play in the sound attenuation in glasses. We will use
defects, and thermal vibrations. Sorting out different con- the term “vibron” to refer to any quantized vibrational
tributions for a given temperature T and sound wave fre- mode in a glass [6]. Our analysis is restricted to the
quency n ­ Vy2p is a difficult task. region Vtin & 1 (the so-called Akhiezer regime [7]),
Experiments show the following features: (i) At tem- where tin is the inelastic lifetime or thermal equilibration
peratures T & 10 K and ultrasonic frequencies (10 MHz time of a thermal vibron. We show that the unusually
to 1 GHz) the sound attenuation coefficient GsT d exhibits strong coupling (measured by Grüneisen parameters g)
a small, frequency-dependent peak [1]. (ii) At higher tem- between sound waves and the low-frequency resonant
peratures, between 10 and 200 K, another peak in GsT d modes explains the features (iii) through (v). As for the
develops whose center increases only moderately when n interpretation of (ii), our calculation shows that confusion
increases. The peak broadens at hypersonic frequencies arises because there actually are two different peaks.
[2] and is not seen above 100 GHz [3,4]. As a func- One is caused by relaxational processes (not addressed
tion of frequency, Gsnd , n at the peak temperatures [2]. here) and dominates below 1 GHz and another is due to
(iii) At hypersonic frequencies, GsT d appears to be almost thermal vibrations and dominates at the lowest hypersonic
independent of (or slightly increasing with) T above the frequencies. A double peak structure should be expected
peak (ii) to at least 300 K [3,4]. (iv) Room temperature at intermediate frequencies. There is some indication for
Gsnd , n 2 from at least 200 MHz [2]; this dependence such structure in measurements on vitreous silica [2]. Our
continues for up to about 300 GHz [3,4] and seems valid calculation is also a prediction: The existing measurements
for any temperature above the peak (ii). Finally, (v) the on amorphous Si [8] report G at too low frequencies
attenuation coefficients for longitudinal (GL ) and trans- (300 MHz) to see contributions of thermal vibrations. But
verse (GT ) waves are similar [2]. even at higher frequencies (say, 30 GHz) one may expect
While the low-temperature behavior (i) of G is un- traces of thermally activated peaks due to various defects.
derstood based on the interaction of sound waves with Recently discovered amorphous Si with 1 at. % H [9] in
tunneling states [1], features (ii) through (v) lack con- which tunneling (and perhaps also relaxational) processes
sistent theoretical justification. The higher-temperature are inhibited would be excellent to test our results.
peak (ii) shows many attributes of a thermally activated In the Akhiezer regime a sound wave passing through a
relaxational process [5], but a calculation shows that to solid can be attenuated by two processes [10]. First, if the
fit experiment, different sets of relaxational processes are wave is longitudinal, periodic contractions and dilations
needed for different n [2]. Also the plateau region (iii) in the solid induce a temperature wave via thermal expan-
is difficult to explain by a thermally activated relaxation sion. Energy is dissipated by heat conduction between
process since numerical fits require unphysically large at- regions of different temperatures. Second, dissipation
tempt frequencies [4]. Further, thermal relaxation pro- occurs as the gas of vibrons tries to reach an equilibrium
cesses give attenuation that increases more slowly than characterized by a local (sound-wave-induced) strain.
quadratic with increasing n, contradicting (iv). Thermal This is the internal friction mechanism. To establish the
vibrations have been overlooked as a sound-wave damp- relative importance of the two processes, consider order-
ing factor on grounds that vibrational modes would need of-magnitude formulas Gh ø sV 2 yry 3 d skT a 2 r 2 y 2 yC 2 d
unreasonably large Grüneisen parameters (g ø 200 for and Gi ø sV 2 yry 3 d sCT tin g 2 d for the heat conductivity
vitreous silica [2]) to account for the measured G. Un- and internal friction processes, respectively [10]. Here

