1: %\documentclass[preprint,preprintnumbers,nofootinbib,aps]{revtex4}
2: \documentclass[twocolumn,showpacs,preprintnumbers,nofootinbib,aps]{revtex4}
3: \usepackage{graphicx,bm,latexsym,amssymb}
4: \usepackage{amsmath,epsf,psfig}
5:
6:
7: \begin{document}
8:
9: \newcommand{\hh}{\hspace{3mm}}
10:
11: \newcommand{\sign}{\mbox{sign}}
12: \newcommand{\ssign}{\mbox{\scriptsize sign}}
13: \newcommand{\smin}{{\mbox{\scriptsize min}}}
14: \newcommand{\smax}{{\mbox{\scriptsize max}}}
15: \newcommand{\sch}{{\mbox{\scriptsize ch}}}
16: \newcommand{\tmin}{{\mbox{\tiny min}}}
17:
18: \newcommand{\la}{\left\langle}
19: \newcommand{\ra}{\right\rangle}
20:
21: \newcommand{\dla}{\la \! \la}
22: \newcommand{\dra}{\ra \! \ra}
23:
24: \newcommand{\prtl}{\partial}
25: \newcommand{\we}{\widetilde}
26:
27: \newcommand{\smfp}{{\mbox{\scriptsize mfp}}}
28: \newcommand{\smol}{{\mbox{\scriptsize mol}}}
29: \newcommand{\sph}{{\mbox{\scriptsize ph}}}
30: \newcommand{\sinhom}{{\mbox{\scriptsize inhom}}}
31: \newcommand{\sneigh}{{\mbox{\scriptsize neigh}}}
32: \newcommand{\srlxn}{{\mbox{\scriptsize rlxn}}}
33: \newcommand{\svibr}{{\mbox{\scriptsize vibr}}}
34: \newcommand{\smicro}{{\mbox{\scriptsize micro}}}
35: \newcommand{\selast}{{\mbox{\scriptsize elast}}}
36: \newcommand{\select}{{\mbox{\scriptsize elect}}}
37: \newcommand{\seq}{{\mbox{\scriptsize eq}}}
38: \newcommand{\scr}{{\mbox{\scriptsize cr}}}
39: \newcommand{\sT}{{\mbox{\scriptsize T}}}
40: \newcommand{\sTLS}{{\mbox{\scriptsize TLS}}}
41: \newcommand{\sd}{{\mbox{\scriptsize d}}}
42: \newcommand{\sext}{{\mbox{\scriptsize ext}}}
43: \newcommand{\scav}{{\mbox{\scriptsize cav}}}
44: \newcommand{\bmu}{\bm \mu}
45: \newcommand{\bE}{\bm E}
46: \newcommand{\bD}{\bm D}
47: \newcommand{\bd}{\bm d}
48: \newcommand{\br}{\bm r}
49: \newcommand{\bS}{\bm S}
50: \newcommand{\bg}{\bm g}
51: \newcommand{\wF}{\widetilde{F}}
52:
53: \title{Electrodynamics of Amorphous Media at Low Temperatures}
54:
55: \author{Vassiliy Lubchenko and Robert J. Silbey}
56: \affiliation{Department of Chemistry, Massachusetts Institute of
57: Technology Cambridge, MA 02139}
58:
59: \author{Peter G. Wolynes} \affiliation{Departments of Chemistry and
60: Biochemistry, and Physics, University of California, San Diego, CA
61: 92093-0371}
62:
63:
64:
65: \date{April 29, 2005}
66:
67: \begin{abstract}
68:
69: Amorphous solids exhibit intrinsic, local structural transitions, that
70: give rise to the well known quantum-mechanical two-level systems at
71: low temperatures. We explain the microscopic origin of the electric
72: dipole moment of these two-level systems: The dipole emerges as a
73: result of polarization fluctuations between near degenerate local
74: configurations, which have nearly frozen in at the glass transition.
75: An estimate of the dipole's magnitude, based on the random first order
76: transition theory, is obtained and is found to be consistent with
77: experiment. The interaction between the dipoles is estimated and is
78: shown to contribute significantly to the Gr\"{u}neisen parameter
79: anomaly in low $T$ glasses. In completely amorphous media, the dipole
80: moments are expected to be modest in size despite their collective
81: origin. In partially crystalline materials, however, very large
82: dipoles may arise, possibly explaining the findings of Bauer and
83: Kador, J. Chem. Phys. {\bf 118}, 9069 (2003).
84:
85:
86:
87: \end{abstract}
88:
89: %\pacs{02.50.-r, 05.40.Fb, 05.30.Pr, 05.45.Df}
90:
91: \maketitle
92:
93: \section{Introduction}
94:
95: Glasses are frozen liquids and thus lack long-range order, yet the
96: differences in material properties between amorphous materials and
97: crystals are often rather subtle. Crystalline samples themselves are
98: rarely flawless and thus contain a number of imperfections such as
99: point defects, dislocations or grain boundaries of various
100: sorts. These tend to further mask the difference. The size of defects
101: in crystals ranges over many scales, while in glasses, the static
102: heterogeneity in the atomic arrangement appears comparable to the
103: molecular size itself. Simple molecular glasses thus seem perfect
104: candidates for description as isotropic continuum, at long enough wave
105: lengths. For instance at cryogenic temperatures, when the de Broiglie
106: wave-length of a thermal phonon at $\sim$ 1K exceeds the lattice
107: spacing by three orders of magnitude or so, continuum theory would be
108: thought to hold to high accuracy. Yet surprisingly, there clearly
109: exist degrees of freedom numbering in great excess of the Debye
110: density of states, leading to extra heat capacity and phonon
111: scattering in all amorphous materials \cite{ZellerPohl}. Here we
112: examine the electrodynamics of these degrees of freedom.
113:
114: Since Rayleigh scattering is too weak to account for the observed
115: magnitude of sound attenuation in glasses, internal resonances must be
116: involved, in the form of {\em anharmonic} structural rearrangments, in
117: order to explain the data. The well known, empirical two-level system
118: (TLS) theory presumes such resonances exist
119: \cite{AHV,Phillips}. Simply postulating a flat energy spectrum and a
120: frequency independent coupling to the phonons accounts for all the
121: gross features of the low $T$ anomalies (for reviews, see
122: \cite{LowTProp,HunklingerRaychaudhuri,Esquinazi}). Direct microscopic
123: evidence of the two-level nature of such entities comes both from the
124: phonon echo experiments \cite{GG} and, relatively recently, from
125: single-molecule experiments at cryogenic temperatures (see
126: e.g. \cite{BTLBO}). At these temperatures, the TLS picture is
127: internally consistent in so far as the structural transitions (ST) can
128: be defined as {\em local}, and thus, tautologically, sufficiently {\em
129: weakly} interacting. One may therefore speak of a multilevel system at
130: the location of each transition whose behavior reduces to a TLS
131: behavior at low enough $T$. We may call this a tunneling center
132: (TC). In 1986, Freeman and Anderson \cite{FreemanAnderson} showed that
133: the magnitude of the TLS density of states is apparently correlated
134: with the phonon coupling. This results in a universality of the ratio
135: of the phonon mean free path $l_\smfp$ to its wave-length $\lambda$:
136: $l_\smfp/\lambda \sim 150$, for all insulating glasses at $T \lesssim$
137: 1K. This universality seems hardly coincidental \cite{YuLeggett},
138: however understanding the origin of the universality requires a
139: microscopic picture of molecular motions in glasses. (The large size
140: of the factor $\sim$150, too, was a puzzle \cite{YuLeggett}.)
141:
142: The Random First Order Transition (RFOT) Theory of the glass
143: transition \cite{KTW, MCT, MCT1, XW, LW_soft} provides an appropriate
144: microscopic picture of the motions in glass. Most commercial and
145: laboratory glasses are made by quenching supercooled melts. In the
146: deeply supercooled regime, most liquid motions are activated
147: transitions between distinct aperiodic states of comparable energy,
148: during which the current structural arrangement in a {\em local}
149: region is replaced by another, quite different arrangement that
150: nevertheless fits its environment. The size of the reconfigurable
151: region, $\xi$, grows with decreasing temperature, and reaches about
152: 5-6 molecular units across by $T_g$, i.e. the glass transition
153: temperature corresponding to the 1 hour time scale. This size is
154: predicted to be universal, within logarithmic accuracy, for {\em all}
155: substances. At any point in time, above $T_g$, the liquid can be
156: thought of as a mosaic of such cooperative regions \cite{XW}, most of
157: nearly the same size, but otherwise with distributed barrier heights
158: and transition energies. Upon freezing, a particular mosaic pattern
159: sets in and undergoes relatively slow changes, called aging. The aging
160: speed depends on the quench depth \cite{LW_aging}. A sufficient
161: fraction of the structural transitions have small enough energy change
162: and barriers so that when they occur, they can account for the
163: cryogenic anomalies, some of which were mentioned above: the density
164: of states and the universality of phonon scattering \cite{LW}, the
165: Boson Peak \cite{LW_BP}, but also the anomalous Gr\"{u}neisen
166: parameter, the so called ``fast'' TLS systems and more
167: \cite{LW_thesis,LW_RMP}. According to the RFOT theory, the
168: universality of the $l_\smfp/\lambda$ ratio directly follows from the
169: universal cooperative region size $(\xi/a)^3 \sim 200$ at the glass
170: transition temperature $T_g$, where $a$ is the molecular length
171: scale. During a structural transition, a relatively large, $\sim 200$,
172: compact set of small units moves in a stage-wise fashion. This
173: corresponds to the motion of the domain wall, which separates the two
174: alternative arrangements, through the compact region. At cryogenic
175: temperatures, these motions occur by tunneling. Despite their
176: collective nature, such tunneling events are possible because of the
177: enormous multiplicity of alternative structural states and of
178: low-barrier paths connecting pairs of states: an amorphous sample
179: actually resides in a high energy density state, well above its lowest
180: energy, perfect crystalline state. Consistent with the facility of
181: tunneling is the smallness of individual atomic displacements during
182: each transition. Their amplitude is roughly equal to the Lindemann
183: length $d_L$. This length is typically one tenth of the characteristic
184: lattice spacing $a$ and is nearly the same for all substances. The
185: precise identity of the ``molecular unit'', or ``bead'' depends on the
186: specific substance, but usually corresponds to a few atoms. (See
187: \cite{LW_soft} for a detailed discussion.)
188:
189: In this article, we use the microscopic picture of the two-level
190: systems provided by the RFOT theory to estimate the coupling of the
191: transitions to external electric fields. Clearly, such a coupling must
192: be present because individual molecular bonds, that possess electric
193: dipoles, rotate during transitions. Since these couplings directly
194: enter into spectral hole-burning experiments \cite{Maier} and can also
195: be directly probed in single molecule experiments \cite{BTLBO}, it is
196: important to know how much collective excitations interact
197: electrodynamically with probes and external electric fields.
198:
199: \section{Interaction of a Single Tunneling Center with External Field}
200:
201: \subsection{Many-Body Origin of the Transition-Induced Dipole Moment}
202: \label{Motivation}
203:
204: To set the stage, let us briefly review the assumptions of the
205: traditional molecular models of dielectric response of insulating
206: media. One often assigns an electric dipole value to an individual
207: molecule, or to a molecular bond connecting distinct atoms in a
208: condensed phase. In a dilute liquid made up of polar molecules, the
209: medium polarizes in a field since the dipoles prefer to orient along
210: the field's direction at the cost of losing their rotational
211: freedom. Even without permanent electric dipoles, a dielectric
212: response occurs due to polarizability: An external field mixes in
213: higher energy molecular orbitals, which generally lack inversion
214: symmetry. Classically, this quantum mechanical response can be
215: imitated as two harmonically bound opposite charges that separate
216: after a field is turned on. In a polar substance, this polarizability
217: also changes the {\em length} or {\em orientation} of the permanent
218: dipoles.
219:
220: While in a fluid the dipoles can freely reorient, during a structural
221: transition (ST) in glass, the dipole is restrained: Each individual
222: bead within a tunneling center, or ``domain'', moves only about the
223: Lindemann length $d_L$, as illustrated in
224: Fig.\ref{domain_dipole}. Suppose, for the sake of argument, one can
225: break up the set of all the beads within the domain into distinct
226: pairs. During a transition, each such pair - and hence the
227: corresponding ``bond'' - rotates about $d_L/a \sim 0.1$ radian. The
228: amorphous lattice generally exhibits no symmetry. There is therefore,
229: typically, excess charge, however small, on each atom. Assume the
230: effective individual charges remain the same during such a
231: transition. One can thus unambiguously assign a permanent, {\em
232: point-like} electric dipole to each ``bond'' introduced above. As
233: schematically shown in Fig.\ref{domain_dipole}, a total dipole moment,
234: $ \bmu_\sT = \sum_i \Delta \bmu_i$, may be generated during a
235: structural transition, that would couple to an external electric field
236: $\bE$ with energy $- \bmu_\sT \bE$.
237: \begin{figure}[htbp!]
238: \includegraphics[width=.95\columnwidth]{domain_dipole.eps}
239: %\includegraphics[width=.85\columnwidth]{domain_dipole.pdf}
240: \caption{\label{domain_dipole} Shown on the left is a fragment of the
241: mosaic of cooperatively reconfiguring regions in the supercooled
242: liquid, with $a$ denoting the lattice spacing (more precisely ``bead''
243: spacing). $\xi$ is the cooperative region size; $d_L$ is the typical
244: bead displacement during a transition. (The shown magnitude of $\xi$
245: corresponds to a temperature near $T_g$ on 1 hour scale.) The two sets
246: of circles - solid and dashed ones - denote two alternative structural
247: states. The expanded portion shows how rotation of a bond leads to
248: generating an elemental dipole change during a transition, where the
249: partial charges on the two beads are $\pm \zeta q$.}
250: \end{figure}
251:
252: In molecular glasses, the bond dipoles are fairly easy to assign.
253: Generally, the assignment of point-like dipole moments to individual
254: bonds is, strictly speaking, non-unique. It would be rather difficult
255: in the case of a highly networked, covalently bonded substance, such
256: as amorphous silica, but as we shall see, other arguments for such
257: systems give similar results. These arguments use the measurable
258: piezoelectric properties of corresponding crystals to unambiguously
259: extract the coupling to fields. In weakly bonded molecular glasses
260: held together by Van der Waals forces, the point like dipole view is
261: already a good approximation.