1478 0031-9007y99y82(7)y1478(4)$15.00 © 1999 The American Physical Society


VOLUME 82, NUMBER 7 PHYSICAL REVIEW LETTERS 15 FEBRUARY 1999

r is density, C is specific heat per unit volume, y is as much as propagating ones to the momentum current.
sound velocity, k is thermal conductivity, and a is the One consequence is that the concept of “minimum kinetic
coefficient of thermal expansion. The ratio Gh yGi ø coefficient,” as introduced for electrical [14] or heat [15]
ska 2 r 2 y 2 dysC 3 tin g 2 d becomes more intuitive when conductivity of disordered systems, is not realized for
putting a ø CgyB (B ø ry 2 is the bulk modulus) internal friction. We generalized [16] DeVault’s theory
and k ø CD, where D is diffusivity. Then Gh yGi ø to include internal strain, the atomic rearrangements in
Dysy 2 tin d. The factor y 2 tin measures the ability of a strained solid. We found that internal strain affects
vibrons to absorb energy from a sound wave of velocity internal friction only by modifying g j , as in the case
y. The difference between a glass and a crystal lies in the of thermal expansion [17]: g j now reflects the change
values of D and tin . In crystals D ø y 2 tin , that is, energy between the initial mode frequency and the frequency of
is carried by phonon wave packets with group velocity y. the mode after the rescaling (scaling parameter 1 1 e)
The ratio Gh yGi is then of order unity. In glasses energy plus the rearranging of atomic positions (to achieve a new
is transferred by diffusion (spreading rather than ballistic equilibrium at strain e). Internal strain is very important
propagation of wave packets [11]) and D is not related for thermal expansion of glasses [17]; we will show that it
to tin [12]. One of the reasons the contribution to Gi of is important for h (and G) as well.
thermal vibrons was previously underestimated is that tin We calculate h and G for the model of amorphous Si
was guessed from thermal conductivity [2]; this gave too based on the Wooten-Winer-Weaire atomic coordinates
small tin . For amorphous Si, D ø 1026 m2 ys [12], y ø [18] and Stillinger-Weber interatomic forces [19], with
8 3 103 mys, and tin ø 10212 s [6] give Gh yGi ø 0.02. 1000 atoms arranged in a cube of side 27.549 Å with
Since these are typical values, Gh can be neglected. This periodic boundary conditions. Diagonal Grüneisen pa-
j j j j
is consistent with experiment: Compared with crystals, rameters gaa y3 ; sg11 1 g22 1 g33 dy3 for this model
glasses have smaller k and yet G can be larger [2]. [20] were given in Ref. [17]; transverse gab are calcu-
Internal friction leads to sound-wave energy attenuation lated here. Vibrational lifetimes tj are extracted from
[10] G ­ sV 2 yry 3 q2 dhabgd qa eb qg ed , where habgd their 216-atom version values [6] (see also Ref. [21]). The
is the internal friction tensor with cartesian coordinates model has sound velocities yL ­ 7640 mys and yT ­
a, . . . , d and q (e) is the wave vector (polarization) 3670 mys [22].
of the sound wave. Summation over repeated indices Figure 1 shows the calculated Gsnd for longitudinal and
is assumed. We will evaluate G for both longitudinal transverse sound waves in amorphous Si from 10 MHz to
(L) and transverse (T ) sound waves with wave vectors 1 THz at 300 K. The attenuation G , n 2 up to about
averaged over all directions: 100 GHz, where the condition for the applicability of ki-
V2 netic theory Vtin & 1 reaches its limit (tin ø 1 ps). Our
GL ­ shaabb 1 2habab d , (1) calculation is not valid beyond this point. In comparison,
15ryL3
the measured attenuation of longitudinal waves in vitreous
silica grows quadratically with n up to at least 400 GHz
V2
GT ­ s3habab 2 haabb d . (2)
30ryT3
The coefficients habgd are the real part of a complex 6
tensor h̄abgd which can be obtained by solving a kinetic 5 T=300 K
equation in relaxation time approximation [7], 4 a-SiO 2 (L) Vacher et al.
j j j j
X gab ggd 2 sḡab ggd
1 gab ḡgd dy2 3 a-SiO 2 (L) Zhu et al.
h̄abgd ­ Tcj tj .
2
LOG(Γ)