262:
263: With this in mind, an order of magnitude estimate of the dipole moment
264: of a tunneling center can be made: The Coulomb charge on a bead does
265: not exceed a fraction $\zeta < 1$ of an elementary charge $q$, which
266: is close in magnitude to the electron charge $e$: $q \sim e$. An
267: individual electric dipole {\em change} is therefore $\Delta \mu \sim
268: \zeta \mu_\smol (d_L/a)$, where $(d_L/a)$ is the rotation angle, as
269: already discussed, and $\mu_\smol \equiv \zeta q a$ is the elemental
270: dipole magnitude associated with each bond. The number of pairs that
271: reorient in a structural transition is $N_\sd = (\xi/a)^3/2$. The
272: $\bmu_T$ distribution is, of course, centered at the origin. Since the
273: dipole forces are a small part of the energetics of the glass
274: transition, the dipole motions are expected to be only weakly
275: connected with each other. Therefore the individual dipoles $\Delta
276: \bmu_i$ make up a ``random walk'' of $N_\sd$ steps (in 3D). As a
277: result, the generic value of dipole change for the transition is given
278: by the width of the total displacements during such a walk:
279: \begin{equation}
280: \mu_T \simeq \zeta (q a) [(\xi/a)^3/2]^{1/2} (d_L/a),
281: \label{mu_Tqual}
282: \end{equation}
283: If the elementary dipole rotations are correlated, one may introduce
284: an additional factor - like the Kirkwood $g$ factor of liquid theory
285: \cite{Kirkwood}. $g$ is typically of order 2. At first we might
286: imagine a great deal of variability for the quantity
287: $[(\xi/a)^3/2]^{1/2} (d_L/a)$ but in fact it is nearly universally (!)
288: equal to unity for all substances. $a$ is typically a couple of
289: angstroms, implying $q a$ corresponds to $\sim 10$ Debye. ($e \AA
290: \leftrightarrow$ 4.8 Debye.) $\zeta$ is expected to be well less than
291: unity, with $0.1$ or less being a reasonable generic estimate. We thus
292: obtain that $\mu_T$ is of the order 1 Debye or less, consistent with
293: experiment. Note the magnitude of $\mu_T$ is rather modest - only of
294: order the size of a typical individual dipole moment - despite the
295: large number of particles constituting a domain. The relative
296: smallness of the two-level system dipole moment is due to the small
297: deflection angle $\sim 0.1$, and the apparent smallness of the partial
298: charge $\zeta q$. There are deep reasons for both: The former stems
299: from the particular magnitude of atomic displacements during a
300: transition: it is equal to the typical thermal displacement at the
301: mechanical stability edge, i.e. the Lindemann length \cite{XW}. The
302: latter is probably related to the intrinsic difficulty of making
303: ionically bonded aperiodic structures, which imposes an upper bound on
304: the value of $\zeta$; this will be discussed in due time.
305:
306: The sublinear scaling of the dipole moment with the domain volume
307: $\xi^3$, in Eq.(\ref{mu_Tqual}), is worth noting: The tunneling
308: transition dipole moment is not a bulk response, and its relation to
309: the material's average bulk dielectric properties, as encoded e.g. in
310: the substance's dielectric susceptibility $\epsilon(\omega)$ in the
311: fluid phase, is not immediately obvious. While it is ultimately the
312: deflections of the same elemental dipoles that both give rise to
313: $\epsilon(\omega)$ and the dipole moment, the causes of the
314: deflections differ. In contrast with the bulk polarizability, where
315: the deflection magnitude is proportional to the field, the transition
316: induced deflections are intrinsic and correspond to distinct local
317: structural states. (To illustrate this distinction further, we point
318: out that a transition can be induced by things other than an AC (!)
319: electromagnetic field - a thermal phonon for instance.) Above $T_g$,
320: the distinct structural states, that evolve into TLS at cryogenic
321: temperatures, are transient metastable structures that live typically
322: as long as the $\alpha$-relaxation time \cite{XW}. The transient
323: structures are (transiently) frozen-in elastic fluctuations.
324: Analogously, the intrinsic generated dipole moment may be thought of
325: as due to frozen-in electric fields - at $T_g$. (A formal connection
326: between generated electric field and mechanical stress is discussed in
327: the following Subsection.) We can use this line of thought to relate
328: the dipole moment to the bulk dielectric properties of the material
329: near $T_g$. Suppose the frozen-in field is along $z$ direction. The
330: corresponding dipole moment is related to this field via an
331: appropriate (frequency dependent) dielectric constant
332: $\epsilon_\sTLS(\omega)$, but also through the fluctuation-dissipation
333: theorem. A self-consistent closure of this relationship gives for a
334: spherical region of {\em volume} $\xi^3$ \cite{TitulaerDeutch}:
335: \begin{widetext}
336: \begin{equation}
337: - \frac{1}{k_B T_g} \int_0^\infty e^{- i \omega t} \frac{d}{dt} \la
338: \mu_{\sT,z}(0) \mu_{\sT,z}(t) \ra \, dt =
339: \frac{[\epsilon_\sTLS(\omega)-1][2 \epsilon_\sext(\omega)+1]}{4\pi[2
340: \epsilon_\sext(\omega) + \epsilon_\sTLS(\omega)]} \xi^3,
341: \end{equation}
342: \end{widetext}
343: where $\epsilon_\sext(\omega)$ is the dielectric susceptibility
344: outside the domain. The dielectric constant inside,
345: $\epsilon_\sTLS(\omega)$, contains both the usual (high frequency)
346: polarizability of the material and the polarization due to the dipolar
347: displacements accompanying the transition. Even though the interior
348: and exterior of the domain are chemically identical, it is necessary
349: to regard the two $\epsilon$'s as distinct, since we know a {\em
350: transition} occurs within the volume $\xi^3$. A similar relation can
351: be written for the volume occupied by a single elemental dipole, with
352: $\epsilon_\sext(\omega) = \epsilon_\sTLS(\omega)$:
353: \begin{widetext}
354: \begin{equation}
355: - \frac{1}{k_B T_g} \left(\frac{d_L}{a}\right)^2 \int_0^\infty e^{- i
356: \omega t} \frac{d}{dt} \la \mu_{i,z}(0) \mu_{i,z}(t) \ra \, dt =
357: \frac{[\epsilon_\sTLS(\omega)-1][2 \epsilon_\sTLS(\omega)+1]}{4\pi[3
358: \epsilon_\sTLS(\omega)]} (2 a^3),
359: \end{equation}
360: \end{widetext}
361: where the factor $(d_L/a)^2$ on the left reflects that the elemental
362: change in polarization, $\Delta \mu_i$, is related to the full
363: elemental dipole $\mu_i$ by the rotation angle $(d_L/a)$, see
364: Fig.\ref{domain_dipole}. The volume $2 a^3$ on the right corresponds
365: to the volume occupied by a pair of beads, as before. One gets, as a
366: result, a frequency dependent generalization of Eq.(\ref{mu_Tqual}):
367: \begin{widetext}
368: \begin{equation}
369: \la \mu_\sT^{(2)}(\omega) \ra = \la \mu_i^{(2)}(\omega) \ra
370: [(\xi/a)^3/2] \left(\frac{d_L}{a}\right)^2 \frac{[2
371: \epsilon_\sext(\omega)+1] 3 \epsilon_\sTLS(\omega)}{[2
372: \epsilon_\sTLS(\omega)+1][2 \epsilon_\sext(\omega) +
373: \epsilon_\sTLS(\omega)]},
374: \label{freq_dep}
375: \end{equation}
376: \end{widetext}
377: where the two-point correlation functions are the $t$ integrals above.
378: Note $\la \mu_\sT^{(2)}(0) \ra = \la \mu_\sT^2 \ra$ and $\la
379: \mu_i^{(2)}(0) \ra = \la \mu_i^2 \ra$. Finally, $\mu_i \simeq \zeta (q
380: a) \equiv \mu_\smol$, as before.
381:
382: Note that two adjacent regions are statistically unlikely to undergo a
383: structural transition at the same time. The physically preferable
384: choice for the external dielectric susceptibility
385: $\epsilon_\sext(\omega)$ is therefore the high frequency, electronic
386: component of the full dielectric response, which we call
387: $\epsilon_\infty$. With this, Eq.(\ref{freq_dep}) becomes
388: \begin{widetext}
389: \begin{equation}
390: \la \mu_\sT^{(2)}(\omega) \ra = \la \mu_i^{(2)}(\omega) \ra
391: [(\xi/a)^3/2] \left(\frac{d_L}{a}\right)^2 \frac{[2 \epsilon_\infty+1]
392: 3 \epsilon_\sTLS(\omega)}{[2 \epsilon_\sTLS(\omega)+1][2
393: \epsilon_\infty + \epsilon_\sTLS(\omega)]},
394: \label{freq_inf}
395: \end{equation}
396: \end{widetext}
397: Furthermore, the two-level systems that are active at low temperatures
398: correspond to the low barrier side of the barrier distribution. This
399: implies one should use the $\epsilon_\sTLS(\omega \rightarrow \infty)$
400: value with regard to the cryogenic phenomena and, therefore, no extra
401: frequency dependence appears in the coupling of the TLS to electric
402: field. As a result, we find no significant reaction field correction
403: to our earlier argument and obtain
404: \begin{equation}
405: \la \mu_T^2 \ra \simeq \mu_\smol^2 [(\xi/a)^3/2] (d_L/a)^2,
406: \end{equation}
407: c.f. Eq.(\ref{mu_Tqual}).
408:
409: \subsection{Piezoelectric View}
410:
411: To treat covalent network glasses, where the assignment of local
412: dipoles or changes is difficult, we now turn to a different way to
413: relate the dielectric response, within a domain, to frozen mechanical
414: stress. First consider a strictly periodic lattice that lacks parity
415: symmetry. Generally, the lattice strain induces internal electric
416: fields giving rise to piezoelectric behavior. In such a piezoeletric,
417: the energetics of the strain, in the lowest order, are described by
418: the free energy density (see e.g. \cite{LLcont})
419: \begin{equation}
420: \wF = \frac{1}{2} \lambda_{ik,lm} u_{ik} u_{lm} - \frac{1}{8 \pi}
421: \epsilon_{ik} E_i E_k + \beta_{i,kl} E_i u_{kl}.
422: \label{quadr1}
423: \end{equation}
424: Here $u_{ik}$ is the standard strain tensor \cite{LLelast},
425: $\lambda_{ik,lm}$ and $\epsilon_{ik}$ are the stiffness tensor and the
426: dielectric tensor respectively, and $\beta_{i,kl}$ is {\em a}
427: piezoelectric tensor. (Note, various sign conventions and free
428: energies have been used in the literature.) The double index summation
429: convention is implied throughout, with the exception of letters $x$,
430: $y$, and $z$, which will be obvious in the context. Finally, the
431: elastic constant $\lambda$ is related to the material's mass density
432: $\rho$ and the speed of sound $c_s$:
433: \begin{equation}
434: \lambda \sim \rho c_s^2.
435: \label{stiffness}
436: \end{equation}
437:
438: In the absence of external field: $\bD = - 4 \pi (\prtl \wF/\prtl \bE)
439: = 0$, - the internal electric field is simply proportional to the
440: strain itself:
441: \begin{equation}
442: E_i = 4 \pi \epsilon^{-1}_{ik} \beta_{k,lm} u_{lm}.
443: \label{E_spont}
444: \end{equation}
445: (Note, the $\beta$ tensor has the dimensions of electric field.) If
446: expressed in terms of strain only, the free energy density reads, in
447: the absence of external electric field:
448: \begin{eqnarray}
449: \wF &=& \frac{1}{2} \lambda'_{ij,kl} u_{ij} u_{kl}, \label{quadr2} \\
450: \lambda'_{ij,kl} &\equiv& \lambda_{ij,kl} + 4 \pi \epsilon^{-1}_{mn}
451: \beta_{m,ij} \beta_{n,kl}. \label{lambda_pr}
452: \end{eqnarray}
453: The total, apparent stiffness, $\lambda'$, can be decomposed thereby
454: into a purely ``covalent'' and a ``coulomb'', i.e. electrostatic
455: component. The latter contribution is ordinarily quite small, owing to
456: the smallness of the charges induced by lattice distortions. Consider
457: $\alpha$-quartz, for example. Here, only $\beta_{x,xx} = 5.2 \cdot
458: 10^4$ esu, and $\beta_{x,yz} = \beta_{x,zy} = (1/2) 1.2 \cdot 10^4$
459: esu, are non-zero \cite{Cady}. As a result, the ``coulomb''
460: contribution to the $(xx,xx)$ component of the apparent stiffness
461: tensor is only about one percent of the covalent counterpart:
462: $\lambda_{xx,xx} = 8.8 \cdot 10^{11}$ dyne/cm$^2$ and $(4
463: \pi/\epsilon_1^T) \beta_{x,xx}^2 = 7.4 \cdot 10^9$ dyne/cm$^2$.
464: (Here, we used $\epsilon^T_1 = 4.58$ \cite{Mason}.) The relative size
465: of the $\lambda$ vs. $\beta$ magnitudes can be understood as follows:
466: $\lambda$ reflects the energy (density) of the elastic restoring
467: force. It is essentially the second derivative of an individual
468: atomic potential. Interatomic bonding, be it coulombic or covalent in
469: character, is ultimately of electrostatic origin. One may therefore
470: associate $\lambda$ with the quantity $\frac{1}{a^3}
471: \frac{\prtl^2}{\prtl^2 r} (q^2/r)|_{r=a} = (q/a^2)^2$, where $q$ is
472: the effective charge giving rise to the bond, and the $1/a^3$ factor
473: in front provides for energy density. (The total {\em first}
474: derivative of the (full quantum) potential energy is zero, of course.)
475: By virtue of being an electric field, $\beta$ roughly corresponds to
476: the quantity $q'/a^2$, where $q'$ would be the partial charge
477: introduced in the previous subsecton. The ratio $q'/q$ corresponds,
478: within the present framework, to the earlier introduced quantity
479: $\zeta$. It follows that in $\alpha$-quartz, the partial atomic
480: charge is indeed about one tenth elementary charge, since the quantity
481: \begin{equation}
482: \zeta^2 = (q'/q)^2 \simeq \beta^2/\lambda
483: \label{zeta2}
484: \end{equation}
485: is approximately equal to 1/100 in silica, as we just saw.
486:
487: Now suppose for a moment that a relation similar to Eq.(\ref{E_spont})
488: exists between the bead displacements within a domain and the internal
489: electric field changes generated during a transition. (We stress, in
490: an amorphous sample such generated field changes are zero, upon
491: spatial average, but here we refer to {\em local} fields at a
492: particular, generally non-centrosymmetric site.) Since $d \wF = - \bD
493: d \bE/4\pi + \ldots$, the free energy change in the presence of a
494: (small) external field $\bD_\sext$ during the transition is given by
495: $\int_V dV \, \Delta \wF = - \bD_\sext \int_V dV \, \Delta \bE/4\pi$,
496: in the lowest order in $\bD_\sext$. (Here we have volume-integrated
497: over the reconfiguring domain that correspons with the two-level
498: dynamics; $\bD_\sext$ obviously varies sufficiently slowly within the
499: domain for realistic frequencies of light, and can be taken out of the
500: integral.) The relation of the field $\bD_\sext$ to the external
501: field proper depends, of course, on the experiment's geometry. We
502: will use an electric field $\bE_\sext (\br) = \epsilon^{-1}
503: \bD_\sext(\br)$ where $\epsilon$ is the average {\em bulk} dielectric
504: susceptibility (which, of course, is uniform and isotropic in an
505: amorphous material). The coupling to this field $\bE_\sext$ is
506: consequently given by:
507: \begin{equation}
508: \bmu_T = \frac{\epsilon}{4 \pi} \int_V dV \, \Delta \bE(\br),
509: \label{muT_formal1}
510: \end{equation}
511: i.e. the generated internal field difference during the transition,
512: integrated over the domain. Eq.(\ref{muT_formal1}), among other
513: things, demonstrates that one may indeed unambiguously assign a
514: collection of point-like dipoles to the bead set within a TC, namely
515: by virtue of the relation $\bE(\br) = 4 \pi c_1 \bmu(\br) n(\br)$,
516: where $n(\br)$ is the (coordinate-dependent) dipole concentration and
517: the constant $c_1 \sim 10^0$ should be chosen depending on a specific
518: way to incorporate the already mentioned cavity effects. In what
519: follows, we will outline the microscopic picture of interaction of a
520: transition with elastic strain, which will naturally lead us to the
521: formula above and the ability to estimate the electric moment via
522: material's piezoelectric properties.