j 1 2 iVtj
(3) 1
The summation is over all vibrational modes j; cj and 0
a-Si (L)
tj denote mode specific heat and relaxation time. The -1 a-Si (T)
j
Grüneisen tensor 2gab is the relative shift of mode -2 a-Si (L) no IS
frequency vj per unit strain eab ; ḡ is the mode average c-Si (L) Mason
of g j weighted with cj ys1 2 iVtj d. The applicability -3 c-Si (T) Mason
of kinetic theory to the problem of internal friction was -4
justified by DeVault and co-workers [13] who obtained h 7 8 9 10 11 12
from a microscopic theory as an autocorrelation function LOG(ν)
of the momentum current density operator. Remarkably,
the microscopic theory shows that the momentum current FIG. 1. Log-log plot of the sound attenuation G scm21 d at 300
K, as a function of sound-wave frequency n (Hz). Calculated G
in a solid is not monopolized by ballistically propagating are represented by lines (IS: internal strain), experimental data
vibrational modes as in the case of the energy current. by symbols. L (T) stands for longitudinal (transverse) sound
Nonpropagating (even localized) modes can contribute waves.

1479
VOLUME 82, NUMBER 7 PHYSICAL REVIEW LETTERS 15 FEBRUARY 1999

[4], suggesting that tin in vitreous silica is several times In Fig. 2 we plot GsT d for different n. A remarkable
smaller than in amorphous Si. This is not surprising since feature is a peak at about 20 K at 1 MHz and below. As n
Si is remarkably harmonic: Room temperature heat con- increases, the peak shifts towards higher T and vanishes
ductivity of crystalline Si is larger by an order of magnitude aboveP4–5 GHz. Two factors cause the peak. (a) The
than that of quartz [23], and a similar relation may hold for sum j cj sg j d2 saturates at much lower temperatures
the corresponding tin of the glassy phases. (about 50 K) than the model Debye temperature TD ø
More surprising is the comparison with crystalline Si. 450 K [22]. This is because the relevant j are resonant
Figure 1 shows that GL is similar for the amorphous modes with small frequencies. (b) For low-frequency
and crystalline cases (measured G for vitreous silica is modes, T tj (after increasing linearly) develops a peak,
several times larger than for quartz [2]). One would before going constant [much like GsT Pd itself]. As the
naively expect the sound attenuation in a glass to be temperature dependence of G follows j cj sg j d2 T tj , the
much smaller than in the corresponding crystal since, peak appears. At large n the peak vanishes because of the
owing to a distribution of bond lengths and bond angles, factor 1ys1 1 V 2 t 2 d in Eq. (3). At T above 100 K, GsT d
anharmonicity of the glass is higher (and tin smaller). is nearly constant, as observed in experiment as a plateau
The same interatomic potential, for example, yields tin for (iii). This again follows from (a) and (b).
high-frequency phonons in crystalline Si at 300 K about We are not aware of any experiment with which we
5 times larger than in amorphous Si [6]. The reason why could compare our calculations. The measurement of GsT d
G in glasses can be of the same order or even higher than of sputtered amorphous Si films reported in Ref. [8], for
in crystals is the internal-strain-induced anomalously large example, was performed at 300 MHz. This is too low
Grüneisen parameters of the resonant modes [17] (see to see any contributions from thermal vibrations. The
also Fig. 3 below). (Resonant modes are low-frequency whole temperature spectrum is dominated by a single peak
extended modes whose amplitudes are unusually large at a of the type (ii), except at very low temperatures. This
small, typically undercoordinated region [17,24].) Atomic peak is expected to increase linearly with n, until thermal
rearrangements caused by internal strain are largest in vibrations become relevant (roughly at 10 GHz), causing
the same regions of undercoordination where the resonant a plateau (iv) that increases as n 2 at higher frequencies.
modes have largest amplitude [17]. This leads to high Even at smaller frequencies one may see some vibrational
sensitivity (measured by g) of the frequencies of these contribution to GsT d at large enough T , since the thermally
modes to strain. If the internal strain is neglected, the activated peak decreases as 1yT at large T .
sound attenuation is an order of magnitude smaller, as Anomalous low T thermal expansion already suggested
seen in Fig. 1. (Since the resonant modes have low [25] very large g values for low v modes. Our large g
frequencies, their tj is longer than an average tin ; this values [17] agree nicely with trends in asT d. Like thermal
adds even more weight to these modes.) Fewer than 1% expansion, G should be strongly sample and model depen-
of the modes are capable of increasing G by a decade. We dent. There is evidence [26] that our highly homogeneous
believe the measured G for vitreous silica is also caused model of amorphous Si becomes free of resonant modes
by the strong coupling of sound waves and resonant when the number of atoms grows to infinity. That means
modes. Vitreous silica is a much more open structure than
amorphous Si so the number of resonant modes should be 20
higher, bringing G above the crystalline value.
1 GHz (L)
Another interesting feature in Fig. 1 is the relative
1 GHz (T)
attenuation strength for longitudinal and transverse sound 1 MHz (L)
waves. While our model of amorphous Si gives GL yGT ø 15
10 GHz (L)
1y3 at 300 K, the measured ratio for crystalline Si is
reversed: GL yGT ø 3 [23]. This again shows how differ-
Γ (cm )
-1