523:
524: Since individual displacements $\bd_i$ during tunneling transitions
525: are only one-tenth of the lattice spacing, one can indeed describe the
526: corresponding additional elastic energy variations, due to the
527: presence of a phonon, by a quadratic form of the type from
528: Eq.(\ref{quadr1}) or (\ref{quadr2}). Further, here one computes {\em
529: relative} displacements, not the absolute atomic coordinates which are
530: generally difficult to calculate. Define $\phi_{ik}$ as the strain
531: tensor due to a (long-wave) lattice distortion of a stable lattice. In
532: addition, define $d_{ik}$ as the ``strain'' tensor corresponding to
533: the set of the tunneling displacements $\{\bd_i\}$. The full elastic
534: energy within the domain, given a particular domain boundary
535: configuration (call it $\Omega_b$) can be written as:
536: \begin{equation}
537: \wF = \frac{1}{2}\lambda'_{ij, kl}
538: [(\phi_{ij}+d_{ij})(\phi_{kl}+d_{kl}) - d_{ij} d_{kl}] + {\cal
539: H}(\{d_{ij}\}, \Omega_b),
540: \end{equation}
541: where the energy functional ${\cal H}(\{d_{ij}\}, \Omega_b)$ includes
542: all the non-linear, many-body interactions giving rise to the
543: existence of the many metastable structural minima within the
544: domain. The construction of a library of the states corresponding to
545: these minima was described in Ref.\cite{LW_aging}. The full
546: multiplicity of the local states reveals itself directly in
547: calorimetric measurements above the glass transition. Here, we are
548: only concerned with the two lowest energy states of a region of the
549: otherwise undisturbed lattice. The two correspond to the two lowest
550: minima of ${\cal H}(\{d_{ij}\}, \Omega_b)$. The size of the domain is,
551: as we have seen, chosen so that one is guaranteed to have at least one
552: alternative structural state of nearly the same energy, and was found
553: to be only slightly larger that the cooperative region size at $T_g$
554: \cite{LW}. Note, by construction, the boundary state $\Omega_b$ is
555: independent of the phonon field $\phi_{ik}$. The effect of an external
556: {\em mechanical} stress on the internal displacements within a local,
557: compact region is passed on through the boundary, and so the
558: interaction of the region with the stress can be expressed through a
559: displacement integral over the region surface \cite{LLelast}. The
560: cross $\lambda_{ij,kl} \phi_{ij} d_{kl}$ term in Eq.(\ref{quadr2})
561: gives the amount by which the energy of a tunneling transition is
562: modified by the presence of a phonon. This therefore gives the
563: TLS-phonon coupling. The latter coupling was estimated in this way in
564: Ref.\cite{LW,LW_RMP}. A direct computation of the
565: $\phi_{ij}$-field-induced change of the transition energy gives:
566: $\Delta E(\phi_{ij}) = \lambda'_{ij,kl} \phi_{ij} \int_V dV \,
567: d_{kl}$, where we have integrated over the domain volume and taken
568: advantage of the elastic strain being $\phi_{ij}$ nearly constant
569: throughout the domain ($\lambda'$ in the latter equation is the
570: domain-averaged value of the atomic force-constant.) The coefficient
571: at the $\phi_{ij}$ gives a (tensorial) coupling of a transition to
572: strain according to:
573: \begin{equation}
574: {\cal H}_{\sTLS, \sph} = g_{ij} \phi_{ij} \sigma_z,
575: \label{H_int}
576: \end{equation}
577: where $\sigma_z = \pm 1$ is the usual Pauli matrix, and
578: \begin{equation}
579: g_{ij} = \frac{1}{2} \lambda'_{ij, kl} \int_V dV \, d_{kl}.
580: \end{equation}
581: The volume integral above indeed reduces to a surface integral of the
582: tunneling displacements. Consider for example, the term $\int_V dV \,
583: d_{xy} = \int_S (d_x dx + d_y dy) dz$ etc. The coupling to the {\em
584: longitudinal} phonons has the most vivid form, since $d_{ii}$ is the
585: divergence of a vector field, i.e. $\bd$. One gets $g_{ii} \simeq
586: \lambda \int_S \bd \, d\bS$. The atomic displacements at $T_g$ are
587: typically near the Lindemann length: $d \simeq d_L$, - but can also be
588: expressed in terms of the elastic constants, since the amount of
589: elastic energy contained in a unit cell is determined by the
590: temperature itself: $a^3 \lambda (d/a)^2 \sim T_g$. (This is by virtue
591: of the fluctuation-dissipation theorem.) Estimating the surface
592: integral \cite{LW, LW_RMP} introduces additional numerical factors and
593: gives, within a factor of two or so:
594: \begin{equation}
595: g \simeq \sqrt{T_g \rho c_s^2 a^3},
596: \label{g}
597: \end{equation}
598: where we have used Eq.(\ref{stiffness}). The result above is easy to
599: rationalize on general grounds: Any atomic motions in a dense liquid
600: (at $T_g$ and otherwise) are either a vibration or an anharmonic
601: motion that is part of a structural transition. The two excitations
602: must coexist and thus be marginally stable against each other, in
603: order for both to be present. Such a marginal stability criterion
604: gives $\la \sigma_z g_{ij} \phi_{ij} \ra \simeq \la \lambda_{ij, kl}
605: \phi_{ij} \phi_{kl} \ra$, as follows from optimizing Eq.(\ref{H_int})
606: together with the elastic energy $\frac{1}{2} \lambda_{ij,kl}
607: \phi_{ij} \phi_{kl}$ with respect to $\phi_{ij}$, multiplying by
608: $\phi_{ij}$ and thermally averaging (at $T=T_g$). Owing to $\la
609: \sigma_z \phi_{ij} \ra \simeq \la |\phi_{ij}| \ra$, Eq.(\ref{g})
610: follows.
611:
612: In the same way that the elastic fluctuations interact with the atomic
613: displacements, polarization waves and external electric sources will
614: interact with the internal electric fields generated in the domain
615: during a transition. As we have seen, the mechanical response
616: characteristics of a transition arise in response to stress
617: fluctuations at $T_g$. Analogously we can say, the electric moments of
618: the two-level system arise in response to the electric field
619: fluctuations at $T_g$. In full analogy with the
620: fluctuation-dissipation theorem context, the TLS dipole moment will
621: interact with an external field source just as it did with the
622: internal electric fields at the moment of freezing. We may thus
623: compute the dipole moment by substituting the generated electric field
624: from Eq.(\ref{E_spont}) into Eq.(\ref{muT_formal1}), bearing in mind
625: that the local lattice and the corresponding tensors are no longer
626: subject to any particular symmetry. Still, since the lattice locally
627: resembles a crystalline lattice, one may choose coordinates, again
628: locally, in such a way that the the $\beta$ tensor is maximally close
629: to a crystalline one. (Clearly, the $m$, $n$ sum in
630: Eq.(\ref{lambda_pr}) is independent of the coordinate choice.)