ently is sound attenuated in glasses and in crystals. The 10


ratio GL yGT can be written as syT yyL d3 sgL2 ygT2 d, where
gL and gT are effective Grüneisen parameters. A crude
way to estimate gL2 and gT2 , suggested by Eqs. (1) and (2), 5
j j j j
is to take mode averages of sgaa gbb 1 2gab gab dy15 2
j j j j 4
and s3gab gab 2 gaa gbb dy30. Our model gives gL2 ø 8
2
3 and gT ø 1. The ratio GL yGT is then about 1:3, in 0
accord with the full calculation. Assuming the same ra- 0 50 100 150 200 250 300
tio gL2 ygT2 ø 3 for vitreous silica (yL ­ 5800 mys and T (K)
yT ­ 3800 mys), transverse and longitudinal waves are
FIG. 2. Calculated sound attenuation GsTd for amorphous
attenuated about equally. This is observed in experiment Si at different frequencies. The thin dashed lines are for
[2]. The explanation of the measured GL yGT in crys- longitudinal waves with the labeled n in GHz. Plotted are
talline Si can be found in Ref. [23]. rescaled values Gyn 2 for n measured in GHz.

1480
VOLUME 82, NUMBER 7 PHYSICAL REVIEW LETTERS 15 FEBRUARY 1999

4 [1] S. Hunklinger and W. Arnold, in Physical Acoustics,


edited by W. P. Mason and R. N. Thurston (Academic,
3 New York, 1976), p. 155.
[2] R. Vacher, J. Pelous, F. Plicque, and A. Zarembowitch,
2 J. Non-Cryst. Solids 45, 397 (1981).
[3] C. J. Morath and H. J. Maris, Phys. Rev. B 54, 203
γ12

1 (1996).
[4] T. C. Zhu, H. J. Maris, and J. Tauc, Phys. Rev. B 44, 4281
0 (1991); H. N. Lin, R. J. Stoner, H. J. Maris, and J. Tauc,
J. Appl. Phys. 69, 3816 (1991).
-1 [5] D. Tielbürger, R. Merz, R. Ehrenfels, and S. Hunklinger,
-10 Phys. Rev. B 45, 2750 (1992).
-20 [6] J. Fabian and P. B. Allen, Phys. Rev. Lett. 77, 3839
-30 (1996).
[7] H. J. Maris, in Physical Acoustics, edited by W. P. Ma-
-40 son and R. N. Thurston (Academic, New York, 1971),
0 10 20 30 40 50 60 70 80 90
Vol. VIII.
MODE FREQUENCY (meV) [8] M. Von Haumeder, U. Strom, and S. Hunklinger, Phys.
Rev. Lett. 44, 84 (1980).
FIG. 3. Calculated transverse Grüneisen parameters g12 for
amorphous Si as a function of vibron frequency. Above the [9] X. Liu, B. E. White, R. O. Pohl, E. Iwanizcko, K. M.
vertical line (ø71 meV) the modes are localized. Jones, A. H. Mahan, B. N. Nelson, R. S. Crandall, and
S. Veprek, Phys. Rev. Lett. 78, 4418 (1997).
[10] V. L. Gurevich, Transport in Phonon Systems (North-
an infinite model would predict G about a decade smaller Holland, Amsterdam, 1986).
than calculated here. Amorphous silicon, however, can be [11] P. B. Allen and J. Kelner, Am. J. Phys. 66, 497
prepared only in thin films, where voids and other inho- (1998).
mogeneities are unavoidable. Voids loosen the strict re- [12] P. B. Allen and J. L. Feldman, Phys. Rev. B 48, 12 581
quirements of a tetrahedral random network (for example, (1993); J. L. Feldman, M. D. Kluge, P. B. Allen, and
by introducing free boundary conditions). Then, as in our F. Wooten, Phys. Rev. B 48, 12 589 (1993).
finite models, regions of undercoordinated atoms will al- [13] G. P. DeVault and J. A. McLennan, Phys. Rev. 138, A856
(1965); G. P. DeVault, Phys. Rev. 149, 624 (1966); 155,
low the formation of resonant modes. While this issue for
875 (1967); G. P. DeVault and R. J. Hardy, Phys. Rev.
amorphous silicon will be ultimately settled by experiment, 155, 869 (1967).
our calculation combined with the existing data on vitre- [14] N. F. Mott, Philos. Mag. 22, 7 (1970).
ous silica strongly suggests the reality of resonant modes. [15] C. Kittel, Phys. Rev. 75, 972 (1949).
Our final note concerns the mode dependence of trans- [16] J. Fabian, Ph.D. thesis, SUNY, Stony Brook, NY,
verse Grüneisen parameters such as g12 . Similar to volu- 1997.
metric gaa y3 [17], transverse g12 in Fig. 3 (g13 and g23 [17] J. Fabian and P. B. Allen, Phys. Rev. Lett. 79, 1885
look the same) is unusually large for resonant modes and (1997).
has scattered values for high-frequency localized modes. [18] F. Wooten, K. Winer, and D. Weaire, Phys. Rev. Lett. 54,
(More resonant modes have g12 negative than positive, 1392 (1985).
which suggests that resonant modes are trapped at highly [19] F. H. Stillinger and T. A. Weber, Phys. Rev. B 31, 5262
(1985).
anisotropic undercoordinated regions whose sizes change
[20] Reference [17] used a slightly different version with
under shear [17].) The 15–70 meV vibrons (diffusons a small amount of residual stress. The main results,
[6]) have g12 ø 0 (average magnitude 0.02), while the however, remain unchanged. J. L. Feldman (private
corresponding gaa y3 are of order unity [17]. Such small communication).
values (zeros in an infinite model) are characteristic for [21] S. R. Bickham and J. L. Feldman, Phys. Rev. B 57, 12 234
diffusons, which are extended modes whose polariza- (1998); Philos. Mag. B 77, 513 (1998).
tion directions (atomic displacements) point, in general, [22] J. L. Feldman, J. Q. Broughton, and F. Wooten, Phys. Rev.
at random. There remains only a short-range correla- B 43, 2152 (1991).
tion between polarization directions which determines the [23] W. P. Mason, in Physical Acoustics, edited by W. P.
diffuson’s frequency vd . If a shear, say, e12 , is applied, Mason (Academic, New York, 1965), Vol. III.
[24] R. Biswas, A. M. Bouchard, W. A. Kamikatahara, G. S.
vd changes to vd0 se12 d. Since long-range order in the
Grest, and C. M. Soukoulis, Phys. Rev. Lett. 60, 2280
diffuson polarization is absent, vd0 se12 d ø vd0 s2e12 d, and
(1988); H. R. Schober and B. Laird, Phys. Rev. B 44, 6746
g12 which is a linear coefficient in the expansion of vd0 in (1991); H. R. Schober and C. Oligschleger, Phys. Rev. B
e12 must vanish. 53, 11 469 (1996).
We thank J. L. Feldman for helpful discussions. The [25] G. K. White, Phys. Rev. Lett. 34, 204 (1975).
work was supported by NSF Grant No. DMR 9725037. [26] J. L. Feldman, P. B. Allen, and S. R. Bickham (unpub-
J. F. also acknowledges support from the U.S. ONR. lished).

1481

You might also like