631: Therefore the latter sum will give a comparable result to that for a
632: crystal. Summing over the displacement tensor $d_{ij}$ (contracted
633: with the $\beta$ tensor) is quite analogous to the argument leading to
634: Eq.(\ref{mu_Tqual}). As a result one obtains the following qualitative
635: estimate:
636: \begin{equation}
637: \mu_T \simeq (\bar{\beta} a^3) [(\xi/a)^3/2]^{1/2} (d_L/a),
638: \label{mu_Tpiezo}
639: \end{equation}
640: where $\bar{\beta}$ is the local value of the piezoelectric
641: coupling. (We remind the reader that glasses are on average isotropic
642: and thus can not have bulk piezoelectric properties. It is only in the
643: frozen state that parity is locally broken.) The simple relation
644: \begin{equation}
645: \zeta q a \simeq \bar{\beta} a^3
646: \label{connection}
647: \end{equation}
648: establishes the connection between the ``molecular dipole'' view of
649: the previous Subsection and the ``piezoelectric'' analysis in this
650: Subsection, c.f. Eq.(\ref{zeta2}). Using the quartz parameters above,
651: one obtains that $\mu_T$ is, again, of the order Debye. Note that
652: formula (\ref{mu_Tpiezo}) uses quantities that can be measured
653: independently for substances which have a crystalline counterpart.
654: The bead size $a$ can be determined from the fusion entropy
655: \cite{LW_soft}. The piezoelectric constants are measurable too. The
656: two views - one based on molecular moments, the other on local
657: piezoelectricity - are somewhat distinct but are highly overlapping:
658: Since the local hybridization pattern on individual atoms is
659: intrinsically asymmetric in amorphous lattices, partial atomic
660: charges, however small, are always expected to be present in glasses,
661: giving rise to both local permanent dipoles {\em and} local
662: piezoelectricity. Mixing in a dipolar species would enhance both
663: effects. According to Ref.\cite{Schickfus_OH}, the two-level
664: systems's dipole magnitude is very correlated - nearly proportional -
665: to the OH$^-$ ion concentration, in amorphous silica with OH$^-$
666: impurities. Yet extrapolation to small ion concentrations shows TLS in
667: silica exhibit an {\em intrinsic} dipole moment, as was later
668: confirmed by an electric dipole echo study \cite{Golding_dipoleecho}.
669:
670:
671:
672: \subsection{Electrodynamics and Electroacoustics: Connection with
673: Experiment}
674:
675: In order to discuss experiments of two-level systems in glasses,
676: involving external fields, let us first recapitulate a few aspects of
677: the traditional phenomenological description, along with the
678: microscopic explanation. The Hamiltonian of an isolated two-level
679: system, as usually written in the low $T$ glass context, is
680: \begin{equation}
681: {\cal H}_\sTLS = \frac{1}{2} \epsilon \sigma_z + \frac{1}{2} \Delta
682: \sigma_x + g_{ij} \phi_{ij} \sigma_z,
683: \label{H_TLS}
684: \end{equation}
685: where $\epsilon$ is the transition energy, $\Delta$ is the tunneling
686: matrix element. (The phonon part of the full Hamiltonian is given in
687: Eq.(\ref{quadr2}).) According to Refs.\cite{LW, LW_BP}, the TLS that
688: are thermally active at cryogenic temperatures have their splitting
689: $\epsilon$ distributed according to a simple Boltzmann-like law
690: $n(\epsilon) = \frac{1}{T_g} e^{-\epsilon/T_g}$. This roughly defines
691: the density of states of the tunneling transitions. The density of
692: states, as seen by calorimetry, is time dependent, because the
693: tunneling matrix elements $\Delta$ are widely distributed. According
694: to the semi-classical analysis in Refs.\cite{LW, LW_BP}, the
695: distribution is $P(\Delta) \propto 1/\Delta^{1+c}$, where $c \ll 1$ is
696: a small constant ($c \propto \hbar \omega_D/k_B T_g$). This
697: distribution is close to, but not precisely the same as the inverse
698: distribution $\propto 1/\Delta$ postulated by the phenomenological TLS
699: model. Including quantum effects reveals that the low splitting
700: two-level systems, i.e. those with $\epsilon \ll \Delta$, are special
701: in the following sense. In such regions, the excess strain energy of
702: the glass is concentrated in the domain wall itself, while the barrier
703: separating the two alternative structural states is not high enough to
704: keep the domain in any of the classical structural states as defined
705: in terms of the classical atomic coordinates. Such TLS, with depinned
706: domain walls, give rise to an extra piece in the combined $P(\epsilon,
707: \Delta)$ distribution \cite{LW_RMP}. They correspond to the so called
708: ``fast'' two-level systems, introduced early on phenomenologically
709: \cite{BlackHalperin} in order to rationalize certain quantitative
710: shortcomings of the original TLS model. The quantum depinning of the
711: domain wall has been also called ``quantum mixing'' \cite{LW_RMP}.
712: Finally, we note the Hamiltonian in Eq.(\ref{H_TLS}) leads to rich
713: relaxational behavior, due to both interaction with phonons and the
714: phonon-mediated interaction with other TLS's. This has been discussed
715: in detail previously \cite{relax}, \cite{HunklingerRaychaudhuri},
716: \cite{Phillips}.
717:
718: The effects of the interaction of the dipole moment with a {\em
719: static} electric field are actually quite difficult to observe under
720: routine laboratory conditions (see e.g. \cite{Stephens_dipole}). The
721: upper limit for the field is given by the dielectric breakdown value
722: and is generically $10^6$ V/cm. For the typical dipole moment of 0.5
723: Debye, this implies an interaction energy of only 10$^{-3}$ eV $\sim$
724: 10 Kelvin. The constant field strength normally employed is actually
725: an order of magnitude weaker, or less. In a field $\bE_\sext$, the
726: transition energy is modified according to $\epsilon \rightarrow
727: (\epsilon - \bmu_\sT \bE)$. Typically, $|\bmu_\sT \bE| < 1$K. This is
728: clearly inferior to the characteristic energy scale of the TLS
729: spectrum, namely the glass transition temperature $T_g$ \cite{LW}. The
730: effect of a constant field thus turns out to be very generic because
731: of the intrinsic flatness of the energy distribution: The angular part
732: of $\bmu_\sT \bE_\sext$ is, obviously, uniformly distributed
733: resulting, again, in a flat distribution of the field-modified
734: transition energy. In order to discern such a small energy variation,
735: a {\em resonance} technique must be employed. Just such an experiment
736: was performed by Maier at el. \cite{Maier}, who took advantage of the
737: possibility to burn very narrow holes - only a few MHz - in the
738: chromophore's inhomogeneous spectrum. These authors turn on the field
739: immediately after burning the hole and observe the (time-dependent)
740: hole broadening, whose overall magnitude depends quadratically on the
741: field strength. Maier at el. report the value of $\mu_T = 0.4$ D for a
742: PMMA matrix.
743:
744: It follows from our theory that light and sound couple to the
745: two-level system transitions in a very similar way, save the dipole
746: character of the TC-photon interaction distinct from the tensorial
747: coupling of the transitions to the phonons. Indeed, the temperature
748: dependence of the speed of light in vitreous silica, as obtained early
749: on in by Schickfus at el. \cite{Schickfus}, nearly coincides with the
750: corresponding ultrasonic data \cite{HunklingerRaychaudhuri}. This
751: strongly suggested, at the time, that both the electromagnetic and
752: acoustic anomalies had the same origin. That the coincidence is not
753: purely circumstantial was shown soon afterwords in a number of elegant
754: electro-acoustic experiments: Increasing the AC electromagnetic field
755: leads to saturation of the structural transitions and a decrease in
756: ultrasonic attenuation \cite{Laermans, Doussineau}. In addition,
757: exposure to the AC field affects the {\em acoustic} impedance of a
758: glass \cite{Doussineau}. Again, the sufficient sensitivity of these
759: experiments is due to the interaction with the AC field being
760: resonant. Finally we mention yet another venue in investigating the
761: TLS coupling to electric fields, namely the electric dipolar echo, see
762: e.g. \cite{Bernard_dipoleecho, Golding_dipoleecho}.
763:
764: The dipole moment magnitudes, reported in all these experiments on the
765: respective substances, are all of the order 1 Debye, although more
766: recent measurements seem to be converging on a fraction of a
767: Debye. Unfortunately, the extracted dipole values do not completely
768: agree between different experiments. So, for example, Kharlamov at
769: el. \cite{Kharlamov_dipole} give relatively low values of 0.2 D and
770: 0.1 D for PMMA and PS respectively, based on their field induced
771: spectral diffusion data. The degree of the quantitative discrepancy
772: is, of course, subject to the detailed assumptions on the distribution
773: of the individual TLS parameters, various angular averagings etc.
774:
775:
776: \section{Dipole-Dipole Interaction}
777:
778: The idea of local structural transitions is internally consistent in
779: that the transitions are indeed distinct, weakly interacting entities.
780: This is easy to understand by considering the moment of vitrification,
781: when a particular pattern of mobile regions sets in: A transition will
782: be found locally, upon freezing, if at $T_g$ it was of marginal
783: stability with respect to external mechanical perturbation, as
784: delivered by stress waves to the given local region. It does not
785: matter, of course, whether the source of these waves is thermal
786: elastic fluctuations or the other structural transitions. Now, upon
787: having estimated the energy spectrum of the TLS and their coupling to
788: the phonons \cite{LW}, one may check the magnitude of the resultant
789: TLS-TLS interaction, mediated by the acoustic waves in the frozen
790: lattice. Such interaction self-consistently turns out to be small
791: \cite{LW_RMP}. We are aware of several observable consequence of the
792: interaction. For one thing, this interaction tends to quench the
793: spontaneous echo generation \cite{BlackHalperin, GG}, by virtue of
794: dephasing each TLS's motion when the TLS precesses about its local
795: field. Another, remarkable effect from the interaction is that it
796: gives rise to a negative thermal expansion coefficient in some
797: glasses, at low enough temperatures \cite{LW_RMP}: The fluctuating
798: entities in the lattice attract in the Van der Waals fashion, via
799: exchanging phonons. This attraction, counterbalanced by the materials
800: stiffness, acts to partially contract the sample. The number of
801: thermally excited transitions increases with the temperature and
802: thereby enhances the degree of contraction. The effect is small,
803: about $10^{-6}$, but nevertheless observable. The often employed
804: dimensionless parameter characterizing lattice non-linearity - the
805: Gr\"{u}neisen parameter - is usually positive and of the order unity
806: in crystals. It was found to be large and negative in many glasses at
807: cryogenic temperatures, see e.g. \cite{Ackerman}.
808:
809: The direct phonon mediated TLS-TLS interaction goes as $1/r^3$, just
810: like the usual dipole-dipole interaction, but is somewhat complicated
811: by the tensorial form of the coupling (see e.g. \cite{BlackHalperin}
812: for a discussion). If, however, the transverse and longitudinal speeds
813: of sound were equal, the interaction would be strictly dipole-dipole.
814: It is therefore often convenient to assume the elastic interaction is
815: indeed purely dipole-dipole resulting in a small {\em quantitative}
816: error. With this simplification, a ``scalar'' phononic Hamiltonian
817: can be used: $\frac{1}{2} \lambda_{ij, kl} \phi_{ij} \phi_{kl}
818: \rightarrow \frac{1}{2} \rho c_s^2 (\nabla \phi)^2$, where $\phi$ is a
819: scalar displacement field polarized in a single direction. The
820: coupling will become $g_{ij} \phi_{ij} \sigma_z \rightarrow (\bg
821: \nabla \phi) \sigma_z$. (The tensorial character of the interaction
822: may actually be important in the Gr\"{u}neisen parameter context, see
823: \cite{LW_RMP} and below). The presence of an electric component to
824: each transition dipole moment clearly leads to another contribution to
825: the total interaction. Since a detailed discussion of the interaction
826: effects has already been given elsewhere \cite{LW_RMP}, here we simply
827: estimate the strength of the electric dipole-dipole coupling relative
828: to the purely elastic counterpart and the rest follows in a
829: straightforward fashion.
830:
831: The elastic dipole-dipole interaction is given by a simple formula,
832: see e.g. \cite{LW_RMP}, \cite{YuLeggett}:
833: \begin{equation}
834: {\cal H}_{\selast} \simeq \frac{g^2}{\rho c_s^2} \frac{1}{r^3}
835: \simeq T_g \left( \frac{a}{r} \right)^3,
836: \label{elast}
837: \end{equation}
838: where, for the sake of clarity, we eschew some numerical constants
839: (these could be found in Ref.\cite{LW_RMP}) and have used
840: Eq.(\ref{g}). The electric dipole-dipole interaction, by
841: Eqs.(\ref{zeta2}) and (\ref{connection}), is, on the other hand:
842: \begin{equation}
843: {\cal H}_{\select} \simeq \zeta^2 \rho c_s^2 a^3 \left( \frac{a}{r}
844: \right)^3.
845: \label{elect}
846: \end{equation}
847: What is the relative value of the two interactions? A useful rule of
848: thumb is that $g \simeq \sqrt{T_g \rho c_s^2 a^3}$ is of the order eV
849: for all substances. In silica, for instance, $\rho c_s^2 a^3$ is
850: several eV, the Rydberg scale being a convenient (and physically
851: justified) landmark. The $T_g$ of silica is 1500 K, i.e. slightly
852: larger than 0.1 eV. We therefore make an interesting observation that
853: the electric dipole-dipole interaction {\em can} be comparable in
854: magnitude to the elastic counterpart for polar enough substances.
855: This is despite the relatively weak contribution (1\% or less) of the
856: coulombic forces to the apparent mechanical stiffness. We will
857: speculate on the physical significance of this observation in the
858: final Section of the article, while for now, we limit ourselves to a
859: formal notion: The elastic dipole-dipole interaction is disadvantaged,
860: compared to the electric counterpart, due to the large $\rho c_s^2$
861: term in the denominator of Eq.(\ref{elast}): Phonons are not true
862: gauge particles.
863:
864: Leggett has emphasized \cite{Leggett} that the dimensionless
865: Gr\"{u}neisen parameter varies ``wildly'' between different amorphous
866: substances, in contrast with the nearly universal $l_\smfp/\lambda$
867: ratio. Lubchenko and Wolynes have argued \cite{LW_RMP}, this stems
868: from the Van der Waals attraction between the tunneling centers, which
869: is strongly enhanced by ``Boson Peak'' excitations. The total
870: attractive interaction consists of several contributions, is
871: temperature dependent, and is expressed in terms of various
872: combinations of the temperature, $T_g$ and the Debye
873: temperature. While it may be argued that there is an intrinsic {\em
874: upper bound} on the value of $\zeta$ (see below), there otherwise
875: seems to be little intrinsic connection between the polar and
876: structural characteristics of glasses, in general. As a result, the
877: electrostatic interaction from Eq.(\ref{elect}) is expected also to
878: contribute to the non-universality of the Gr\"{u}neisen parameter.
879: This notion is consistent with the sensitivity of the Gr\"{u}neisen
880: parameter in silica to the concentration of polar impurities
881: \cite{Ackerman}. Further, the magnitude of the negative thermal
882: expansivity is indeed larger in more polar mixtures according to
883: Ref.\cite{Ackerman}.
884:
885:
886: {\em Symmetry vs. Transient Piezoelectricity.} We have so far focused
887: on the dipole moment of one of the structural states {\em relative} to
888: the other, namely $\sum_i \Delta \bmu_i = \mu_\sT$, since it is what
889: determines the coupling of the transition to external field. We next
890: inquire whether there is a correlation between the degree of polarity
891: of a state with its absolute energy. For example, suppose for a moment
892: that the lower energy state is completely non-polar, so that the
893: two-level system dipole moment is completely due to the excited state.
894: Such a situation could be exploited experimentally: One could
895: supercool a liquid just enough so that it does not crystallize readily
896: (the way glassblowers do), then expose the sample to strong electric
897: field for a sufficient while, and then quench the sample below its
898: $T_g$. After that, remove the field. Clearly, the number of dipole
899: moments along and opposite to the field direction will differ. In
900: other words, a number of dipoles will be actually lined up in a
901: preferred direction, leading to a (weak) ferro-electric order. As a
902: consequence of this, removal of the field in the procedure above
903: should lead to sample's contraction (which is the sample's way of
904: minimizing the ferroelectric energy.) Some residual polarization will
905: appear as well. If the field is removed sufficiently fast, the sample
906: will also heat up some. (This, in a sense, represents an antithesis to
907: adiabatic demagnetization.) Such polarization will not take place if
908: the degree of polarity is uncorrelated with the energy of a structural
909: state, at least within the relevant energy range.
910:
911: While we can not, at present, rule out {\it \`{a} priori} the
912: ``pyroelectric'' scenario above, it seems rather unlikely for the
913: following reason. An amorphous sample, at least at long enough times,
914: should have inversion symmetry, on average. With this symmetry
915: present, no piezoelectricity, let alone pyroelectricity, can take
916: place \cite{LLcont}. The question therefore is whether an external
917: field breaks the inversion symmetry (locally) without inducing
918: crystallization. Perhaps it will, on short enough time scales, before
919: considerable aging takes place. An experimental study would settle
920: this issue. At any rate, even if present, ferroelectric order would
921: not affect the thermal expansion properties at low temperatures
922: (c.f. the discussion of the Gr\"{u}neisen parameter in
923: \cite{LW_RMP}). This attraction mechanism is temperature independent
924: and simply contributes to the effective molecular field at each TLS
925: site.
926:
927:
928: \section{Closing Remarks}
929:
930: In this article, we have outlined the microscopic origin of the
931: coupling of the intrinsic structural transitions in amorphous solids
932: to electric fields. The coupling stems from rotation of the molecular
933: bonds, within the region encompassing the transition, which generates
934: a net electric dipole. The molecular constituents of a glass, even if
935: intrinsically non-polar, are strained due to disorder. Therefore small
936: partial charges on each atom are expected leading to the presence of
937: electric dipole moments associated with individual bonds.
938:
939: A local structural transition occurs by moving a domain wall through
940: the region. The domain wall is a mechanically strained region
941: separating alternative structural states. Such strained regions are
942: frozen-in thermal fluctuations of the lattice. (Above $T_g$, the
943: life-time of such as frozen-in structure would not exceed the typical
944: $\alpha$-relaxation time in the liquid.) In an analogous fashion, the
945: dipole moment can be thought of as frozen-in polarization
946: fluctuations, or as the local piezoelectric response to local strains.
947:
948: The glass transition is driven by steric, i.e. mechanical
949: interactions, not primarily by electrical ones, which is reflected in
950: the smallness of the partial charge, given in the theory by the
951: dimensionless quantity $\zeta$. So, for example, the elastic modulus
952: can be represented as $\rho c_s^2 (1 + \zeta^2)$, where $\zeta^2
953: \lesssim 0.01$ is the contribution of the ionic forces to the overall
954: material stiffness. (Here, we discriminate between ionic (or hydrogen)
955: bonds, as in NaCl or H$_2$O, and covalent bonds, as in diamond.) In
956: spite of its apparent small contribution to the material's structural
957: integrity, the coulomb component could actually be comparable to the
958: elastic component of the interaction between distinct {\em
959: transitions}. Note such interactions are unimportant as far as the
960: identity of each transition is concerned, because the corresponding
961: regions are statistically far apart \cite{LW_RMP}. Nevertheless, this
962: weak interaction may be viewed, loosely, as the successful attempt of
963: the system to have avoided entropically costly local ferroelastic
964: order. With this in mind, and the relatively strong coulomb
965: interaction of structural fluctuations, the smallness of the partial
966: charges in many glasses may simply reflect the fact that purely ionic
967: compounds, such as sodium chloride or water, do not vitrify readily
968: but instead, form low density lattices. (Note, ice does amorphize
969: under high pressure, see e.g. \cite{Strassle}.) Conversely, this
970: imposes an upper bound on the $\zeta^2$ value.
971:
972: Consider now Eq.(\ref{lambda_pr}). In order for the lattice to be
973: stable, the $\lambda'$ matrix (with an $ij$ pair considered a single
974: index) should be positive definite. This means, if a particular
975: $\lambda$ is negative, the corresponding $\beta$ could be, in
976: principle, quite large. This would imply large induced dipole moments.
977: As just argued, the coulomb component is likely to be small in a
978: glass, because the latter is rather homogeneous. Defects in crystals,
979: however, can be more extended and can be highly anisotropic. In this
980: regard we wish to mention the single chromophore studies of spectral
981: drift in a Shpolskii matrix, by Bauer and Kador \cite{BauerKador}.
982: These authors have seen a transition (presumably due to a structural
983: defect) that generated a remarkably large dipole moment of 8
984: Debye. Note, the Shpolskii matrix is polycrystalline, not strictly
985: amorphous.
986:
987: Finally we remark there is more to electrodynamics of amorphous solids
988: than what could realistically be discussed in this communication. Here
989: we have analysed only the case of good electric insulators. Distinct,
990: interesting phenomena take place in semiconductors and metallic
991: glasses. As an example, let us mention an old experiment of Claytor
992: and Sladek \cite{ClaytorSladek}, who found enhanced ultrasonic
993: attenuation in As$_2$S$_3$ glass upon removal of electric field. This
994: extra attenuation was greatly reduced by infrared radiation, which
995: suggested that there is, we quote, ``atomic relaxation accompanying
996: electronic transition in gap states where injected carriers have been
997: trapped''. We leave this for future work.
998:
999: {\em Acknowledgments:} P.G.W. was supported by the NSF grant CHE
1000: 0317017. V.L. and R.J.S. gratefully acknowledge the NSF grant CHE
1001: 0306287 and the Donors of the Petroleum Research Fund of the American
1002: Chemical Society. We are happy to dedicate this bagatelle to Bob
1003: Harris, whose elegant contributions are many, but more than scan the
1004: range from the electrodynamical properties of matter, parity, and
1005: tunneling systems, all of which enter the question discussed here.
1006:
1007: \bibliography{/home/vas/tex/lowT/lowT}
1008: %\bibliography{lowT}
1009:
1010: \end{document}
1011: