l2 is dictated by the interactions between the chains with differently oriented dipole moments (separated by the minimum distance bl2). Interchain interactions are constituted by two components falling off with distance by the power and exponential laws. The former arises only with the proviso that the dipole moment orientations are noncollinear with respect to chain axes and can be described in the continual approximation. To this end, a chain of identically oriented molecules is regarded as two uniformly charged closely-spaced parallel threads with the charge densities qla and -qla, where the effective charge q and the dipole length / are related by the dipole moment value, //
3.3. Chain orientational structures on planar lattices
70
= ql. For the orientational structure presented in Fig. 2.4, the displacements of the two threads relative to one another, if projected onto the axes z and y, amount to /cosf9 and /sin£feinp (by multiplying these projections by q/a, the corresponding projections of the dipole moment density are obtainable). The electric field induced by two threads at the distance y reckoned along the axis y, is expressed as: E(y) = ^-(o, ay
sin 0 sin p, cosfl).
(3.3.6)
The interchain interaction of dipole moments coUinear to the chain axes decays exponentially with the interchain distance (see Eq. (2.2.11)) and has no contribution descending by the power law y'2 in Eq. (3.3.6). Summation over lattice sites and interchain distances involves the Riemann zeta-function
with/? = 3 and 2 (g(3) » 1.202, £(2) « 1.645). In the ground state determination for dipole systems, exponentially small terms are of prime importance: it is because of them that the interchain interaction energy depends on chain longitudinal displacements. On the other hand, exponentially small terms contribute only about 10% to the interchain interaction energy and can thus be neglected in the energy estimation. By virtue of this simplification, the interchain interactions are describable in a very simple analytical form. Indeed, the explicit expressions for the force matrix components G^- (/' = 1, 2) and <£,2 <S>jj• * co) \ 1 +
J
[cos2 8j + (l - 3 cos 2
*j |cos2 0j - sin 2 6jsin 2
+
<512 »
3
ab2
(3.3.8)
[cos^j cos 02 - sin #i sin 0 2 sin ^i sin p 2 ]
depend only on the parameters cq and %j f° r t n e molecular vibrations concerned, molecular orientations in a couple of chains, the separation a between the neighboring lattice sites in the chains, and the interchain distance b/2.128 The approximation exploited involves no dependence of force matrix components on the
3.3. Chain orientational structures on planar lattices
71
mutual displacement of lattice sites in neighboring chains, which substantially extends the applicability of relations (3.3.8). Let us define the Davydov splitting of spectral lines, AQ D a v , as a frequency difference between the symmetric and antisymmmetric collective vibrations. For a herringbone structure with &i = B\ and <% = n - substituting Eq. (3.3.8) into (3.3.2) and the use of the inequality AQDav « &b provide: AQ Dav = fis - n A * ^12. „ 12^(2 W0 -AT fcos2 G - sin2 9 sin 2 J . J *>o ab2 l
(3.3.9)
Now we pass on the analysis of the relations derived focusing on several particular cases of importance which enable us of correlating the calculated values with the available experimental data. CO molecules adsorbed on the (100) face of a NaCl crystal reside at the sites of a square lattice (a = b/2 = 3.988 A); at sufficiently low temperatures (T < 25 K), they have inclined orientations (0 = 25°) with alternating dipole moment projections onto the axes of the neighboring chains (q>\ = 0, q>i = 180°).28 For this system, the Davydov splitting of vibration spectral lines is determined as: Afi Dav «3.935(ar/a 3 )cos 2 .
(3.3.10)
The accurate allowance made for dipole-dipole interactions leads to:81 AQ Dav =3.995^/a 3 )(l-1.058sin 2 6»),
(3.3.11)
which relation only slightly differs from the approximate one (see Eq. (3.3.10)). Substituting &b = 2155 cm"1 and x = 0.057 A 3 (vibration polarizability of C O molecules in gaseous phase) into Eq. (3.3.10) yields AQ D a v «7.8 cm" 1 , in good accord with the observed quantity A Q D a v « 6 cm" 1 . 2 8 C 0 2 molecules adsorbed on the same surface form a rhombic lattice ( a = b/2 = 3.988 A ) with the orientational parameters (9= 65°, q>\ = 49.5°, ^ = 130.5 0 . 1 3 " 2 6 1 2 3 With such values o f the angular variables, expression (3.3.9) goes negative; with the corresponding vibrational parameters entered (ab = 2349 cm' 1 and % = 0.48 A 3 in gaseous phase 1 2 9 ), the Davydov splitting becomes about -26 cm" 1 . At the same time, a materially smaller absolute value of the quantity concerned w a s recorded in experiments: Afi D a v » -9.2 cm' 1 . 1 2 3 T o apprehend the discrepancy, one should take into account that the Davydov splitting scale is dictated by the difference of trigonometric functions (see the bracketed expression in formula (3.3.9)) which is
3.3. Chain orientational structures on planar lattices
72
very sensitive to the relation between surface-parallel and surface-normal components of dipole-dipole interactions. These components are differently screened by a substrate. To estimate the screening effect, we employ a simplistic model based on image force of point dipoles. Consider two point dipoles A and B, with the dipole moments jiA = jv'A + nA and nB = n B + ji B , the heights above the surface zA and zB, and the mutual displacement along the surface r. The overall interaction between them is contributed by the interactions with point dipole moments of their images, fiA- = n A - \i% and n B ' = n ' B - j i B , induced in the substrate bulk at the distances zA and zB from the interface between vacuum and a medium with the dielectric constant s:
f'Dip / D i p = Z ^ ( r •ff.fi K^ 1 1 [z)^(r + 2 z A K ^ + Z ) ^ ( r - 2 z B K ^ ] , a,f} = x,y, +— 21 + £
(3.3.12)
At s -» 1 or at zA, zB » r, the second summand approaches zero, as might be expected. In the other limiting case, at zA = zB = z « r, the following power series results (accurate to terms (z/ry): 2s ^Dip
-
\ +£
1-9
e-\(z\
MAMB
(3.3.13) 2 + 1+e
s
ap-
'z* J\'
J
VAM£
^?L,a,f}
a
= x,y.
As is obvious from the above relation, the screening effect of the substrate on interactions causes, in the limiting case z —> 0, the change in the surface-normal components by a factor of 2e/(\+£) and in the surface-parallel components by a factor of 2/(l+f), i.e., the renormalization ratio is equal to£. As the substrate dielectric constant increases, the interaction of surface-parallel components monotonically decays to zero, whereas the interaction of surface-normal components is enhanced reaching the maximum, viz. the interaction with a double dipole moment, for a metal (at e -> oo).130 Analogous renormalization factors are also of significance in the treatment of island-like particles on a dielectric substrate.131 With such effects included, formula (3.3.9) can be rewritten as:
3.3. Chain orientational structures on planar lattices
AQDav»
^ ' * ftJgcos2<9-sin2<9sin2 L l + e ab2
73
J,
(3.3.14)
J
which results in the understated absolute magnitude of the Davydov splitting, AQDav * -3.3 cm'1, for the system CO2/NaCl(100) (s = n2 * 2.31 with n « 1.52 denoting the refractive index of NaCl in the spectral range of stretching C02-vibrations). Analyzing the corrections of the order (z/r)2 in Eq. (3.3.14), one can notice that they somewhat attenuate the screening effect, so that the result tends to the experimental value, AQDav »-9.2 cm"'. As another illustration, we consider the two-dimensional aggregates of benzothiazolocarbocyanine dyes (BTCC) adsorbed on the (111) face of AgBr crystal. The face of interest presents a triangular lattice constituted by Ag+ ions, with a lattice constant a equal to 4.082 A. The separation between neighboring molecular planes within the stacks (regarded as chains), r0 = a • sin cp (with q> = 60°), matches the separation of Ag+ ion rows so that r0 « 3.536 A.132 The neighboring stacks can have either the same ( ^ = #>i)132 or different ( ^ = n- ) (here ju is a matrix element of the transition dipole moment):128 1 .^ 3£"(2),u2sinc> ^ 2 = - A O D a v « - hK'Hr2 • •i
.„ „ , . . (3.3.15)
nr0L
At q>i = q>\, the polarization P A vanishes and the only spectral line of symmetric collective excitations is observed. At cpi = n-
7l>
t a.2 n>.
(3.3.16)
The dipole moments of the first electronic transitions in polymethine dye molecules can be derived in the context of the long-polymethine-chain approximation:128134135
3.3. Chain orientational structures on planar lattices
74
y u=-(
2
A 2 ^c-c( /7 + 1 + 1 ) -
(3.3.17)
Here r c _ c « 1.247 A is the projection of a chain C-C bond onto the long molecular axis; n is the number of C atoms in the chain; X is the effective length of the dye end-groups which is easily calculable from their topological characteristics ,134"136 Given n = 3 and X - 4.58 for BTCC molecule, we come to /i = 10.4 D which estimate agrees perfectly both with the value observed in the experiment (/4xp =10.5 D) and with that calculated by the Huckel LCAO-MO method (j^ak = 10.2 D).132 For two stacks having the same molecular orientations, the numerical factor 34(2) in formula (3.3.15) should be substituted by 2; as a result, with cp = 60° and the parameters indicated above, the frequency shift yielded in the point-dipole approximation is co\2 = - 445 cm'1. On the other hand, the computer simulation of two parallel stacks, each consisting of ten BTCC molecules, gave the frequency shift arising from the interstack Coulomb interaction: cofj0 = - 446 cm"1.132 The excellent accord between these two values suggests that the long-chain approximation is quite efficient in the calculations of individual molecular characteristics of polymethine dyes and the point-dipole approximation is likewise justified when applied to the treatment of interchain interactions in aggregates. The intrachain dipole-dipole interactions of BTCC molecules are responsible for the frequency shift contributing to the first of Eqs. (3.3.8) as 2 ^ (3)//(l 3cos2(p)/(ha3), with its sign changing at
= 2<Wl2
* ^ 7 0 cm"1 which value exceeds the experimental
one by no more than 10%. To sum up, the analytical model proposed presents a rare instance when a rather simple treatment enables not only qualitative but also quantitative relationships between structure and optical properties to be revealed. We now turn to a more complex instance of the Davydov-split spectral lines corresponding to the bending vibrations of C0 2 molecules in a monolayer adsorbed on the NaCl(100) surface. A molecule C02 in gaseous phase exhibits two degenerate bending vibrations in the plane perpendicular to the long molecular axis.
3.3. Chain orientational structures on planar lattices
75
On adsorption, this axis is parallel to the unit vector specified by Eq. (3.3.1) and the orientations of the bending vibration appear as follows: *js
• [- sin (pj, cos cpj, O), eJp = {- cos Gj cos
./ = U
(Eqs. (3.3.1) and (3.3.18) define a triad of mutually orthogonal vectors). Evidently, the adsorption potential induced by the substrate removes the degeneration of the bending vibrations, oriented parallel to (e^) and inclined at the angle of 90° - 9 to the surface (ejp). The singleton frequencies were obtained by means of isotopic mixture experiments, that is, recording the IR spectra of mixtures of 12 C 16 0 2 and 13 I6 C 0 2 as a function of the concentrations. In the limit of infinite dilution of 12 16 C 02, the IR spectrum showed a single band atftfo= 2349.0 cm"1 in the range of the asymmetric stretching vibration, whereas in the range of the bending vibration the extrapolation yielded ax, =660.0 cm"1 and ax, =655.5 cm"1.137 Since the singleton frequency &b of the asymmetric stretching vibration on the surface is much higher than the corresponding values ax, and ax, of the two bending modes, the solutions of equation (3.1.16), with allowance made for several molecular modes, become approximately separated for v = 0 and v = s,p, respectively: i
fij'v
J'v'J
=njs.PJ •
(3.3.19)
J"
Eigenvalues and eigenvectors for the asymmetric stretching vibration of herringbone structure with &i = Q\ = 0and cpz = n- q>\ = (pare described by relations (3.3.2) and (3.3.8). In order to determine the eigenvalues Q/ and eigenvectors Sjv i of the two bending modes s and p we have to consider the following 4x4 matrix:
O
=
®\s,2s
(5 1.5,1 P
\s,\s
_<5 \s,2p
ls,lp
\s,2p
\p,\P
\s,2p
-3>\s,\p
\p,2P
®\s,2p ^ - O \s,\p 0)ip,2P
(3.3.20)
l/>>l/> )
It is convenient to introduce the two pair indices X = ±1 and p. = ±1 instead of/. We can now specify the eigenvalues and eigenvectors by means of the following analytical expressions:123
3.3. Chain orientational structures on planar lattices
76
n%i
=T[*1*,1,
+*Ip,lp + *{®U,2s
-®1/»,2/»)+MA1
?DJXn ^ (S/PUM
=•
(3.3.21)
XDXfj
*>ls,lp-M>l3i2p
S
2WXM v
-X
j
with A
X = {$>\s,\s -*>\p,\p+rip\s,2s
D X/x
®\s,ls -VlpAp 2
Substitution of SJvj
-®iplpf
+M®ls,2s -%p,2P)
+4®ls,lp
-M>Ut2pf
. (33 22)
+ M^X
'
>
(®l*,lp-'W l*,2p)
in Eq. (3.1.25) by the corresponding vector components given
by Eq. (3.3.21) leads to the integrated absorption of the spectral line with indices X and//: K2N
A
Afi =
2C0F cos 3
WxM
(3.3.23)
• K ^ D x M f a , +te2s)+(op
Jxp~\p\p
The peak frequencies, integral intensities, and Ay'
-Xe2p)\ Ay'
ratios for stretching and
bending vibrations of the CO2/NaCl(100) system are listed in Table 3.1 together with the measured values.123 The calculations involve the above-presented parameters of the system as well as the vibrational polarizabilities in gaseous phase, xo = 0.48 A3 and & = Xp = °-24 A3.129 The comparison of calculated and experimental data shows a good agreement both for the polarizations (except for the absolute value of the Ay'
Ay'
ratio for row 5 in Table 3.1) and for the
frequencies, especially in the case of bending vibrations. Some discrepancies for stretching vibration frequencies may be attributable to the already discussed screening effect of a substrate.
3.3. Chain orientational structures on planar lattices
77
Table 3.1. Calculated IR peak frequencies, integral intensities and A}s> j Af'
ratios for stretching (A
and S) and bending (X, //) vibrations in CO2/NaCl(100) monolayer using the parameters given in the text. The measured values are written in parentheses.'"
1{^H)
A (cm"1)
4>)
4»)
4*)/#)
A S
2354.7 (2349.06) 2332.7 (2339.89) 660.5 (659.26) 653.5 (652.78) 663.8(663.24) 654.6 (655.79)
0.074 (0.079) 0.102(0.105) 0.0028 (0.0045) 0.0030 (0.0050) 0.0010(0.0045) 0.0037 (0.0035)
0.044 (0.046) 0.133(0.133) 0.0016(0.0025) 0.0017(0.0035) 0.0134(0.0080) 0.0021 (0.0025)
1.695(1.72) 0.768 (0.79) 1.742(1.80) 1.742(1.43) 0.073 (0.56) 1.738(1.40)
0.1) (1,-1) (-1, 1) (-1,-1)
Chapter 4 Temperature dependences of spectral line shifts and widths Treating vibrational excitations in lattice systems of adsorbed molecules in terms of bound harmonic oscillators (as presented in Chapter III and also in Appendix 1) provides only a general notion of basic spectroscopic characteristics of an adsorbate, viz. spectral line frequencies and integral intensities. This approach, however, fails to account for line shapes and manipulates spectral lines as shapeless infinitely narrow and infinitely high images described by the Dirac 5-functions. In simplest cases, the shape of symmetric spectral lines can be characterized by their maximum positions and full width at half maximum (FWHM). These parameters are very sensitive to various perturbations and changes in temperature and can therefore provide additional evidence on the state of an adsorbate and its binding to a surface. Among the accepted mechanisms governing spectral line shapes are inhomogeneous broadening inherent in disordered systems and homogeneous broadening arising from the interaction of a localized or collectivized excitation with the environment. For vibrational excitations in an adsorbate, the environment is represented by the phonon thermostat of a substrate. Harmonic coupling between the vibrational displacements of atoms in the adsorbate and in the substrate can lead to a finite broadening of adsorbate spectral lines only with the proviso that the corresponding frequency falls within the quasicontinuous spectrum of the substrate. This condition is met for the low-frequency vibrations of adsorbed molecules which involve rotational or translational (along the surface) degrees of freedom. To account for the spectral line broadening for high-frequency stretching vibrations of adsorbed molecules, allowance should be made for the anharmonic coupling of these vibrations with the low-frequency modes of the same molecules (the latter have finite lifetimes due to the harmonic coupling with the quasicontinuous spectrum of the substrate). The situation becomes even more involved as a consequence of the collectivization of the high-frequency and low-frequency vibrations due to lateral intermolecular interactions. This wide range of questions is to be elucidated in the present chapter. The bulk of attention is given to the effects induced by the collectivization of adsorbate vibrational modes whose low-frequency components are coupled to the phonon thermostat of the substrate. This coupling gives rise to the resonant nature of lowfrequency collective excitations of adsorbed molecules (see Sec. 4.1). A mechanism underlying the occurrence of resonance (quasilocal) vibrations is most readily 78
4. Temperature dependences of spectral line shifts and widths
79
comprehended using the example of a system of harmonic oscillators in which the states of a certain oscillator, due to its coupling with all the others, have finite lifetimes. These problems are considered in detail in Appendix 1. Adsorption potential and lateral interactions convert rotational orientational states intrinsic in molecules in gaseous phase to hindered-rotation states typical of adsorbed molecules. Small vibrations around the minima of the hindered-rotation potential can in the rough be regarded as harmonic only providing sufficiently high reorientation barriers. Then it is admissible to treat orientational excitations as quasiparticles obeying the Bose-Einstein statistics and to apply the corresponding retarded Green's function technique. Actually, reorientation barriers are not, however, too high and a comparably small number of orientational vibration quanta fits in their heights. In view of this, thermally activated reorientations and subbarrier tunneling should be taken into account. Sec. 4.2 presents the corresponding technique for computing spectroscopic characteristics; it includes thermally activated reorientations as a stochastic process of a system's departure from subbarrier equidistant energy levels. Here we introduce the Markov approximation for retarded Green's functions which underlies an exact expression for a spectral line shape in the case of biquadratic anharmonic coupling between two oscillators in the phonon field of a substrate (the exchange dephasing model). Also, the perturbation theory for the Pauli equation is presented which refers to the subbarrier group of states. The results obtained are adjusted to the limiting case of weak interaction between the subsystem concerned and the phonon field of the thermostat so as to apply them to a description of one-sided temperature broadening for Hbond molecular complexes. For a better understanding of this section, Appendix 2 provides the classical picture of thermally activated reorientations and quantum tunnel relaxation of orientational states in the phonon field of a substrate. In Sec. 4.3, we generalize the exchange dephasing model to various cases of anharmonic coupling between high-frequency and low-frequency modes, and to the case of collectivized excitations in adsorbate. For the system H(D)/C(111), the highfrequency and low-frequency modes differ in frequencies by a factor of about 2. Therefore, cubic anharmonic coupling of these modes contributes materially to spectral function characteristics of high-frequency vibrations. The systems with sufficiently strong lateral interactions of high-frequency molecular modes are distinguished by a strong dependence of spectral function parameters on the molecular axis inclination to the surface plane. Lateral interactions of lowfrequency molecular modes also exert influence on the spectral function of highfrequency vibrations; an illustrative example is provided by the analysis of temperature-dependent shifts and widths of the Davydov-split spectral lines for 2x1 phase of the monolayer CO/NaCl(100). All relevant mathematical derivations (based on double-time Green's functions in the representation of Matsubara's frequency space) can be found in Appendix 3.
80
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules Here it is our intention to show that for a system constituted by substrate phonons and laterally interacting low-frequency adsorbate vibrations which are harmonically coupled with the substrate, the states can be subclassified into independent groups by the wave vector K referring to the first Brillouin zone of the adsorbate lattice.138 As the phonon state density of a substrate many-fold exceeds the vibrational mode density of an adsorbate, for each adsorption mode there is a quasicontinuous phonon spectrum in every group of states determined by K (see Fig. 4.1). Consequently, we can regard the low-frequency collectivized mode of the adsorbate, ft^K), as a resonance vibration with the renormalized frequency a>n and the reciprocal lifetime IK-
K
—
K<1
—
As]Bv
Fig. 4.1. Coupling of the adsorbate low-frequency mode with substrate phonons: K level of the adsorbate (a); initial quasicontinuous phonon spectrum of the substrate not perturbed by the adsorbate, bold lines designating the levels which correspond to the specified wave vector K (b); level shifts in the K subsystem caused by the coupling of the adsorbate K mode and substrate phonons (c).
Represent the Hamiltonian for low-frequency modes of the adsorbate and substrate as a sum of three terms:
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
He = H<{mo])+Hs
+
Hmt,
81
(4.1.1)
the first of them accounting for laterally interacting adsorbate modes: tfj.no.) = ^PM R
+
/m
1 £ [m(COjSR,R. Z
^
+
*tM{R
- R')] «,(R) «,(R')
R,R'
=2^f(K)(^+(K)6e(K)+l/2). K Here K/(R) and />;(R) denote the shift and generalized momentum for the molecular vibration of the low frequency a>9 and reduced mass m,, at the Rth site of the adsorbate lattice; 6/+(K) and 6/(K) are creation and annihilation operators for the collectivized mode of the adsorbate that is characterized by the squared frequency c»/2(K) = cof + 5> /jat(K)/m/, with O /,iai(K) representing the Fourier component of the force constant function <J>/iat(R). Shifts w/(R) for all molecules are assumed to be oriented in the same arbitrary direction specified by the unit vector e/; they are related to the corresponding normal coordinates, u( (K), and secondary quantization operators:
2
R
«,(R) = ( i V V o H ] T « , ( K y * - , ut{K)=
(
ti V / 2 / \ ^2co- e(K) ^ | (^(K)+^ + (-K)), (4.1.3)
where N0 is a number of adsorbate lattice sites in the main area. The second summand in the right side of Eq. (4.1.1) describes substrate phonons:
^ = Z to ^ii.^W + v 2 }
< 4 - L4 )
k||j(T
In the above relation, quantum states of phonons are characterized by the surfaceparallel wave vector kg, whereas the rest of quantum numbers are indicated by a; the latter account for the polarization of a quasi-particle and its motion in the surface-normal direction, and also implicitly reflect the arrangement of atoms in the crystal unit cell. A convenient representation like this allows us to immediately take advantage of the translational symmetry of the system in the surface-parallel direction so as to define an arbitrary Cartesian projection (onto the a axis) for the
82
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
shift of a substrate atom with the mass M residing at the Rth site of the adsorbate lattice:
^ ( R ) = (MVo)- 1 / 2 2 S u n ,, e ' k l l ' R ^( k ll'^ k„,
t M5(k,|,
(here S£
.
V/2,
(415>
.
2cos(kn,ff)
are unitary matrices affording the switch to normal coordinates
Ks(k||, (T)). In this representation, the third summand in the right side of the Eq. (4.1.1) which describes a harmonic coupling between low-frequency vibrations of the adsorbate and substrate assumes a substantially simplified form: "int=
Z<e"
M
^
R
^(
R
)
R,a,J3
-{Mm^
£
e^tt^M^^WY.^^
•
The first Brillouin zone for vectors k||, being determined by the crystal unit cell, it can be larger than that for the adsorbate lattice, and hence the sum over R entering into the Eq. (4.1.6) is found as 2V(VK)R="oZVK,B, R B
(4-1-7)
where summation over B=«|Bi+«2B2 is performed over all inequivalent integer linear combinations of the vectors Bi and B2 of the adsorbate reciprocal lattice. This allows the classification of the wave vector set, kg, into vector subsets B-K, and expression (4.1.6) can hence be rewritten as follows: "int 4 l k ^ ( K K ( K . " ) ^ K > ? ( K f e ( K , v ) ] , K,v where
(4.1.8)
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
XK: = {Mm()~
Z/
f
83
(4.1.9)
int^B-K.cr
and the new quantum number v serves instead of the quantum number couple
S
s(K>v)=^CKv,K<,XKq
•
(4.1.10)
Unitary matrices C^q and C KvK? are diagonal in K, the former transforming from adsorbate to low-frequency system normal coordinates and satisfying the equations: K ? -®< (KJ-SKFKJ.MC.K, =0.
!-gk(4=<
CK,KJ
= 1 , (4.1.11)
where \XK,V\
*K(*) = E e-4(K,v)' v
(4.1.12)
Let us now invoke the conception of systems with a quasicontinuous spectrum developed by Lifshits.139 Since energy levels SYJ^CO^ for a perturbed system and levels £Kv=ct)s%K,v) for the corresponding unperturbed one alternate (see Fig. 4.1), the position of an arbitrary level £Kll on the interval (eKv, sKv + AeKv) can be reckoned from the level sKv and characterized by the parameter £Kq varying from 0 to 1: £Kq-£Kv+CKq&£Kv The function gK(£k<j) ar>d its derivative are hence expressible as:13 £ K ( * K ? ) = P K ( % V ) + *J7k(*Kv)«>t
(4.1.13)
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
84
^SK{£KV)
1
W=i-
SK^K?
^K,
d?Kq (4.1.14)
Detailed derivation of formula (4.1.13) with an emphasis on the origin of the term cotTT^Kq is given in Appendix 2 (see Eq. (A2.83)). Formulae (4.1.14) are obtained by substituting characteristics of the quasicontinuous spectrum into the first of Eqs. (4.1.11), with regard to Eq. (4.1.13) and the relationship between perturbed and unperturbed interlevel gaps, Ae^'1 = [AeKv'1 - d^ki/d^K,,]"1. The quantities fj K(e) and P K(e) are calculable by the equations: VK(£) =
^gK{£
+ iO) = Y\zKv\
#(£-£Kv\ (4.1.15)
V
J
E-E
where Re and Im denote real and imaginary parts of a complex function, respectively. Fig. 4.1 schematically depicts the alternating levels £Kl,and %, which belong to the group of K-dependent states. Incorporating expression (4.1.13) into the equation gK(£Kq)
= £
Kq ~0}i'(K)
which follows from relation (4.1.11) and calculating the derivative in Eq. (4.1.11) in view of Eq. (4.1.14) we arrive at:l40 ^K.(£Kq)
QKKq\
AEiKq
(4.1.16) WKq^Kq 2
2
PK, - « / ( K ) - % ( % ? ) ] + ^ ( ^ 1 where
»W(®K9-<»K;'7K)A«,K9.
4.1. Resonant nature of low-frequency collective excitations of adsorbed molecules
K&rih~
2
? ,„
85
(4.1.17)
is the resonance function, co^ is the resonance frequency obeying the equation a>l-col(K)-PK(a&)=0,
(4.1.18)
7K=—^K(4).
(4.1.19)
and the parameter
if multiplied by 2n, accounts for the full width of the resonance function. As the Fourier transform of the resonance function (4.1.17) takes on the form 00
« M =— 2K
Mco-ri)e-ia,tda> = —e~n^2, J In
(4.1.20)
— 00
the lifetime of the resonance mode with the wave vector K is found as rK = 2/^ K . The quasimode approximation is justified only in the case of ^K « <WK, i.e., for long lifetimes of resonance modes. Then resonance function (4.1.17) is narrow enough and all slowly varying functions of conq can be substituted by their values at the point t%? = <%. It has been just this procedure that has led us to the last approximate equality in relation (4.1.16). Importantly, the summation over q in expressions containing IC^Kql2 corresponds, in view of relation (4.1.16), to the switch to integral sums involving the resonance function which, in turn, can be replaced by integrals. Thus, if F(a)) is an arbitrary slowly varying function of co, then 00
X H ^ C K . K ^ q
\F{a>)fR{a>-a>K;TiK)da> = F(a>K)
(4.1.21)
-oo
and the result is independent of 77K. In the special case that F = 1, we revert to the corresponding normalization condition. Explicit expressions for the dispersion laws COK and resonance width rfo can be deduced for a model system, with the substrate simulated by a semi-infinite elastic
86
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
continuum and the adsorbate overlayer presented as an array of point masses connected to the surface by harmonic springs.141142 These expressions are strongly coverage-dependent and predict relaxation rates in good quantitative agreement with available experiments. A remarkable feature of this treatment is that no energy transfer from the adsorbate to the surface is possible at K > coilc-y (cT is transverse sound velocity) and hence rjK = 0 in this region.143 Averaging over all the wave vectors K of the first Brillouin zone furnishes the resonance width of a single adsorbate (see formulae (4.2.25) and (A1.109)).
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase Practically, any experimental study of an arbitrary system reduces to measuring the response of some physical characteristic A of the system to a probing external action which perturbs, in the general case, another characteristic B. The required response is then described by a conveniently calculable retarded Green's function (GF) that contains sufficiently complete information about the states of the system:144
G(t)^-i9(i^A(t\B{0^,
A{t) =
expUHt\Acxp(-^Ht\
(...) = Sp(p ...), p = exp(- ///r)/Sp[exp(-
H/T\
where A and B are the operators of the physical quantities A and B, and p is the equilibrium statistical operator of the system determined by its Hamiltonian H and the absolute temperature T (in energy units). In many cases one is interested in the properties of a molecular subsystem with a Hamiltonian H^moli which is coupled through an interaction HmX = V with the remaining large part (surface reservoir) of the total system described by the Hamiltonian Hs. We have then
H = Mmolhffs + V
(4.2.2)
and the operators A and B occur only in the Hamiltonian # ' m o " . This fact by itself does not lead to any simplification of Eq. (4.2.1), since the operator V containing the variables of the subsystem and of the reservoir does not commute with //( mo1 ) or Hs .
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
87
On the other hand, there exist well-developed methods for calculating states of subsystems using the Markov approximation for the reduced density matrix (statistical operator) of the subsystem p^mo' = Trp (Tr indicates the trace over the variables of the reservoir).145 The average value of a physical quantity A{t) will now be determined by the trace of the product of operator matrices for the subsystem only:
A{t) = Sv{p(mo^A)
(4.2.3)
and the response to an external action described by the time-dependent operator Hg(t) will be included in the operator / r m o >(t) satisfying the equation
h
dt
J
drTv
AV,
\ Is exp
-(£( m o , ) + £ 5 ) r AV, (4.2.4)
xexp
AV = V-(v)
•$t^hHs)r],fH)
where p$ is the equilibrium reservoir operator (the expression with a nonvanishing operator (v\
is derived in detail in Ref. 61).
We find the linear response of a subsystem in contact with a reservoir to an external perturbation involving some variable B of the subsystem and depending on the time through a function F(t), so that the corresponding perturbation operator can be written in the form HB{t) = -BF(t).
(4.2.5)
We substitute (4.2.5) into Eq. (4.2.4) and look for a solution of the latter in the form
p(moi\t)=p0+Sp(t),
88
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
where p0 is the equilibrium statistical operator of the subsystem. The matrix elements of the linear response Sp(t) to the perturbation HB(f) represented in the basis of the eigenstates of the Hamiltonian H
#( mo1 )
{8pqq\t)={q\Sp(f\q'Y
| ) = ^ 1 ^ ) ) w ' " satisfy the following equation:
dt
+
SPgg^+^Pgg^W^'
=
W)
l V f e ' - A,H).
(4.2.6)
qq
Here p^ is the diagonal matrix element of the equilibrium statistical operator p0 of the subsystem and the matrices Q and T with four indices are given as follows (see, e.g., Ref. 61 where these quantities are obtained taking into account the nonvanishing operator (v)
and the frequency shifts arising from principal-value
integrals which are usually neglected in other sources): ^•qq'qq'
^qq'°qq°q'q'
it)?/™-(%S1'«
^
F 1
qnq{<*>) ,
. Fqrrq'i®)
«*+»""
Q.q'r-0)
S~w+
dco
27th
^ -
'8,
qq
F
q'q'qq{a>) Pq'q'qqi^)}
Qqq+O)
Q?V-0)\'
qq'qq' ~ 77T1 /LiI W \ qr rq'q' +
2h
qrrq'^p-q'r J°qq J
{r ~ q'q'qq V w /
F
q'q'qq v^q'q'lr
00
Fqq'qM= fa™ (fVff.^V^
(4.2.7)
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
89
We introduce the GF (with four indices) for the left-hand side of Eq. (4.2.6):l46
-^gqq'gg'(t)+ X ZimWi'i^Wl'
+r
<7,«?2W')= - ^ ) * W < W
(428)
<7l92
in terms of which one can easily express the linear response Spgg'(t):
9?'-oo
Substituting (4.2.9) into (4.2.3) we are led to the usual way of writing down the time-dependent average value of a physical quantity A{t) as the linear response to the perturbation (4.2.5): m 00
A{t) = S p ( p 0 ^ ) - i jdt'G(t-t')F{t')
(4.2.10)
in which the required GF (4.2.1) is determined by the following expression:
(')=' Y,Aq,ggqW{t)Bw(p?-pg).
(4.2.11)
Together with Eqs. (4.2.7) and (4.2.8), Eq. (4.2.11) solves the given problem of finding the GF of a subsystem in the Markov approximation.
4.2.1. The exact solution for the exchange dephasing model To begin with, we consider how Eqs. (4.2.7), (4.2.8), and (4.2.11) reproduce the main results provided by the known exchange dephasing model147148 as regards the spectral line shape for a high-frequency local vibration.149150 The Hamiltonian of this model with a high-frequency mode coh and a single low-frequency exchange mode coi is given by Eq. (4.2.2) in which components have the following form:
90
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
//(mo1) = ticoh^>t,bh +1/2)+ hafpfb, +1/2)+ hyblbhbtb,, As = Y.h0)Mbk
+ V2),
(4-2.12)
V^^blbk+xlkbt). k
The high-frequency mode is coupled through the anharmonicity coefficient y with the low-frequency mode which is a resonant one due to harmonic interaction with the surface reservoir. For simplicity, the last term is written in the Gaitler-London approximation (compare with more general Eq. (4.1.8) where this restriction is absent). The required GF of the high-frequency mode can be obtained from the general Eq. (4.2.11) for A = bh, B = b% and the GF gqq'qq> with
q-{nh,ni)
( b hh\ n h) = nh\nh). bfbt\«/) = n,\ni)): G
(0='X^/»/(/)^»/'
g
»i»'i(')=^i"/,o«/,i«;Jo«;(0-
(4.2.13)
n,n}
In this formula, we have used the inequality hcoh » T which makes it sufficient to consider only the states with nh = 0 and 1, and also to put p\n = 0 (the quantity p0n/ is denoted by pn/). Calculating the matrix elements (4.2.7) by Eqs. (4.2.12) and substituting them in Eq. (4.2.8) we arrive at: ^S/i,*; (')+{'(©* +?"/)+fo + («/ + ^ K k i f , i i ; ( 0 - («/ + iHg», +i, n) (0 - "/^oSn, -1,»; (')=-^('K,!,; w
o = *k»i )/0 - 5 ) .
J
2
7(®/)= *X! k* I ^
(4.2.14)
ffl
" *)
it
# = exp(-»a>//7'). The structure of the left-hand side of Eq. (4.2.14) is the same as the homogeneous equation of Ref. 148 for the matrix elements PQH j n / of the reduced density matrix of the subsystem considered. A numerical solution of that equation was given in Ref. 151. Exact analytical expressions for the spectral line shape of a highfrequency local oscillation were obtained in terms of this model by the generating-
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
91
function149 and temperature-GF150152 methods. (The latter approach leads to an integral equation representing the boson version of the equation written down many years ago by Nozieres and De Dominicis'53 to solve the x-ray edge problem). Using the approach of Ref. 149 one can obtain an analytical expression for the GF (4.2.13) in the case of several low-frequency exchange modes &>; (j = \,2,...,p) which interact with one another neither directly nor through the reservoir:
*=f l - # y e x p ( - i v / )
(4.2.15)
Here we have \
( i
2
4= i + 4 -r 1 + 2/i+4 r M = - + ;— 24 1~4V 1-4 n 1-4
r
2\
(4.2.16)
(to simplify the notation, we have omitted the index j of the quantities £ y, t], /u, £, , X, and v). Eq. (4.2.15) can be simplified in the practically important case ofp degenerate exchange modes which is realized, e.g., for molecular complexes with hydrogen bonds. The high-frequency mode coh is represented by the valence or deformation vibration of a hydrogen atom which is anharmonically coupled with four (p=4) degenerate libration modes whose low frequencies are caused by the relatively weak interaction between two molecular fragments through the hydrogen bond (Fig. 4.2). On expanding the denominator in (4.2.15) in a series and integrating over the time, the spectral function of the high-frequency vibration near coh for p degenerate exchange modes takes the form S((o)=
\mG(a>), n KP-iy^O
(4.2.17)
"•
m-coh-pX-nv
92
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
CH>
OO
(K>«
CM)
Cy
(X?
j =4
cxA
d-o
Fig. 4.2. Low-frequency modes of librational vibrations of a molecular complex with a hydrogen bond.
Forp = 1, Eq. (4.2.17) reduces to the results obtained in Refs. 149 and 150 and in the high-temperature limit {T » hd).} gives, at rj«y, a simple formula'21154 for the strong one-sided temperature broadening of the spectral lines of complexes with hydrogen bonds (see Fig. 4.3): ^hr^tw^y^, \p-H\y\T
z = ^r{a>-coh). yT
(4.2.18)
Note that the molecules in gaseous phase do not exchange energy with the thermostat so that expression (4.2.17) corresponding to the limiting case rjj —• 0 is consistent with the concept of hot electrons (with the frequencies coh + X#§/(l-§)) which accounts for the fine structure of spectral lines of gaseous H-bond complexes.155 The one-sided broadening of the spectral line 2r is proportional to the
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
93
shift of its maximum Aco for various H-bond complexes in different phases (2T ~ KAO>, K ~ 1 -s- 1.5), both quantities increasing with the temperature quadratically at small T and linearly at large T, in accordance with the approximated temperature dependence coh + £#§/(!-§) o n different temperature intervals. For instance, tertbutanol manifests the characteristic dependence 2T ~ 1.32xl0'3 T2 (T is measured in cm"1) at 15 < T< 290 K (*•= 1.0 at T< 150 K and K* 1.4 at T> 150 K).156'157
S(,co)
c o
£
o (A
a
<
o
a>h
Fig. 4.3. One-sided temperature broadening of spectral lines typical of H-bond complexes.
Formula (4.2.18) represents an envelope of spectral function (4.2.17) at high temperatures (T > hcoi) when integration instead of summation over all nt is possible in Eq. (4.2.13). The typical parameter values in Eq. (4.2.18) for complexes with hydrogen bonds are at ~ 30 cm"', y~ 3 cm"1, and p = 4.155-157 The shift of the spectral function maximum specified by Eq. (4.2.18), Act) - |« m a x - (oh\ = (p - l)yT/tia)i, is proportional to the temperature; at T = 300 K, the value Aco ~ 60 cm"1 agrees with the experimental evidence. The proportionality factor K is defined as a difference of two roots of the equation x = 2^~p) ex~l : K = x2 - x, (*i<x2), and varies from 2.45 at p = 2 to 0.788 at p = 10, the calculated value K= 1.38 atp = 4 for terf-butanol closely coinciding with the measured one. At the temperature T = 290 K corresponding to the phase transition of tert-butanol from the liquid to the crystalline phase, the spectral line narrows jump-like by about 30 cm"1 156157 on account of the drastic enhancement of tj exchange in crystalline phase, this effect being accompanied with the accordingly
94
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
decreased shift Aco of the spectral function maximum. On adsorption of H-bond complexes, the fourfold degeneracy of low-frequency libration modes should be removed, as the substrate influences atomic displacements in adsorbed molecules unequally in two perpendicular directions. As a result, four pairwise-degenerate libration modes originate and the coefficient K grows from 1.38 to 2.45 together with the degeneration removal parameter.158 The exact Eq. (4.2.17) takes into account the effect of the reservoir (the condensed phase) on the spectral line shape through the parameter 77. Consideration of a concrete microscopic model of the valence-deformation vibrations makes it possible to estimate the basic parameters y and 77 of the theory and to introduce the exchange mode anharmonicity caused by a reorientation barrier of the deformation vibrations; thereby, one can fully take advantage of the GF representation in the form (4.2.11) which allows summation over a finite number of states. Indeed, torsional vibrations are strongly anharmonic; they are characterized by well-defined values of the energy barriers AU separating equivalent (with a rotation angle q> = 2TI) or nonequivalent (with AU stochastic reorientation processes dominate causing a spectral line broadening by the magnitude of the average reorientation frequency.159 The simplest way to take into account the anharmonicity of the torsional vibrations thus consists in considering a limited number of orientational states (with e < AU), which leads to the observed Arrhenius-type temperature dependence of the line width dictated by the factor exp(-ALVr). The pre-exponential factor depends on the relations between the parameters of the system which were established in the classical consideration.160 The problems discussed here are closely related to the problem of calculating the rates at which a particle leaves a potential well and which govern the rates of chemical reactions. The most consistent description of low-temperature chemical reactions that included tunnelling and dissipation processes was given in Ref. 161. We shall be interested only in the thermally activated contribution which dominates for many systems at not too low temperatures.
4.2.2. Valence-deformation vibrations of a molecular subsystem in condensed phase As a simple model which takes into account valence and deformation vibrations of a molecule imbedded in the condensed phase, we consider a diatomic molecule with two degrees of freedom corresponding to valence (in the radial variable r) and torsional (in the angular variable
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
A(mol)_A
h2
d2
h2
32
,
95
N
(4.2.19) 2
U(r,(p)»-m*o}l{r-r0) +Uar[h(r)+-AU(r\l-cosp
p = 1, 2, ...,
where m is the mass of the atom involved in angular deformations, m* is the reduced mass of the molecule, r0 is the equilibrium radial distance, and the approximate expression for the potential U(r,
Hr(p = ha>h\ bhbh+-\2 + H +-——'-g> 2 V J coh m h2
bhbh, (4.2.20)
*2
H(p=—" +-AC/(r 0 )(l-cosp < g), 2mr0 d
is the characteristic frequency of torsional vibrations.
In Fig. 4.4 we schematically show the eigenvalues sn
a
of the Hamiltonian
Hr(p which depend on the radial, nh, and the deformation, a, quantum numbers. For £0a «At/ the levels are grouped so that the gaps between groups are approximately the same and equal to hoi whereas the tunneling splitting which occurs forp > 1 in each group ofp levels is exponentially small in the parameter 4AU/ha>, . 6U21
96
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
Fig. 4.4. Transitions leading to a departure of a molecule from hindered-rotationsubbarrier states for n* = 0.
Neglecting the tunneling splitting we can assume the eigenstates |0cr) to be localized in the wells of the deformation potential so that e0cr = ha)i(a +1/2), a - 0, 1, ... «N + \&AU(rQ)/ha>i, (OCT' I cos = ( ^ m ^ ) ^ a^Sa.a_x + (a + l j K t f ^ ,
(4.2.21)
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
97
(Tunnel relaxation of orientational states in the phonon field of a substrate is considered in Appendix 2). When a molecule has a single equilibrium orientation (p = 1) the deformation potential is also characterized by a well-defined barrier AC/ which separates the equivalent minima. That is w h y , the subsystem Hamiltonian (4.2.12) used in the exchange dephasing model 1 4 7 , 1 4 8 with
^
r = 2mr0
coh
2mrQ al
(bl+bt)
(4.2.22)
correctly describes only the subbarrier deformation states (the operators ty and bf are n o w defined by the standard expressions for the matrix elements o f the subbarrier states only). W h e n w e describe the high-frequency response with regard to the molecular orientation, w e must put the operators A and B in the G F (4.2.1) as: A = bf,cos
B = bf, cosq>.
(4.2.23)
B y virtue of the properties of Eq. (4.2.21), the matrix elements Aq>q and B^
in
Eq. (4.2.11) will be diagonal in the quantum numbers a only for the subbarrier states, and the operator V takes the form (4.2.12) for those states. Indeed, it is shown in Ref. 164 (see also Ref. 6 1 ) that the interaction o f a reorienting molecule with a solid matrix (in which it is embedded due to the rigid coupling through only one of its atoms) can be written as the energy of the d'Alembert force, - w i i , in the noninertial frame of reference connected to the molecular center of mass which undergoes the acceleration ii due to vibrations of the solid: V = m r • i i . Expressing the deformation vector u in terms of the second quantization operators o f phonons o f the reservoir and using the fact that the operators bt and bi+ arise due to the presence of sm
( m(o 3 \/l k Zk=-
(4.2.24)
where p is the density of the medium and V is the volume of the main area. In the case of a Debye spectrum for phonons of the reservoir which are characterized by
98
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
the average sound velocity c or the Debye frequency atD and the mass M of the unit cell, the width of the resonant deformation mode takes the form164 3
_ mcoj _ 3/r m a, (Oi. Anpi? 2 M mD)
(4.2.25)
(For an alternative derivation of this formula, see Appendix 1). For real systems, we have m < M, (pi/copf « I , fij2mr^ooi«\ and the estimates (4.2.22) and (4.2.25) satisfy the conditions TJ, y« Q}t formulated for the Markov approximation. Transitions between superbarrier states {a > N, see Fig. 4.4) involve nondiagonal in a elements of the density matrix which give, at AU > T, a small contribution to the spectral function for frequencies of the order a, + £ A ( 0o- ~£0a')/h ( c * c r ' ) . As we are interested in the spectral function at frequencies near co^ , we can neglect this contribution. Thus, the required expression for the GF of the high-frequency mode taking into account the anharmonic coupling of the latter with the exchange deformation mode (characterized by a well-defined value of the reorientation barrier AU) takes the form (4.2.13) with restricted summation over the quantum numbers «/ = a = 0, 1, ... « N of the subbarrier states. It is then expedient to rewrite Eq. (4.2.14) in the following form: N 8 -r-g
(4.2.26)
v ^ / ^ + woOv + l ) ^ ,
(4.2.27) (4.2.28)
-fa +
tya'.a+l-V&a'.a-l}'
W
= 0, 1, .... N,
which corresponds to the Pauli equation with the transition rates perturbed relative to Waa'. The perturbations va determine the rate at which the molecule leaves the subbarrier states (see the area enclosed by the dashed lines in Fig. 4.4) and consist of two contributions: departures due to the anharmonic coupling of the valence and the torsional vibrations and those due to the strong anharmonicity of the torsional vibrations (the anharmonicity is determined by the magnitude of the barrier AU in
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
99
the present model). The latter contribution proportional to WQ(N + ])% was neglected in Ref. 151, since an equation like Eq. (4.2.26) with finite N was used only for a numerical approximation of the case N -><x>. The next section will cover elucidation of properties for the GF of the Pauli equation with unperturbed transition rates Waa' which satisfy the principle of detailed balance; on this basis, a perturbation theory in small corrections to the transition rates will be constructed, and the results will be considered in detail for a model with vCT and Waa< of the form (4.2.27) and (4.2.28).
4.2.3. Perturbation theory for the transition rates in the Pauli equation We consider a subsystem characterized by the states with the energies sa and the rates wqq' for transitions from a state q' into a state q that satisfy the principle of detailed balance. Pa
w
«"^-
ex
P{-£a/T)
*-2>q4/r)-
<4229)
The Pauli equation165 for the probability pq(t) of finding the subsystem in the state q at time t is conveniently written to perform further transformations:
The principle of detailed balance which is also valid for the quantities Wqq< enables the diagonalization of the nonsymmetric matrix fVqq- with nonnegative elements:
IX^V^^V,
(4.2.31)
where the eigenvalues p„ are also nonnegative and the orthonormalization relations for the matrix elements Cqvare defined with weights pq:
YJCqvCq'v = Pq8qq., J X ' C . / V ^ ' V -
(4-2.32)
100
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
If we denote the initial probabilities for state occupations at time /=0 by pq(0), the solution of Eq. (4.2.30) takes the following form:
A?M = Xp*'V(°)X C V C <7V ex P(-/V)-
(4.2.33)
The quantities introduced have a number of properties necessary for thermodynamic equilibrium to establish between the subsystem and the reservoir at t -> oo . First of all, we note that by virtue of the definition (4.2.30) the summation of the matrix elements W^ over the first index gives zero. Therefore, summation over q of both sides of Eq. (4.2.31) makes the product / ^ V Cqv vanish. From the linear independence of the rows Cq of the transformation it follows that there exists at least one eigenvalue /uv equal to zero. We denote the corresponding index v by zero, and then po = 0 and 2_, Cqv=0 for v * 0 . Summing the first of Eqs. (4.2.32) over q and using the fact that 2_. Pq - ' /i0=0,
CqQ = p q ,
we
f"ind
J^Cqv=Sv0.
(4.2.34)
q
These properties lead to the physically obvious consequences of the solution of Eq. (4.2.33):
9
q
pq(t) = pq forp^O) = pq, andpq(<x>)= pq for arbitrary initial conditions. The frequency Fourier component gfl{co) of the GF of the unperturbed Pauli equation (4.2.30) (satisfying Eq. (4.2.30) with -S(t)5qq' has, in view of Eq. (4.2.32), a pole at
w
^—,0)
+ lUv
CO
™
on the right-hand side)
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
101
Here £qq'(a>) is determined by the general expression for g^\{co) in which the term with //„ = 0 is excluded from the sum over v, so that, for instance, ^ - ( o ) , with the weight factor p taken into account, is a pseudoinverse matrix with respect to Wq(i and satisfies the following identities:
YjVm"eq"M = Sm'-P^
2>w'(°) = °-
i"
?
(4-2.36)
We now consider perturbations of the transition rates, adding the diagonal in q contribution vq8qq> to W^. The perturbed GF gqq'{co) will then be related with the unperturbed one through the Dyson equation146
In the second-order perturbation theory in vq, the required retarded GF (25) is equal to G{t)=-i0(t)exp(-Tt),
f ^ v ^ - ^ v ^ o j y / v
(4.2.38)
and determines, accurate to the factor -id(t) , the probability that the subsystem leaves the given group of states (e.g., those enclosed by the dashed line in Fig. 4.4). The quantity F acquires the meaning of a generalized leaving rate, since vq and F can take on complex values. The spectral function corresponding to Eqs. (4.2.13) and (4.2.26) becomes Lorentzian: S{co) = --\m n
=•. co-co/j+ir
(4.2.39)
The first term in expression (4.2.38) for F has a simple physical meaning: it sums the perturbed leaving rates from each level of the subsystem taking into account the equilibrium probabilities for their occupation. On the other hand, the second term depends on the unperturbed rates for transitions between states of the subsystem and is inversely proportional to them by virtue of the definition of
102
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
We evaluate the pseudoinverse matrix ^ ' ( o ) of the transition rates (4.2.28) between N+\ low-energy states of a harmonic oscillator which interacts resonantly with a phonon reservoir. With this aim in view, it is necessary to: 1) write g^ '.{co) as the ratio of the appropriate cofactor to the determinant of the matrix icoSqq' - Wqq' ; 2) expand this quantity in a Laurent series in co; then the principal part of the Laurent series gives the first term of the right-hand side of Eq. (4.2.35) and the terms of zeroth degree in co from the regular part of the Laurent series determine the required £qq'(o) matrix; 3) calculate resulting determinants and their derivatives with respect to co for co = 0 using relatively simple recurrence relations obtained by expanding determinants of a quasi-triangular shape (like the matrix (4.2.28) for fFqq) with respect to an arbitrary row or column. As a result, we get
eaa<0h—ea-a(0)
=
yr*-i, y \-^~k * =1
1
k
*=?+1
k
(4.2.40)
S k=\
Here q,q' = 0, 1,..., N and it is implied for q = q' = 0 and q = N that sums vanish if the upper limit of summation smaller than the lower one. It is easy to check that the expression (4.2.40) derived here obeys the identities (4.2.36). As far as our concern is with the second-order perturbation theory specified by Eqs. (4.2.38) and (4.2.39) and valid for y « ?j and
where
2
e
r = Rer = (N + lh-^^AAf)+~j^-yBN({),
(4.2.41)
Aa> = hnr = r-£-CN{<S),
(4.2.42)
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
1 (" + ^N+]
A (A
BN(Shl-
r (A
{N + l)(\-Z)22{shl 1"£ N+\
r (A
Mz)-
CN<Shl-{N]l^$"[2AN®-l DN{$) = -
l-Z
N+\
N-
v i z d
w +
103
•DN(Z)
(4.2.43)
DM,
r(i-^l
For a single subbarrier level we have A0(£) = 1, 2?0(£) = C 0 (# = 0, and the halfwidth of the spectral function is determined solely by the reorientation rate, which is equal to the transition rate to the first excited state of deformation vibrations:164 r = 7 n{a),),
n{a),) = [exP(hco,/T)-1]-1.
(4.2.44)
For two subbarrier levels (N = 1), Eqs. (4.2.43) represent, at £ « 1, an approximated result of the two-level problem that holds true for any values of £,:
^ ) = - M i ^ ) l + 3 £_(l + 6 £ + £ 2 ) K l 4
*>(#) =
(W) 2 (l + 6£ +
(4.2.45)
C,fe) =
2^2
i-
l +
2£
l + 6£ + £ 2l>2
Finally, in the other limiting case, N -> oo, the functions 5/v(£) and CN(£) are approximately equal to unity, and Eqs. (4.2.41) and (4.2.42) reduce to the results of Ref. 151 for y« r/. We draw attention to the fact that in the first-order perturbation theory we have AfJ^g) = 1 and Bfl%) = 0 in Eq. (4.2.41) for any N, and the expression for the leaving rate of the molecule from the subbarrier states reduces to two well-known special cases. The first of these corresponds to the low-temperature limit £ -> 0 for which166 77 = TVo(l-
r = (^ + lKexp(-A£//r).
104
4.2. Markov approximation for Green's functions of molecular subsystems in the condensed phase
The second special case corresponds to the classical limit l - £ -> hcoi/T -> 0 and gives the Kramers' result:160
T = w0(AU/T)exp(-AU/T), in which the parameter w0 serves as a "viscosity" coefficient. In the second-order perturbation theory, the coefficient Au(%) refines the first-order result. For N» 1 and £ < 0.5 , we derive the asymptotic expression from Eq.(4.2.43):
4v(£)«i-#/(i-£). The main results of the present section which are helpful in describing spectra of high-frequency local vibrations of molecular subsystems in the condensed phase are contained in Eqs. (4.2.17), (4.2.18), and (4.2.41) to (4.2.43) which involve two parameters, yand TJ. According to the estimates of Ref. 155, molecular complexes with hydrogen bonds are characterized by the anharmonic coupling coefficient y~ 3 cm'1 and the frequencies of the libration vibrations cot ~ 30 cm"1 much less than the Debye frequency coD,. This fact leads to the inequality TJ « ythus corroborating the validity of Eq. (4.2.18), in agreement with the experimental data of Refs. 156 and 157 (a discussion of this problem is given in Refs. 121 and 154). Estimation of the parameters y and t] for surface groups of atoms using Eqs. (4.2.22) and (4.2.25) or similar relations in Ref. 151 shows that y/rj ~ 0.1. If we put ylt] = 0 . 1 in (4.2.41), the reorientation contribution to the broadening becomes dominating over the anharmonic contribution with the coefficient y, starting at temperatures T > 0.5 ha>i for N= 3 or T > 0.9 hcot for N = 5. Assuming that the values of the barriers AC/ for CO bridge groups on Ni(l 11) correspond to N < 3 and involving the realistic estimates co/ «184 cm'1, coh «1900 cm"1, y~ 1 cm"1, 77 ~ 30 cm"1, the reorientation contribution to Eq. (4.2.41) can explain the previously observed151 spectral line broadening for the valence CO vibrations. Apart from the contribution (4.2.42) from the fourth-degree anharmonicity (r-r 0 )V> e v e n m e third-degree anharmonicity of the form {r-r0)q? and also other low-frequency vibrational modes of the substrate167 will give a comparable contribution to the shift of the maximum of the same line. Neglecting the other low-frequency modes and taking into account only a single harmonic (N -» 00) mode, one would overestimate, by an order of magnitude, the value of the only anharmonic parameter y in the description of the observed spectra.150151 The purely reorientational broadening mechanism with a single threefold quasidegenerate subbarrier level is characteristic of the valence vibration spectral line for OH groups on Si0 2 surface. Equation (4.2.22) describes the observed temperature
4.3. Generalization of the exchange dephasing model
105
dependence of the line halfwidth for co, « 200 cm"1 and 77 w 4 cm"1 (for this system, y a 4 cm"1 holds and in Eq. (4.2.22) we have flrj» 0.2 cm'1, which is much less than the value 7 ). 61121 - 164 Eqs. (4.2.41), (4.2.42), and (4.2.45) are applicable to the system of OD groups on Si0 2 surface, since in this case one can put the number of subbarrier levels equal to two due to the doubled mass of the reorienting atom. Importantly, the value of the results gained in the present section is not limited to the application to actual systems. Eq. (4.2.11) for the GF in the Markov approximation and the development of the perturbation theory for the Pauli equation which describes many physical systems satisfactorily have a rather general character. An effective use of the approaches proposed could be exemplified by tackling the problem on the rates of transitions of a particle between locally bound subsystems. The description of the spectrum of the latter considered in Ref. 135 by means of quantum-mechanical GF can easily be reformulated in terms of the GF of the Pauli equation.
4.3. Generalization of the exchange dephasing model to various cases of anharmonic coupling between high-frequency and low-frequency modes, and to the case of collectivized excitations in adsorbate In Sec. 4.2 the exchange dephasing model was considered which takes into account only the biquadratic anharmonic coupling between the local and low-frequency vibrations (see Eq.(4.2.12). The need for due regard to cubic terms of anharmonic coupling was pointed out more than once.1168"169 First, the contribution of the cubic anharmonicity to relaxation processes is of particular significance in the cases when the doubled frequency of a resonance molecular mode is found close to the frequency of the local vibration, as for instance, in the system D/C(l ll).' 7 0 1 7 ' Second, the cubic anharmonicity also contributes to the dephasing mechanism of line broadening and results in the renormalized constant of the biquadratic anharmonicity.1 Such a renormalization of the quartic anharmonicity coefficient in terms of the cubic one was invoked by Ivanov, Krivogiaz et al. for consideration of high-frequency vibrations in crystals as early as in the sixties.172'73 The necessity of taking into account cubic corrections in the processes involving four vibrational excitations is easy to comprehend from the analogy with an anharmonic oscillator treated in the perturbation theory. It is common knowledge that contributions from the quartic anharmonicity in the first-order perturbation theory and from the cubic anharmonicity in the second-order perturbation theory are of the same order of magnitude. Vibrational dephasing and relaxation in four-phonon processes are determined by the squared quartic anharmonicity coefficient. Therefore, the fourthorder perturbation theory is needed to adequately describe these processes.
106
4.3. Generalization of the exchange dephasing model
Intermolecular lateral interactions and resulting collectivized vibrations of individual adsorbed molecules greatly add to the complexity of description for local vibrational excitations in adsorbates. Fig. 4.5 schematically demonstrates that these interactions on a simple planar lattice of adsorbed molecules which vibrate with high (coh) and low (a)/) frequencies lead to the emergence of the corresponding energy bands, with energy levels classified by the wave vector K. Isolated molecule:
Molecular ensemble:
t^ p
l" ^ ^s^
^
^
-" ^-"^ s^~ ^
S*
•^
^
t-
^s-
y S '
^
^
Lateral interactions
=> <2>S(K, V)
<» S (K, V)
Substrate phonons
Substrate phonons
N;(K)
Fig. 4.5. Collectivization of vibrational excitations in adsorbed molecular ensemble due to intermolecular lateral interactions.
Considering only biquadratic anharmonic coupling, the dephasing of local vibrations was treated in the special case that only high-frequency modes underwent collectivization174 and subsequently with the allowance also made for collectivized low-frequency modes.138175 It should be emphasized that the possibility for
4.3. Generalization of the exchange dephasing model
107
orientational phase transitions to occur in adsorbed molecular systems attaches special significance just to due regard for low-frequency mode collectivization, since dispersion laws for these modes are particularly sensitive to phase transition parameters. Nonetheless, these subtle effects, if included alone, without various anharmonic contributions, can hardly furnish the picture accounting for experimental evidence. The objective of the present section is to develop a consistent theory properly taking into account both lateral interactions of highfrequency and low-frequency modes of adsorbed molecules and all possible cubic and quartic anharmonic couplings between them. Consider an arbitrary two-dimensional Bravais lattice, with its sites R occupied by adsorbed molecules and molecular vibrations representing two modes, of a high and low frequency. Frequencies o\h reduced masses mhj, vibrational coordinates w/,,/(R)> a n ^ momenta p/,,i(R) a r e accordingly labeled by subscripts h and / referring to the high-frequency and the low-frequency vibration. The most general form of the Hamiltonian appears as140 # (mol) = tfjmol)
+
^(mol)
+
^(mol) >
(4 3
,}
*C° = Z T ^ 4 X [ V * V R R ' +*Wte(»-R,)]«*1/(RK/(R') R 1
Z
RR'
A;
( 4 3 2 )
KL
J
HJff> = V3 + V4, K
3 = £ [d)3o^(R)+<S2lMA2(R)w/(R)+,2^(RK2(R)+«>o3«MR)] •
(4.3.3) <4-3-4>
R v4 = ^ [ C D 4 O ^ ( R ) + < I > 3 1 ^ ( R ) M / ( R ) + C I > 2 2 ^ ( R ) W / 2 ( R )
R
(4.3.5) 3
4
+ CD13^(R)M/ (R)+(D04M (R)].
The second equality in Eq. (4.3.2) demonstrates that the harmonic component of the Hamiltonian of the molecular subsystem is diagonalized by the Fourier transform in terms of wave vectors K of vibrational excitations:
4.3. Generalization of the exchange dephasing model
108
/2
(4.3.6)
R
/v(K)=kW EsA./(K>* . where No is the number of adsorbate lattice sites in the main area. These collectivized excitations are characterized by the dispersion laws: (4.3.7)
^,/,lat(K) = J > , , , / , , a , ( R K ' K R .
(4.3.8)
In the particular case of dipole-dipole lateral interactions between molecules with the same dynamic dipole moment fi, we have:
r ~ \ r
DaP{R),
(4.3.9)
A a,P~x,y,z\ hJj
D°0(R) =
S
afi
RgRp
iRl3
iRl5
(4.3.10)
(the Fourier components of the tensor Dafi(R) for various planar lattices were analyzed in detail in Sec. 2.2 and 3.2). Anharmonic coupling of high-frequency and low-frequency modes on the same lattice site is presented by relations (4.3.3)(4.3.5). The Hamiltonian of molecular vibrations (4.3.1), along with the Hamiltonian of substrate phonons (4.1.4) (which can also be represented as a sum over K and v variables) and the operator of harmonic coupling between substrate phonons and low-frequency molecular modes (4.1.8), constitutes the full Hamiltonian of excitations in the adsorbate and the substrate. It represents a generalization for the above-considered Hamiltonian (4.2.12) in the following three respects. First, highfrequency and low-frequency molecular modes are collectivized as a result of lateral interactions in the adsorbate. Second, anharmonic coupling between high-frequency and low-frequency modes does not reduce to a biquadratic term and includes all possible cubic and quartic anharmonic contributions. Third, the generalized Hamiltonian is free of assumptions implied in the Heitler-London approximation. It
4.3. Generalization of the exchange dephasing model
109
is clear that with these generalizations, handling the problem exactly is no longer possible, but in practice it is often sufficient to apply the results provided by the fourth-order perturbation theory for two-time retarded Green's functions in the coordinate-momentum representation. Due to the occurrence of various types of anharmonic coupling and energy bands of molecular vibrations, the energy conservation law admits various processes contributing to the spectral line broadening for local vibrations (Table 4.1).
Table 4.1. Various processes contributing to the spectral line broadening for local vibrations. Frequencies of collectivized local vibrations QK (solid arrows) are supposed to exceed phonon frequencies coKq (dashed arrows): QK > max o>Kq. For an extremely narrow band of local vibrations, diagrams A and B respectively refer to relaxation and dephasing processes, whereas diagrams C account for the case realizable only at the nonzero band width for local vibrations. The number of concerted excitations
A
B
3
C
—
1 --«.
--'
2
1 "-».
--'
2
4 **
~~
~~~
• * .
As the detailed mathematical description of these processes is rather tedious,140 here we confine ourselves to the temperature-dependent dephasing and three-particle relaxation (processes of types B and 3-A in Table 4.1) contributing to the shift AO(^K) and width r B (Q K ) of the spectral line for a local vibration Q K at Q K =coh{K)»coK ~T:
A0(QK) =
mhQKN0
-E
g)(2ef(K,K,K,) / mea>K
\
(4.3.11)
110
4.3. Generalization of the exchange dephasing model
hf(K.KK2)f
7th"
rB(nK)=
X
+G)K
- « K - K ,-K 2\lK2
, + ?
U
U
?K-K,-K 2 )>
(4.3.12)
r A (" K )=
, 5R(fiK - *>K, - ^ K - K . ^ K , +1K
7ih<&\2
4mhm,nKN0
•KJ (4.3.13)
^K.^K-K,
K
x
+
[4%) 4°K-K l )+ll
where the renormalized anharmonic coefficient can be represented as (Dg i) (K,K 1 ,K 2 )=a> 22 +3
3flK+QKi
^30^12 m
h
-^K_K|
[ " K - ( % , +"K-K,)2]["K-("K, -^K-K,)2]
3 Q K + Q K , -^K-K, + 3 021*03 m t [^K-(^K, +^K-K,r][^K-(^K, -«K-K,) ]
tii+al 'K
*21 K,=K 2 ,
mh
*22
[Q2K -{coK.
-til 'K-K,
+ f i K - K / ) 2 ] [ " 2 < -(VKj
Q
(4.3.14)
-^K-K,)2]
K+4,-4-K,
M
< [QK-(<% +«K-KJ)2]["K-(<»K, -<%-K,)2]
It is noteworthy that relation (4.3.12), with factors (2^/77 )$R(a>; 77) substituted for 9?(#;7), accounts for additional spectral line shifts Ai(QK) for local vibrations which arise from the corresponding process. However, this contribution does not comprise the total line shift in this high-order perturbation theory just as contribution (4.3.11) does not provide total shift in the second-order perturbation theory. It is expedient first to analyze the contribution from cubic anharmonicity in the simplest case, i.e. for a system free of lateral interactions, when QK and COR are
4.3. Generalization of the exchange dephasing model
111
independent of K and amount respectively to Q 0 = coh and co^ = coh With regard for the inequalities TJ0« COOrelations (4.3.11)-(4.3.13) yield: (4.3.15)
rB(n0)
=
L(eff)f X^—Ln(co0ln{a)0)+\], no
M>? 2
rA("o) =
2
rP =m m/Q co
2
h
0
0
[
"W+T
m
2}(n0-2o)0Y+7ji
(4.3.17) '
2*21
mhQ.l
Q 0 - 2»i o ( D 0 - 2 « o ) 2 + ' 7 o
•,021*03 , 2
(4.3.16)
mA(4n§-a^)
Q
(4.3.18)
0+2«0
The theory developed permits spectral line shift and width to be calculated from Taylor power series for interatomic potential energies in a concrete system. Various methods of tackling this problem can be found in the literature140'169171176' 180 (see also survey 181 and references cited therein). Here we invoke a realistic model for the coupling of two mutually perpendicular vibrations which was reported by Burke, Langreth, Persson, and Zhang.1 As in Ref. 1, write the Hamiltonian for the interaction between the modes uh and Ui in polar coordinates r and 6, where 6 is the angle between the adsorbate bond and the perpendicular to the surface plane: tf(mol)
=
Pr ,
,
2m
Pi 2m(r0+ury-
2
+\kee
2
+Ueui+^ul+... 2
(4.3.19)
(r0 is the equilibrium length of the adsorbate-surface bond). Expanding the factor (r0+ur)'2 in ur and taking into account that p$j mr0 has the same amplitude as ks62, we are led to the following values of the parameters involved:
4.3. Generalization of the exchange dephasing model
112
(4.3.20) <J>22 = 3 O T ® 0 I2r0 •
*21 = ° 0 3 = 0. *12 =-ma)o/r0,
(In contrast to Eq. (3.13) of Ref. 1 where ur(r0+ur)sm9, we assume that the force constant kgof the frustrated rotation with the changing angle #is independent of ur). On substituting values (4.3.20) in Eq. (4.3.18) and accounting for the doubled contribution from two modesfflomutually perpendicular in the surface plane, we obtain: (eff) _ 3hco0
(4.3.21)
mrQQ0 S = -(o0
Ky
Q 0 - 2a>,'0
(Q0-2(o0Y+7]2
r?W =wr ^Q 2
0
«(
0
(4.3.22)
^O + 2 » 0
(4.3.23)
2j(n0-2aj0f+rj2
Evidently, we would arrive at the same result if the problem were considered in the Cartesian system of coordinates, with accordingly different values of anharmonic coupling coefficients:
-l^o)/^
(4-3-24)
containing contributions from central forces (proportional to QQ an£ l ®i)- However, it follows from the structure of the relationship (4.3.18) that these contributions to yfi ' should be completely compensated, just as in the limit Q 0 » a>o treated in Ref. 1. To derive Eq. (4.3.23) in Cartesian coordinates, we should invoke the energy conservation
law, Q0=2fflb,
before
using
Sokhotskii's
formula,
(x + iO)
= P\x~ j-i7rS(x), which introduces vibration relaxation in the system concerned. Thus the model specified by Eq. (4.3.19) serves for the verification of relations (4.3.21)-(4.3.23) gained in the high-order perturbation theory. For the case of CO/Pt(lll), Persson and Ryberg176 regarded a parallel frustrated CO translation as a low-frequency parallel mode and derived the estimate
4.3. Generalization of the exchange dephasing model
113
(?ers) r
=-tico0Q0/4Eh
(4.3.25)
which was also in nice accordance with experimental data for the systems H/Si(l 11)182 and H/C(l 11).171 To arrive at Eq. (4.3.25), they in fact introduced a new Morse potential which accounted for potential energies of two mutually perpendicular displacements uh and uh with the independence included only in the Eh parameter, so that the free term in the expansion of this potential in uh specified the harmonic potential energy /nft>2H2/2 of the vibration ut:m
£(
p
-)(M„Mf) = [ £ , - I w f t , 0 2 M f j ( e - 2 ^ - 2 , - ^ ) ,
« =j ^ -
(4-3.26)
(Here Eh and a are the parameters of a normal Morse potential, and m = mh = mi in the framework of the model discussed). Although this combined potential (4.3.26) provides plausible estimates, it can hardly be substantiated in terms of the theory of intermolecular interactions and contains, in addition, only biquadratic anharmonic i i
i i
coupling -(mco0a /2)uhue (leading to Eq. (4.3.25)); thus, it cannot be involved as an intermolecular potential in the approach defined by relations (4.3.21)-(4.3.23) which also takes account of the cubic anharmonic coupling. Of interest is to compare Eqs. (4.3.21) and (4.3.25); for this purpose, express kr and 3>3 in terms of the parameters Eh and a of the normal Morse potential and introduce the ratio of the equilibrium adsorbate-surface bond length to a']: K- arQ. Then B=-y{Vm)Q.<J i? a*, 2Q>3ro/kr = -K and we are led to (efl)
=
6fr-
-^(Pers)
(4
3
2?)
The parameter values K and 8 for light adatoms are listed in Table 4.2. The proportionality factor for ^ e f l ) and / P e r s ) (see Eq. (4.3.27)) proves to be of the order of unity, which makes our estimates close to those by Persson and Ryberg.176 Thus, it is hardly surprising that the results of the approach presented in Eqs. (4.3.21)-(4.3.23) are consistent with the experimental evidence. The specificity of this model becomes noticeable if the value Q 0 is close to 2ft^ and hence the resonance factor 8 grows large enough. For the system H/C(l 11), positive values of 8 somewhat decrease the absolute value of ^ e f f ) , while the system D/C(lll) is characterized by 8
4.3. Generalization of the exchange dephasing model
114
absolute value of the parameter 8 increases, so does the contribution of the relaxation process T^ ' which becomes predominant for the systems H(D)/C(111).
Table 4.2. The basic parameters of the models under discussion for some adsorption systems. The dephasing ( T T ^ )
and the relaxation (2r^ 3 ) )
contributions to the full spectral linewidth for local vibrations as well as the experimentally observed values of this parameter (2r* exp *) are presented for the temperature 7*=300 K. Data are taken from a, Ref. 176; b, Ref. 169.
Parameter n 0 (cm"') Eh (eV)
H/Si(lll) 2086
H/C(lll) 2835
D/C(lll) 2110
3.5
3.5
3.5
210
1140
1140
2.21
2.17
2.20
52
120
120
0.014 -3.88
0.579 -28.6
-1.58 -21.3
-5.7
-21.6
-73.3
a
-5
-23"
-
1
2 l f > (cm' )
1.13
0.03
0.38
2r<,3) (cm"')
0.01
5.19
28.0
5.7"
30"
©o ( cm "') K
rj (cm -1 )
S Pers > r< r<*>
rtp)
(cm"') (cm"1)
(cm"1)
a
2r ( e x p ) (cm"')
l
4.3.1. Contribution of dipolar dispersion laws to dephasing of high-frequency collective vibrations Here we focus on the effect of dipolar dispersion laws for high-frequency collective vibrations on the shift and width of their spectral line, with surface molecules inclined at an arbitrary angle 6 to the surface-normal direction. For definiteness, we consider the case of a triangular lattice and the ferroelectric ordering of dipole moments inherent in this lattice type.56109 Lateral interactions of dynamic dipole moments \i - / / e (e = (sin&os^, smfcirup, cos#)) corresponding to collective vibrations on a simple two-dimensional lattice of adsorbed molecules cause these vibrations to collectivize in accordance with the dispersion law:121
4.3. Generalization of the exchange dephasing model
fa
1
nl=n2h+m a
h
a,/3=x,y,z
115
\ f dup^
duh)
ba%).
du'h t Jo
(4.3.28)
Here
baP{k)^DaP{R)e-
/k R
(4.3.29)
is the Fourier component of the tensor of dipole-dipole interactions (4.3.10), R is measured in units of the lattice constant a; mh and uh are the reduced mass and the displacement of the vibration with the frequency Qh. Since the spectral lines observed in the IR region are determined by the value k=0 at which the tensor DaP(6) - D(o)Sap is isotropic for symmetric lattices, the corresponding spectral shift caused by dipole-dipole interactions can be written as:
AQo=Q 0 -Q,,«Q d i p 5(o)l 2
5(0)=-(3COS 6>-I)D0,
Q dip=-!\Qh, 2a'
zv = mh
(4.3.30)
Z>0 «11.034,
where X\ i s the vibrational polarizability of the molecule. Thus, the positive and negative shifts result at G < 54.7° and 9 > 54.7°, respectively. The approximate equality in Eq. (4.3.30) holds at |AQ 0 |«Q/„ which is mostly the case for real systems. With the electronic polarizability taken into account, the value Qdip is renormalized by the factor (1 + j e 5 ( o ) / a 3 ) - 1 . ' 8 3 The parameter Qdjp can also be influenced by image dipole effects which are sometimes quite noticeable even on insulator surfaces. An additional shift and broadening of the spectral line for high-frequency collective vibrations in the framework of the widespread dephasing model 147148 ' 51 are describable in terms of the anharmonic coupling of these vibrations with a lowfrequency resonance mode a>i characterized by finite lifetimes 1=2/77 (l ' s the full resonance width). With the anharmonic coupling specified by the function ft, the maximum position Qmax and the width T of the spectral line at |ft| « 7+|AQ0| are defined by the following relationships:
4.3. Generalization of the exchange dephasing model
116 Q
max= Q 0+ro/ 2 + A . A = A0 + A,, = n(al\r{(ol)+1}—
£-
A0=y0n(co,)
/k
1
(4.3.31)
(4.3.32)
where JV is the number of adsorbate lattice sites in the main area, and n(co) = [exp(ha)/kBT)-1]~
(4.3.33)
is the temperature-dependent factor of the Bose-Einstein statistics. Persson, Hoffman, and Ryberg were the first to derive an expression like Eq. (4.3.32) which involved the coefficient y independent of the wave vector k and concerned the case of biquadratic anharmonic coupling between the collective and the resonance mode.174 For the same type of the anharmonic coupling considered in the lowtemperature limit (kBT«ha)i, y, TJ, and AQ0 are arbitrary), the following expression is valid:184 f\\
r
v;
Re W, W = yn(; f°l -Im
N4-*n -Qk+in k 0
(4.3.34)
(This formula is derived in Appendix 3). With regard to various cubic and quartic anharmonic interactions, the quantity ft is characterized by a certain combination of these anharmonic contributions and becomes dependent on k (see Eq. (4.3.14) for a related quantity and Ref. 140). However, this dependence is insignificant compared to the k-dependence appearing in the denominators of Eqs. (4.3.32) and (4.3.34). Therefore, spectral characteristics defined by formulae (4.3.32) can with good reason be regarded as proportional to certain functions of lateral interaction parameters and of the resonance width TJ:
V^rv where
Re F, -Im
1
ph(Q)dQ.
F.±J — f JV^Qo-^k k
-Q + irj
(4.3.35)
4.3. Generalization of the exchange dephasing model
117
/>/»(") = ^ I > ( Q - Q k )
(4.3.36)
is the spectral density function for collective vibrations. On the other hand, the same functions enter in the low-temperature approximation (see Eq. (4.3.34)): ynjcoi)
(\-rFAY+r2F?
yFv
(4.3.37)
For functions (4.3.35) to be calculated, rather complicated specific dispersion laws should be known for the excitations induced by long-range anisotropic dipole forces. That is why, the approximate expression F(apPr)=(AQo+/?7)-
(4.3.38)
is frequently used in estimations. Here all characteristics of lateral interactions are accounted for by a single parameter AQ0 defined in Eq. (4.3.30). Substituting Eq. (4.3.38) in (4.3.37), we arrive at the known low-temperature formula of Erley and Persson: ^(appr) -(appr)
2
"fo)
AQ0-y
{AQo-yf + n
2
(4.3.39)
In what follows we will establish under which conditions and to which accuracy the approximation (4.3.38) is justified for the cases of normal, parallel, and inclined molecular orientations with reference to the surface plane.186 The curves 1 in Figs. 4.6a and b show the functions Fr and F& calculated by formulae (4.3.35) and (4.3.38) for the case of normal molecular orientations (e || Oz) and plotted versus the argument AQ/(^+AQ). The dimensionless argument and functions of this kind normalized with respect to the sum of the resonance and the band widths were introduced so as to depict their behavior in both limiting cases, AQ.«rj and AQ.»TJ. The deviation of the solid lines from the dotted ones indicates to which degree the one-parameter approximation defined by Eq. (4.3.38) differs from the realistic dispersion law. As seen, this approximation shows excellent adequacy, but for the region A Q » ^ , where the asymptotic behavior of the approximation (4.3.38) and Eq. (4.3.35) are as follows:
4.3. Generalization of the exchange dephasing model
118
.2 ^
^r(appr) ,
1.2 1-1.44-2AQ A£T
tt—Uzbrt-* AQ
"
4AQ
F ( a p p r ) « 1 44
^
(4.3.40)
ACT
,Fr*-In 2AQ2
4eAQ
(4.3.41)
As the behavior of the spectral line width, 77^(1/77), can roughly be regarded as linear in 77 and the coefficient value 21n2 * 1.386 is close to 1.2, the approximation (4.3.38) holds workable for normal orientations even at AQ»T].
0.5
An/(Ti+4fi)
0.5
Fig. 4.6. Dependences of dimensionless spectral characteristics Fr(rp-AQ) (a) and F^TT+AQ) (b) (as defined in Eq. (4.3.35)) versus the dimensionless parameter ACV(ip-An) at 0= 0, Afi = 13.2410^ (1); 6= 90", AO = 11.586ndlp (2); 6= 54.7", AfJ= 6.080Qdip (3). Solid and dotted curves correspond to Eq. (4.3.35) and to the approximation specified by Eq (4.3.38).
4.3. Generalization of the exchange dephasing model
119
The case of parallel orientations (e 1 Oz) differs radically from the previously considered one, since the frequency dependence of the spectral density function is specified by the fractional power law:109
Ph (Q) =
{ZjlnT
r W V f "dip AK
Q>Q0,
^ | = - r = « 7.255,
Q-Q0 ' Q dip
(4.3.42)
yLa 0.263
(Here T(l/4) « 3.626 is the special value of the gamma function so that the numerical value of the coefficient is 0.0531). On substituting Eq. (4.3.42) in (4.3.35), we obtain, at TJ « A Q , the following relationships:
K
^ S
1/4
(3/2^y/2
r2(i/4W/2rf "diP
Q dip
(4.3.43)
With 7] « A Q , the distinction between solid and dotted lines 2 in Fig. 4.5 becomes significant suggesting that the approximation defined in Eq. (4.3.38) is inapplicable to the case of parallel orientations. The angular dependences F r AQ and FAAQ at 7->0 calculated for inclined molecular orientations are represented by the curves 1 and 2 in Fig. 4.7. The spectral density function (appearing as FTln at ^->0) takes nonzero values if the angle Granges from 46.7° to 90°, since in this 6 range the value Q 0 is neither the upper (as at 0<46.7°) nor the lower (#=90°) edge of the collective vibration band. The maximum values Fr found in the vicinity of the angles 6 =49° and 0 =63° correspond to the regions of dispersion laws with small values |5Qk/3k|. The asymptotically exact result with a nonzero spectral density function can be obtained for molecules which are only slightly inclined to the surface plane: /~ \ -v3 cos# ph{Q0) = —T———, cos0 « 1 . 2xy± "dip
(4.3.44)
For the value FA to be calculated, the integration over the whole Brillouin zone should be performed which yields, in the limiting case of #=90°, the value F A « 5.44/Qdjp (in the framework of the isotropic approximation).
4.3. Generalization of the exchange dephasing model
120
Fig. 4.7. Angular dependences of dimensionless spectral characteristics /-VACi (curve 1) and FAAQ (curve 2) calculated at rj= 0.
As a main point of this subsection,186 it is shown that due to sufficiently strong lateral interactions of adsorbed molecules on a two-dimensional triangular lattice, the spectral line for collective vibrations manifests some characteristic peculiar relationships between its dephasing-induced broadening and the resonance width 77 for the low-frequency mode: The line width changes as 77^(1/77) for surface-normal and as TJV4 for surface-parallel molecular orientations, and takes nonzero values (independent of 7) for inclined molecules with the inclination angle ranging from 47° to 90°.
4.3.2. A simple model for collective high-frequency and low-frequency molecular modes In the framework of the low-temperature dephasing model, the dependence of spectral line shifts and widths for local vibrations on real dispersion laws for highfrequency (Qk) and low-frequency (cok) molecular modes is defined by the complex function (see Appendix 3):184
n Wk =
r
1
M
^of -^)'
(4.3.45)
4.3. Generalization of the exchange dephasing model
121
^ 1 ) = — X K -"k+k 2 -k, + ">k, ~«k 2 +'(7k, +1k2)/2ll 0
•
(4-3.46)
k2
Here k, kh k2 are dimensionless wave vectors of vibrational excitations in a lattice of adsorbed molecules (with N0 molecules in the main area). Formulae of Erley and Persson185 (4.3.38) and (4.3.39) follow from general relationships (4.3.45) and (4.3.46) with k = 0 provided the dispersion of lowfrequency modes, o\ and rjk, is neglected and the differences Q k - £ \ + | < _[< are substituted by the corresponding values of spectral shifts AQk. To estimate the effect caused by the dispersion of low-frequency molecular modes on the local vibration spectrum, we substitute the difference o\ -o\ in Eq. (4.3.46) by the parameter Aa\, the effective halfwidth of the band of low-frequency vibrations corresponding to the high-frequency spectral line with the wave vector k. With regard to the sign change resulting from the permutation of the summation variables k, and k2 in this difference, we arrive at the following generalization52187 of the Erley-Persson formulae:
*-f^.
*°,
^ T
2-
(4-3.47)
Owing to the additional parameter Ao^ that enters into the above expression, the sign of ReFk is opposite to that of AQ k at Aa)k 2 > AQ k 2 + rf. This fact proves to be fundamentally important in the analysis of temperature dependences of the spectral line shifts for the system CO/NaCl(100). Consider a 2x1 phase of the monolayer of CO molecules forming a square lattice on the NaCl(lOO) surface. The orientational inequivalence of two CO molecules in the unit cell of the two-dimensional lattice (see Fig. 2.3) results in Davydov-split spectral lines for radial C-O vibrations with the frequencies Q s = 2154.86 and fiA = 2148.58 cm'1 at T = 11 K (see Fig. 2.5) that correspond to symmetric surface-normal (S) and antisymmetric surface-parallel (A) normal vibrations." The dispersion law for these vibrations, if calculated with regard to geometrical features of the system (molecular orientations inclined at an angle of 25° to the surface normal with alternating azimuthal angles, = n, along any of the square lattice axes), reproduces the observed magnitude of the Davydov splitting, Q s - QA « 6.3 cm'1, and yields the following values for spectral shifts: AQs « 5.1, AQA » -1.2 cm"1.81 Temperature-dependent contributions to shifts and widths of split spectral lines in the temperature range from 11 to 20 K are presented in Fig. 4.8 and can be approximated as follows:
4.3. Generalization of the exchange dephasing model
122
0,4
0,12
0,08
I 0,04
Fig. 4.8. Temperature dependences of shifts (a) and widths (b) for Davydovsplit spectral lines of local vibrations in the 2x1 phase of CO molecules adsorbed on the NaCl(lOO) surface.
ReWs « 1.90«(ft>s),
- I m ^ s * 0.361w(<»s),
ReWA « 4.50n(a)A),
- ImWA * l.473n(coA),
coA « 39cm"1.
(4.3.48)
The nature of these dependences changes dramatically at T > 20 K, as an orientational transition from the 2x1 to the lxl phase of CO molecules results at T * 20 K.
4.3. Generalization of the exchange dephasing model
123
To describe the temperature dependences observed, one can first attempt to take advantage of well-known formulae by Erley and Persson185 (4.3.39) written for symmetric (/-S) and antisymmetric (/-A) vibrations: ReW,-yrkcoA
AQ y - y
1+7 7
J
[
r
[AQJ-Y)
-lmWj=r2>b»j)l
\
2
w
(4.3.49)
2.
Since it follows from Eq. (4.3.48) that \lmWs\ < |ImffA|, the second formula in Eq. (4.3.49) leads to the inequality |AQS - Y\> |AOA -r\- To account for positive shifts, it is necessary that y > 0. Moreover, AQS > 0 and AQA < 0. However, with this conditions included, the first formula in Eq. (4.3.49) can yield only ReWs > Re^ A , whereas just the reverse is true for real systems. To determine the characteristics of the 2x1 phase in the system CO/NaCl(100) from general formulae (4.3.47), we equate expressions (4.3.47) and (4.3.48) thus deriving four equations in four unknown parameters, y, ij and Aft)/ withy = S, and A. It is noteworthy that for the spectral lines associated with local vibrations S and A, the vector k assumes two values: k = 0 and k = kA (kA is a symmetric point at the boundary of the first Brillouin zone). The exact solution of the system of equations provides parameter values listed in Table 4.3. The same parameters were previously evaluated by formulae (4.3.49) without regard for lateral interactions of low-frequency molecular modes." As a consequence, the result was physically meaningless: the quantities y and rj proved to be different for vibrations S and A (also see Table 4.3). Table 4.3. Parameters of Davydov-split spectral lines for the 2x1 phase of the system CO/NaCl(100) calculated by the formulae of Erley and Persson (4.3.49) and by the generalized formula (4.3.47) accounting for the dispersion of low-frequency CO vibrations.
Parameters
Values calculated by Eqs. (4.3.49) (cm"') Surface-normal Surface-parallel symmetric antisymmetric vibrations vibrations (/=A) o=s) -1.9 4.6
Acfly
Y
n
1.5 2.8
5.2 15
Values calculated by Eqs. (4.3.47) (cm"1) Surface-normal Surface-parallel symmetric antisymmetric vibrations vibrations 0=S) 0=A) 5.1 -1.2 6.8 3.6 3.4 3.4 0.63 0.63
4.3. Generalization of the exchange dephasing model
124
The resonance width for low-frequency modes r/q averaged over wave vectors is given in the Debye approximation as follows:143
*=7T2>«=74'
(4 150)
-
where m/ is the reduced mass of low-frequency vibrations for adsorbed molecules, and p and c, are density and the transverse sound velocity for the substrate material (see Eq. (4.2.25) and the alternative derivation of this formula in Appendix 1). Thus, Eq. (4.3.50) permits the parameter 77 to be independently estimated. As an example, for the system CO/NaCI(100) (m, * 2.6810"23 g, a, * 37.5 cm"1, p « 2.18 g/cm3, c, « 2.2 105 cm/s), it is found that 77« 1.2 cm"'. With reference to this value, Eq. (4.3.47) provides a more adequate value of the parameter 77 than Eq. (4.3.49). Consider vibrational excitations giving rise to small angular azimuthal deviations v the following expression for the excitation Hamiltonian in the harmonic approximation: H
ex = ~Yj 7 W
+J
(qW
(4.3.51)
where / = m(2d) sin 9 is the molecular moment of inertia (2d is the length of molecular bond) and the function [4fi3 cos2/9 + (2B3 +fl4)sin26>]sin26>cos;c
j(q)= j(qx,qy)=
+ 2(2S3 + B4)(cos2 0 - sin 2 0)cosqy
(4.3.52)
-2fi4COS 2 0sin 2 0-4(£2 + 2fl3 + B4+2fl 5 )sin 4 0-8AC/4 represents the dispersion law for orientational vibration provided that J(JI,Q) =
-4(B2 +4B3 +2B4 +2S 5 )sin 4 <9-8A£/ 4 > 0 .
(4.3.53)
4.3. Generalization of the exchange dephasing model
125
At AC/4 = 0 and s i n # « 1, the value J(nfi) goes to zero, which gives rise to a Goldstone mode in a dipole-like system with the degenerate ground state, with the degeneration removed by thermodynamic fluctuations at nonzero temperatures.70 At sufficiently small negative values of the reorientation barrier AU4, an energy gap arises which corresponds to orientational vibrations of the frequency coi for isolated molecules, so that the dispersion law for orientational vibrations in the case of pure quadrupole-quadrupole interactions takes on the form:52 a>l =ft>/2+3 ,],
(4.3.54)
where t- sin #and
con=-
1 U 2d\m
(4.3.55)
is the characteristic frequency for the quadrupole system. With the parameters coi = 37.5 cm 1 , cog = 9.3 cm"1, / = 0.1786 (0= 25°) inherent in the system CO/NaCl(100), the dependenceft>qis plotted in Fig. 4.9.
(0,0)
<*.0)
(*.*)
(0, it)
(K.0)
Fig. 4.9. Dispersion laws for orientational vibrations of inclined quadruples (0= 25") on a square lattice relative to the ground state for the 2x 1 phase of CO/NaCl( 100).
126
4.3. Generalization of the exchange dephasing model
Gently sloping spectral regions arise from small values of the coefficient of cosqx and evidence for a stronger anisotropy of quadrupole interactions compared to dipole interactions. The average band halfwidth for collectivized orientational vibrations proves to be of the order 10 cm"'. Thus, the orders of magnitudes of effective band halfwidths for low-frequency vibrations, A&>;, (see Table 4.3) are consistent with the estimated band halfwidth for orientational vibrations of CO molecules which value is dictated by intermolecular quadrupole-quadrupole interactions. The values Aa>s exceeding AcoA evidence for a lower density of lowfrequency states in the band of symmetric vibrations compared to that of antisymmetric vibrations.
Appendix 1 Local and resonance states for a system of bound harmonic oscillators The Hamiltonian function for a system of bound harmonic oscillators is, in the most general form, a sum of two positively definite quadratic forms composed of the particle momentum vectors and the Cartesian projections of particle displacements about equilibrium positions:
i
'
ij,ap
where m, is the mass of the ith particle; $ denotes the force constants; a, /} are the projections of corresponding vectors on the Cartesian coordinate axes (summation over Greek indices will be indicated explicitly because it appears in the expressions more than twice). The equations of motion for the displacements u" that correspond to the Hamiltonian function (Al.l) are
«,«?+£«#»«/=<).
(A1.2)
Both quadratic forms in Eq. (Al.l) can be diagonalized simultaneously by changing to new (normal) coordinates x:
qv
where C?v is the real orthogonal matrix with the indices /' and q ranging from 1 to N (N is the number of particles in the system) and the indices a and v ranging from 1 to d (d is the dimensionality of the space of displacements):
XC£VCV ia
=S
qq'SW > X C ™ C J q qv
127
=S S
V a/3 •
( A l A)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
128
Since p" = THJU" , the kinetic energy in (A 1.1) takes the form
(A1.5) ia
'
aq'vv'
qv
Using the formula (A1.3) we derive the potential energy as follows: (A 1.6) ij,ap
qq'vv'
ijafi
For the potential energy to be diagonal with respect to xq , we must require that Y(mm £JX"I'"J)
Yl2d>aPcavCPv,' 1
^ij
^iq^jq
=a>2 5 .8
(A 1.7)
w
qvuqquv
ijap
By multiplying both sides of the Eq. (A1.7) by Cjff and summing over q',v' (with the Eq. (A 1.4) taken into account), we have:
2("i»irV2*7csv=*jvcg',
(A 1.8)
i.e., the columns number qv of the matrix C" v are the eigenvectors of the matrix (OT;/M/)~
'
2
^ , ^ ,
]=Q
(A1.9)
By substituting the equality (Al .7) into the relation (Al .6) and adding the result to Eq. (A1.5), we deduce the Hamiltonian function expressed in terms of normal coordinates:
*4l[feN^W qv
(ALIO)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
129
and the corresponding equations of motion for xq : xvq+co2qvxvq=0
(Al.ll)
which coincide with the equations of vibrations of a harmonic oscillator with the natural frequency coqv. An important and convenient characteristic of the system is its Green function (GF) that describes the response of the system to an instantaneous perturbation. We introduce the GF corresponding to Eq. (Al.ll): Gvq(t)+w2qvGvq(t) = -S(t),
(A1.12)
with the minus sign for the 8(t) corresponding to a standard definition by which (as will be shown below) \mG{co) < 0. The solution of Eq. (A 1.12) is: Gq{t)=-e(t)smcoqvt/G)qv.
(A1.13)
The frequency Fourier component of the GF is defined by the relation: 00
C»=
\GVqWmdt.
(A1.14)
If expression (A1.13) is directly substituted into Eq. (A1.14), the primitive of the integrand is easy to find, but the substitution of the integration limits, 0 and oo, by the Newton-Leibnitz formula results in the uncertainty at the upper limit such as lim e^°~miv'
. With the attenuation "k-+ +0, the uncertainty is removed, and we
/->oo
arrive at: Gq(a))=[G)2-a)qv+iQsignG)f
,
(A 1.15)
where ) l at0 signco = < 6 1-1 atft><0
(A1.16)
130
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
The presence of an infinitesimal, purely imaginary addition in the GF denominator turns out to be very important and to have a deep physical meaning. Its origin is related to the theta-function in Eq. (A 1.13) that represents the causality principle: The reaction of a system can be caused only by perturbations at preceding instants. Note that the relation (A 1.15) (though without the imaginary addition to be completely defined later) is derivable by the Fourier transform of Eq. (A 1.12). Since 00
(A1.17)
s{,)=±
j.-~< dco
(the last identity is one of the integral representations of the 8-function), then Gg(o))= \co -o)gVJ
. We show how the inverse Fourier transform of Eq. (A 1.15)
performed by the theory of complex variable function takes us back to the original expression (A 1.13). Since exp(-iwt)=exp(-itRecot)exp(tlma> t), the integrals over the semicircle of an infinitely large radius are zeroes in the upper and the lower half-planes at respectively t < 0 and t > 0. Therefore, the integral along the real axis to may be replaced by the integrals round the following closed circuits (see Fig. Al.l):
Fig. A 1.1. Integration contours in formula (A 1.18).
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
2K
Gtth
fa(a,yim'dco,
131
t<0, (A1.18)
1 — \GvAa>yiai'da>, t>0. 2n
The poles of the integrand co- ±coqv - i'0 lie in the lower half-plane (see Fig. Al. 1), so that at t < 0 the case is reduced to an integral of the analytic function round a closed circuit in the upper half-plane which is equal to zero. At t > 0, the value of the integral is found by the theory of residues: G"{t) =-2xiR.es
ha>*
1 -sin
ieat 0}.
v?,
(A1.19)
qv
so that we again arrive at (A 1.13). Thus, the infinitesimal imaginary addition at the pole of the GF (A1.15) really arisesfromthe theta-function in Eq. (A1.13). We now introduce the GF G? (?) for the displacements uf that satisfies the equation
mffl(*)+,Z*7Gp)!S-StovS<*
(A 1.20)
JY
and determines the response of the displacements uf (t) to the external forces F" (which should be substituted in the right-hand side of Eq. (A 1.2)):
u?(t) = -YlGf(t-t')Ff(tyt'.
(A1.21)
It is straightforward to show that Gff{t) is related to the GF G^{t) by the relation Gf(t)=(mimjy^Gvq(ty:rqvcf; qv
.
(A1.22)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
132
For this purpose, we only need to substitute Eq. (A1.22) into (A1.20) so as to deduce, using formulae (A1.4) and (A1.8), Eq. (A1.12) for Gq(t). Formula (A1.22) relates the GF of the non-diagonal representation for the initial displacements uf to the GF of the diagonal representation for normal coordinates xq . Evidently, the same relation holds for the frequency Fourier components of the GF, so that
Gf{a,h(!nlmjyy^-2 ^ . qv « - f t ^ + fOsigno;
(A1.23)
Invoking Sokhotskii formula, we obtain the imaginary part of the GF: ImGf(«) = -^(m,my)-1/2sign«^C«,'C^(«2-«2v).
(A1.24)
qv
Two significant properties of lmG^(a>)
following from Eq. (A 1.24) deserve
mention: 00
\co\mGf
{o))dco = (A1.25) 00
l2
v
-7r(mimJY s\^co'YJCfq C^q
v
qv
^mt
2
2
\\a>\8\co -(o qv)dco = - — 8^8^ , -oc
'
\mGf(co)=-7migncoYJS\(02
ia
-co2v).
(A1.26)
qv
Introduce a normalized distribution function for squared frequencies alternatively called a state density: 00
5 2
P^yj^L ^
qv
-^X
JP(®2W = 1 o
(A1.27)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
133
(N is the total number of particles; d is the dimensionality of the space of displacements). According to Eq. (A1.26), this function is related to \mGff(a>) at ea>0: P ^ - ^ ^ l m G ^ i c o ) .
(A1.28)
ia
The physical meaning of the function p(a>2) is that the quantity p(co2)d(o2 specifies the number of squared frequencies a^v (divided by their total number Nd) falling within the interval from a>2 to co2 + Sa>2. This essential characteristic of the frequency spectrum is also defined by the GF of the system. The above treatment needs to be summarized. Transform (A1.3) for displacements and relation (A1.22) for the GF make it possible to express the quantities desired in terms of simple solutions of Eq. (A 1.11) for independent vibrations of harmonic oscillators and their GFs (see Eqs. (A 1.13) and (A 1.15)). In so doing, the transformation matrices C" v and the frequencies coqv are found to be the solutions of the eigenvalue problem (see Eq. (A 1.8)). To explicitly derive these quantities is impossible in the general case, as it would require solving a system of Nd linear equations or a single equation in co2 (see Eq. (A 1.9)) of the degree Nd. Below we will examine three particular cases of importance in which the natural frequency spectrum is deduced in an explicit form. Consider a system of two bound one-dimensional oscillators whose interaction with each other and with the wall is effected by two different springs (see Fig. A1.2).
Fig. A1.2. Two coupled oscillators.
Then in the initial Hamiltonian function (A 1.1), all the indices, a and p, of the projections onto the Cartesian axes are the same and thus can be omitted, while the
134
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
indices of particle numbers assume only two values, i = 1,2. Designating the spring force constants as kx and ku, we come to
H=-^
2mx
+
2m2
^-+X-kxu^X~kn{u2~uxf, 2
(A 1.29)
2
so that the force constants Oy in Eq. (Al. 1) are: *ll!B*l+*12.
^12 =-*12,
(A 1.30)
«>22 = *12 •
The equation (A 1.8) becomes as follows:
®!2
jmxm2
(A1.31)
C
\q + / =C2q ~°> ^mxm2
•-(On
Clq +
''$22
2^
(A 1.32)
c2g=o.
m2
j
Hence, we have: 1/2 2
f<
1
S>\\
Di
$22^
+ 4 q>?2
, (A1.33)
/M,/M2
n'/2
cjq=(-iy^Hu(-iy^B)
(A1.34) fi =
(J>22 ^ «2
0H m
ID,
j,q = 1,2.
\
We simplify the expression (A1.33) for a case when k\ ~ £)2 and m2 « easily expanded in the small parameter mxm2
M
(m:+m2f
m2
mx
(
\
an.
,
m2 «mx
mx. It is
(A 1.35)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
135
which is the ratio of the reduced mass m = m\m2IM of two particles to their total mass A/= m, + m2. As a result, we obtain, accurate to terms of the order (mIMf:
M
m
\
m M
%\ m M1
2\
, m ^m m2 « m 1+ +2 ; M MA
(A 1.36)
and the formula (Al .33) takes the form: 2 1
( h 1- m H M M2 hi V
2 m2 ~
k
U
m
^
1 + m 1 k\ M k12
(A 1.37)
So, at m 2 « mi, the natural frequencies of the system correspond to the independent vibrations of the mass M on the spring k\ and of the reduced mass m on the spring kn- As the parameter B tends to unity for m2« mu the relative displacement of particles 1 and 2 is approximately described by the normal coordinate x2: 2 ~u\ ~ X2/\m
u
.
Thus, in view of the inequality m2« mu we derive a physically obvious result: the motion of a light particle 2 can approximately be regarded as vibrations about a relatively immobile particle 1. The squared frequency of these vibrations would be equal to ki2/m2 thus differing (by virtue of the approximate equality (A 1.36)) from k\2/m2 by the terms of the order mIM. Besides, the second formula in (A 1.3 7) shows that the approximate notation of the sought-for frequency as knlm is preferable because it is valid accurate to terms of the order (mIMf. As another example illustrating an explicit switch to normal coordinates, we consider a three-dimensional monoatomic simple lattice. In such a system, masses of all particles are the same and the positions of their stable equilibria are at the lattice sites which are given by radius vectors n (called lattice vectors). Instead of an unsystematic particle numbering (/' = 1, 2, ..., AO, it is now convenient to distinguish them by the lattice sites they belong to and to designate them by the index n. The periodicity of the particle positions implies that the force constant Ojjjj, for a pair of particles n and n' should be the same as that for a pair of particles n+a and n'+a, where a is an arbitrary lattice vector. Thus, „+st n . + 9 but, by the n+a,n +a
arbitrariness of a, we can put a = -n' and then O^j, = ^°_nt
0
= *(n - n'), i.e., the
force constants are dependent only on the difference n-n'. The displacement of the lattice as a whole should not result in changed force constants, which is equivalent to the condition:
136
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
^ 0 a / ? ( n - n ' ) = O.
(A1.38)
Note that in the above-discussed model (see Fig. A 1.2), the particle 1 is "bound" by the spring k\ to an immobile wall, so that the condition (A 1.38) was not met (see Eq. (A 1.30)). The Hamiltonian function H—^vl^^^-^yy^Q n
(A1.39)
nil'
is easily diagonalized by the transform (A 1.3) in which m, = m, q is the vector running through a quasi-continuous spectrum of values (at N » \) provided cyclic boundary conditions: "n+A'.a^Un'
' = 1,2,3
(A1.40)
(ai, a2, a3 are the basis lattice vectors; N\, N2, N} are the large integers that fix the main area for a lattice of N = NiN2N3 sites). The matrix C^„y has the complex matrix elements CnqV/=-7=^(q)exp(/qn)
(A1.41)
obeying the unitarity relations
Z
cavc*av' L
nqL
g g nq
°qq'°w
na
•STcavc*0v ' ^Lnq qv
L
n 'q
g g
fA142,>
°nn °ap
\J\i.tt)
which generalize the orthogonality conditions (A 1.4) for real matrices. As a result, the transform (Al .3) becomes: u °
n
= ' mN
^
2/q e av(q)exp(/'q • n) qv
and then formula (A 1.40) is equivalent to the condition:
(Al .43)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
q-a =-j?-g„
gt=0,±l,±2,...
137
(A1.44)
The solution of this equation for q is well known: 3
— b , , b i = — a 2 x a 33, V0 '
b2 2 = — a33 x a i , F0 ''
j=lNt b
a
3=— l
xa
2-
(A 1.45)
a a
^0= l 2a3=ai[a2xa3]
^0
and explicitly defines the spectrum of values for q. The vectors b/ are called the basis vectors of a reciprocal lattice and satisfy (as it follows from Eq. (A1.45)) the relations: a,--b<=2^-.
(A1.46)
The quantity V0 is equal to the volume of a lattice unit cell. Since exp(/'qn) is a periodic function with the period In, the integer values gt may be restricted by the inequalities: -Nil2
i = \, 2 , 3 .
(A1.47)
The range of q corresponding to the above inequalities is referred to as the first Brillouin zone. It is now easy to prove the following two equalities of significance: — £exp[/(q-q')-n] = £qqS n
_]Texp[*(n-n')q]=<W q
(A1.48)
in which the summation variables n and q run through exactly TV above-specified values. Substituting Eq. (A 1.41) into Eq. (A 1.42) with regard to Eq. (A 1.48), we obtain the unitarity relations for the matrices e a Xq):
2 e «v( ( l)w(
(A1.49)
138
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
Eq. (A 1.8) that defines the frequency spectrum cojjq) for the system assumes the form:
-E*^(q>>(q)=«»?(q> 0 .>(q)>
(ALSO)
m
p
where the third-rank matrices «%)=£ CD"/* (n>r'l n
(A1.51)
n
are the coefficients for the expansion of force constants in a Fourier series in terms ofq:
(A152)
q
Note that i>a^(n) = °^(q) = *#*(-q) since values °^(n) are real. As a consequence, *^(q)=0^a(-q), so that the matrix <£>
(A 1.53)
(q) is Hermitian, its eigenvectors e„(q) are orthogonal (in the
sense of the first equality in Eq. (A 1.49)), and eigenvalues <wv(q) are real (the positivity of Q)y (q) followsfromthe positive definiteness of the quadratic form for the potential energy). We subject Eq. (A1.50) to the complex conjugation operation and take into consideration the property (A 1.53). Then the quantity eav(q) will represent the eigenvector eav(-q) with the same eigenvalue cov(q), i.e., a)l{-q) = a>l{<\),
eav(-q)=e*av{q).
(A1.54)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
139
For inversion-center lattices, a stronger condition, aP(q) is real and symmetric and hence the eigenvectors ej[q) may be considered to be real. The GF (A 1.23) of the system in question is
p G *>,«) = — - > — exp(*qn m N t o ^ -«»?(q)+»0signfl>
(A1.55)
and on expanding it in a Fourier series in q (as in Eqs. (A1.51) and (A1.52)), the resulting coefficients appear as
G p{q,co) = - \ - 2
TFT
:
'
(AL56
)
m y (o -o)y(q)+ /"Osignoj The quantity (A 1.56) at co> 0 is associated with the so-called spectral function
a
v
which characterizes the absorption spectrum if the radiation with the frequency a> and the wave vector q is absorbed by the system concerned. The definition of S\f\,co2 J gives its relation to the state density p(o?) (see Eqs. (Al. 27) and (Al. 28)) and the normalization condition: 00
2
S q w2
p(« )=^X ( ' q
)'
jfs(q,<»2)*»2 = 1 -
(A1.58)
o
Importantly, from Eq. (A1.51) it follows <S>aP{q=0) = 0 (in view of Eq. (A1.38)) and hence Eq. (2.50) yields a v ( q = o ) = 0 . Vibrations characterized by this property are said to be acoustic. The symmetry properties expressed by Eq. (A 1.53) result in the fact that there are no terms linear in q in the expansion of aXq) in small q. The expansion will, therefore, begin with quadratic terms which have the following general form for an isotropic lattice:
140
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
-0^{q) m
+ Xlq2Saf).
= ^qagfi
(A1.59)
Substituting Eq. (Al. 59) into Eq. (Al. 50), it is readily seen that «l2(q)=(^+^)72> *>2,3(q) = 4 7 2 ,
e1(q) = q/;
(A1.60)
e 2 (q),e 3 (q)le 1 (q).
(A1.61)
The time dependence of the normal coordinates x„ obeying Eq. (A 1.11) is given by the factor exp(-icov(q)t), so that the displacements of the particle, u„, corresponding to this normal vibration are related, according to Eq. (A 1.43), to the coordinates n and to the time t by the cofactor exp[/(qn-ftj v/ (q)?)j. This cofactor describes a plane wave propagating in the direction of the vector q. The phase velocity of the wave propagation is specified by cv =cov(q)/q . For small q, these waves are acoustic (because the wavelength 2nlq notably exceeds the size of the lattice unit cell) and the expansion coefficients A, and A^ in Eq. (A1.59) are related to acoustic wave velocities. The solution (A 1.60) accounts for longitudinal waves (in which the particle displacements are parallel to the wave propagation direction), and Eq. (A 1.61) describes two degenerate (since o>i = coi) transverse waves. The velocities of the longitudinal (c>) and the transverse (c,) acoustic waves are expressible in terms of A! and A2 taken from Eqs. (A1.60) and (A1.61): c
l = TJAI+1%,
cr=^2
(A1.62)
with ci > c,. The dependence of the frequency on the wave vector q is called the dispersion law; it is linear and the corresponding distribution function for squared frequencies (see Eq. (A 1.27)) can easily be calculated by replacing the integral over q (with N -» °o) for the sum over q. We demonstrate this technique below. The summation over q implies summing thrice over the integers g, (/' = 1, 2, 3) in Eq. (A 1.45). Changing to the integration over dgidg2dg3, note that the Jacobian of the transformation from gt to qa is (according to Eqs. (A 1.44) and (A 1.45)):
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
141
where V = NV0 is the total volume of the main area. As a result, we are led to a frequently used relation:
2-7^/--
(A 1.64)
(In the case of a (/-dimensional space, a pre-integral cofactor changes to V/(2xY , with V denoting the "volume" of the main area in a of-dimensional space). The quantity p{a>2) is now easy to find by Eq. (Al .27) (or Eqs. (Al .57) and (Al .58)):
(A 1.65) 6TT'
*
a.Tr n 4x c~
J__l
(A 1.66) c
/
w y
Of course, as q increases, the dispersion law co^cfi deviates from linear, so that the integral over q in Eq. (A1.65) should be limited from above. It is reasonable to so limit the admissible frequencies co that the state density normalization (A 1.27) is preserved: G>D
\p(a)2)dG)2 =
"°
6TT 2 C 3
<*-•.
(A 1.67)
Hence, we obtain:
tOD=c
<7D=(67r2AoJ
•
(A 1.68)
Such an approximate description of acoustic vibrations is referred to as the Debye approximation and the limiting frequency coo is called the Debye frequency. The
142
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
corresponding limiting value of the wave vector q is of the same order as the reciprocal lattice constant. The GF (A1.55) satisfies the equation:
n>
which represents a frequency Fourier transform of Eq. (A1.20) for a periodic lattice of particles with identical masses. Suppose that the particle at the site n = 0 has the mass m c = m - Am different from the mass m of the other particles. The equation determining GF G^,(w) for such a system is conveniently written as follows: W«
2
Gn^(W)-Xo^(n-n")G?n-M =Am^Gnf,(«Ko+^n„^.
(A1.70)
The left-hand sides of Eqs. (A1.69) and (A1.70) are of the same form, which enables the perturbed GF with the mass defect Am to be readily expressed in terms of the unperturbed one: G^^G^yAmco^G^ico^U. r
(A1.71)
Assuming n = 0 in Eq. (A 1.71), we come to: £kr '^G^p^GfM-
For an isotropic lattice, the matrix G^(a)
(Al-72)
is also isotropic and, on the
strength of Eqs. (A1.27) and (A1.55), expressible in terms of the state density: G
oo (®) = G o o ( « K r .
G
oo{0>) = ^^Yjt°2
- « ^ ( q ) + ' 0 signwf
qv
m m J mL - ™Z 4- ift c o n m
* co —S + /0 signco
L \
/
\ /J
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
143
In the latter equality, the Sokhotskii formula is used and the designation P(o)2) is introduced for the principal integral value:
(A 1.74) 0 °>
-
a
Thus, the lattice isotropy permits a straightforward relation between the perturbed and the unperturbed GF which is obtained without solving the system of linear equations (A1.72) in the general case: Gg{a>) = Ggf (<»)/[l - Amco2Gw{a>)}.
(Al .75)
By the very definition of the GF, the real parts of the poles of its frequency Fourier component correspond to natural frequencies of the system (see, for example, Eqs. (A1.23) or (A1.55)). Consequently, the spectrum of natural frequencies of the perturbed system, mp, should fit the equation (Q2p
sk-*wr'-£
3N^,rp
qv
-VVM/J
Am
< AI ™>
corresponding to the vanishing denominator in the GF (A 1.75), with ReG 0 o(«) taken from Eq. (A1.73). The qualitative behavior for solutions of Eq. (A1.76) is conveniently exemplified by several equidistantly disposed nondegenerate frequencies o)y(q). A graphic solution of this kind, with 37V = 6, is presented in Fig A1.3. First of all, it is noteworthy that the parameter m/Am = (l-mc/m)" can assume any negative value at mc > m and any positive value greater than unity at mc < m. In this physically admissible range of the values mlAm, Eq. (A 1.76) has exactly 37V roots, as is to be expected for a system with 37V vibrational degrees of freedom. At mc -> m, we obtain mlAm -> ±oo and the spectrum of natural frequencies coincides with <w^q). At mc * m, each of the values cop is necessarily between two neighboring squared frequencies l((\) (with different q and v) have the same value (the value of o)v(<\) is then said to be s-fold degenerate),
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
144
then the degree of Eq. (A 1.76) with respect to cop will decrease down to 2>N - s + 1 and we seemingly lose s-\ roots. As a matter of fact, regarding the s-fold degenerate squared frequency as a limiting case of s very close values, it is clear that s-\ "lost" roots are just equal to the degenerate value cov(q), since the alternation principle leads them to lie between values tending to each other. Am/m
Fig. A1.3. Graphical solu ion of Eq. (A1.76) with six equidistant non-degenerate squared frequencies a>l{<\ (3/V=6). The dash-dotted line connects the bending points of solid curves and defines quasilocal vibration frequencies.
Among 37V values of co „ , it is reasonable to distinguish the most sensitive to a change in the mass mc of a "defective" particle. With mc < m, the largest changes are exhibited by the maximum value of cop that tends to infinity as mc -> 0 {mltsm -» +1). The vibration corresponding to this value cop is called the local vibration of a "defective" particle. The implication of the term is that the displacement amplitudes for particles which vibrate with the local frequency a>p drop fast away from a "defective" site n = 0. Indeed, the analysis of Eqs. (A 1.70) and (A 1.71)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
145
demonstrates that the displacement amplitude u"(n) for a particle at the site n is related to that for a central particle (of mass mc), ua(0), by the unperturbed GF : ua(n)=Amco2^G^{coy{co). P
(A 1.77)
With mc « m, we have cop »(m/mc)la)v(q)) » maxa)v (q) and the GF G^(a>) defined by Eq. (A 1.55) decreases fast with increasing |n|. The law governing this decrease is given by a specific form of the functions (o2(q) and e^q) but the decrease as it is can easily be noticed even in the limit G^ff (©)-» \mo)pJ SasS„Q. At mc > m, it is also possible to distinguish thefrequenciescop most sensitive to a change in mc. This is less trivial than for local vibrations, because the spectrum mp at mlAm < 0 has no split-off frequencies and alternates with the spectrum co2(q) (see Fig A1.3). Transform the sum in Eq. (A1.76) to a form convenient for a subsequent consideration. Moreover, the transformation in question will present an illustrative example for a direct deduction of formula (Al .73) for G00(»). Label all squared frequencies co2(q) of the unperturbed spectrum in increasing order with the index m: cov(q) = coom, so that »om+l > ^Ow ('"=1,2, ..., max, the subscript 0 indicating valuesfromthe unperturbed spectrum). In so doing, note that some values co^m may be associated with several different vibrations (i.e., several different normal coordinates in Eq. (ALIO)), with the degree of degeneracy for the value a)Qm designated by 9m. We also introduce the quantity 8co§m for the gap 7
7
7
between neighboring squared frequencies: Sco0m =#>o/n+l ~O)0m • Then an arbitrary 7
1
1
value cot falling within the interval [a>op,fttyp+llmay be represented as: wj = a>lp + zpSa>lp ,
(A 1.78)
where zp is the dimensionless parameter ranging from 0 to 1. With this notation, we rewrite the desired sum in Eq. (A 1.76): l y 1 2 ™^co p-col{q)
1 y 1 3N2p
1 $p 3N ZpSa>2p
+
146
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
1
3JV
°>lp - °>lm (<»lp ~ <4m + Zp8<»lp )(«0p - mlm )
">\*P.
(A 1.79)
fl»
_L y 3N
ZpScop
J^p)<°lp-<»lm
1 (2
3Mfa>,'Op
2
P) VOp-MOm
Sea, .2 Op
\
2 2 MOp-VOm
+ Z„
Scolp
P
Consider the bracketed sum. Its terms decrease fast as m moves away from p, which provides a number of simplifications concerning the behavior of the functions 9m and co0m near m = p. First, we set 9m = 9p
and second, regard the spectrum
of WQm as a equidistant one (these assumptions are quite warranted in a small vicinity far from band boundaries and singular points). The sum concerned then takes the form:
y
_ 1 yJ__Y
1
^-in{n + zD) „*n
v
"I—
I
z'D' W^-in ^~Ln +: O " n*0' PVn±Q ntn
P'
00
1
•*-' n + 2 P\n= \n=-ao
'P n*0
1 ~z
4
K
(A 1.80) COt^TZ„.
Z
P
We now note that the distribution function for squared frequencies p\a>2p j (see Eq. (A 1.27)) is related to SO)QP and 9p as follows:
prlr T^S^r2 ~ ^(q))33N^ S ^ W * + aom +
z d( °l P °0p
qv
(A1.81)
* J^H5« m
]dzpSV>0p + <°lm + Zp8(»lp) = o
INScoL
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
147
An approximate smoothing with regard to the gap scales in Eq. (A 1.81) is implied by the very definition of the state density involving a ^-function and it becomes exact in the limit N -» oo, Sa>Qp-> 0. The first summand in Eq. (A 1.79) is also expressible in terms of pica2):
— Y -T^T= 3N W
mPp) QP - VOm
3m
Y
myp)
SC 0m 2
2
°
2
WScQ0m a>0p- co0m
" K - 2 h ^ T - (Al-82) J Q
co0p - a
which is in accord with the above-introduced notation (A 1.74). Substituting formulae (A1.80) - (A1.82) into Eq. (A1.79) yields the final expression valid for N -> oo:
T^T
2
X
2(,
= p{<°lpY*p{«>lP)x>tnzp
.
(A1.83)
We can now conveniently select, by substituting Eq. (A1.83) into (A1.76), such a value from the spectrum of cop that is the most sensitive to a change in the parameter s = Am/m. For this purpose, differentiate the displacements zp with respect to er. dzp
sin2xzp 7T2colpP\G>lp)
G*IPP(
^-colpPyalp)]
(Alg4)
+n2G)%pp1(a>lp)
and require that the absolute magnitude of the derivative dzjds be maximum. As evident, this is ensured by the following equivalent equalities: <°0pP(«>0p}= m/&™,
ZP = 1/2 •
( A1 - 85 >
The graphic interpretation of the above conditions consists in selecting the branch of the cotangent curve in Eq. (A1.83) that intersects the straight liney = m/Am at the inflection point (see Fig. A 1.3 where the inflection points are connected by a dotted line). For the Debye spectrum, we have
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
148
PW
1 co
col
0)0
In
2aD _ I - 3&>5 + 3 co I coo, \a>~2 + 3cop/5co4,
+0)
(Op -CO
(A 1.86) o)«Q)D o)»coD
and the graphic solution for Eq. (A 1.85) is shown in Fig. A 1.4. co2P(co2
Fig. A1.4. Graphical solution of Eqs. (A1.85) and (A1.99) for a Debye spectrum.
Two roots lie in the range -0.635 < m/Am < 0 (mc > 2.57 m), the root with a smaller value cop corresponding to a smaller value p( cop), so that it is this root which is the most sensitive to a change in mc on the strength of the criterion (A 1.84). Vibrations with the frequencies meeting the condition (A 1.85) are said to be resonance or quasilocal ones. The imaginary part of the GF (A 1.75),
lmG$(co)--
np\a>2) m
Am 2 J 2\ + co P\coL) m
[^-c°2^\a)2)
(A 1.87)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
149
which is proportional to the absorbed radiation power for the frequency a and, in view of the fluctuational-dissipation theorem, to the mean-root-square vibration amplitude for an impurity atom is indeed of the resonance nature. It should be pointed out that the term with c t g ^ in Eq. (A 1.83) separates out the contribution to the sum that varies fast on a small scale of displacements SCOQ„ (cf. Figs. A 1.3 and A 1.4). Adding an infinitesimal imaginary quantity to zp (involved in Eq. (A1.73)) followed by roughening with respect to Zp (as in Eq. (A 1.81)) results in zo\.{mp + <0) replaced by -r.
fcotta J v p
+
*>W, r
p
= l l n ^ t i O ) = l l n ( _ , ) = _,-. K sin(/0) n ' v
0
;
(The minus sign for /' in the last equality is chosen to correspond to the sign of the imaginary part ctg(^zp + iff)). Thus, Eq. (A 1.83), if roughened with respect to zp, turns to Eq. (A1.73). For positive values mlAm > 1 (jnc > m), we get cop>a)D, p\(OpJ^>0emd return to the local vibrations whose frequencies are specified by split-off branches in Figs. A 1.3 and A 1.4. The studies in the effect caused by defects on lattice vibrations were pioneered by Lifshits139'188. Much later a related research of applied character was performed by Montroll et al.189"191. The discovery of the M8ssbauer effect192 gave impetus to investigating defect effects on dynamic properties of crystals and, most importantly, to predicting resonance vibrations193"195. A more detailed account on the theory for local and quasilocal vibrations of a lattice defect can be found in monografs.196197 Vibrations of an impurity atom or an impurity with internal degrees of freedom were examined in Refs. 198, 199. In conclusion of the present Appendix, we consider the vibrations of an impurity oscillator incorporated into a system of other bound oscillators. Assume the impurity particle C to be harmonically bound to a main system of oscillators numbered by / = 0, 1, 2, ... through a single particle with the number i = 0. Fig. A1.5 shows particles labeled by / at cubic lattice sites. The complete Hamiltonian function of the system under discussion is represented as follows:
L
a
1=0
'
ij
a
(A 1.88)
150
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
Fig. A1.5. An impurity particle C in a crystal with a cubic lattice.
The last term of the above equation amounts to the work done by the external force F(7) swaying the impurity bond. The equations of motion for frequency Fourier components of the displacements u(co) and the force F(co) that correspond to Eq. (A1.88) have the form: -mcco2u§+ka^~uS)=Fa{co), - m^uf
+ £
+ ka («? - 2£ )slQ = -Fa {w]Si0.
(A 1.89) (A 1.90)
By Eq. (Al.89), we express UQ -u" in terms of u" : -mcco2uZ+ka(^-u$)=Fa{a>)
(A1.91)
and substitute the result into Eq. (A 1.90);
(A 1.92) jfi
mca>
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
151
Compare Eq. (A 1.92) with the frequency Fourier transform of Eqs. (A 1.20) and (A 1.21). The right-hand side of Eq. (A 1.92) may be regarded as the Fourier component of the external force, but then it turns out that
"?(-)=2>3»
mcco^F^j^+k^)
(A 1.93)
rriQCO
where Gf^yco) is the frequency GF of the main oscillator system that is defined by Eq. (A 1.23). Rewrite Eq. (A 1.93) for / =0 as:
I
s
ap~
P
mca2-kp
G$(o>) p mca)
-kp
and immediately arrive at the equation for the frequency spectrum cop of a system with an impurity oscillator:
det
In the limit cop »co "
mcfOpkp
V
pi
\
= 0.
, we obtain GQ{?[cop)-+[mQCOp f 5an and Eq. (A1.95) M "max
yields, as it must, three roots, copa =ka/m,
*
"
"'
that correspond to the vibrations of
oscillators with reduced masses m^mcm^j(mc
+m0)
and force constants ka
(along the axes a = x, y, z). In this limiting case, ka lm » co for ReG"^ \copj in Eq. (A 1.95) is equal to -mccop coD ~ co
(A 1.95)
mca)p-kp
, but the cofactor
in the frequency range
and the situation reduces to the above-regarded vibrations of a defect
with the mass m0 + mc (cf. Eq. (A 1.72) where Am is represented by -mc). An important comment is relevant here. The degree of Eq. (Al .95) with respect to co2p is seemingly equal to 3(3N + 1), whereas a system with an impurity particle of the mass mc has only 3N+ 3 degrees of freedom. Indeed, numerators of diagonal elements of the determinant, with the whole sum over qv fractions in ReG"g [copj
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
152
(see Eq. (A1.23)) reduced to a common denominator, produce a polynomial of the degree 3N + \ in wi . On the other hand, calculation of the determinant 3x3 will result in a 3 times larger degree of the polynomial. In actual fact, the determinant (A 1.95) represents a polynomial of the degree 3N + 3 in cop and has 3N + 3 roots. To gain a better insight into the case, we make use of the known transformation139 of the general expression:
D{z) = te\[dap + ap {z)Gap (z)], Gap (z) = £
lalfi
(A 1.96)
kq-z
where A,q with different q has different values. The correspondence between the above relation and Eq. (A 1.95) (or the determinant that follows from Eq. (A 1.72) and refers to the situation with a mass defect) is quite obvious. Further, Eq. (A1.96) can be written in a detailed form: Z)(z)=l + £ a a ( z ) G a a ( z ) + i £ a
a/3 a G
a aa
a
G
a pa
•aPr ^a
^ya
a
PGafi a G p pp apGyp
a
G
a
a
G
a
a
P
a Pa
G
a/3
G
p PP
ayGay a G r Py
ayGn
W ^IXZT^T+TZ^Z ap
aa
lalp
(K, - z )(4,ii - z )
<7i<72"
9 |
a
I
Qi Ri <Jj
aPy
9l?2?3
q]
ia
ia
<1\
(A 1.97)
1 \
• Z )(A? 2 - z)(Ag3
a
l
2 1\ }P \P ~ Z) \ 1l
f
1\
lr
2
f1-i 1*
3
f3
With the same values of g, (/ = 1, 2, 3), determinants with / " have the same columns and go to zero, so that no denominator in Eq. (A 1.97) can have a cofactor (Aqi - z) to a higher power than the first. Decomposing the expression obtained into simple fractions
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
( \
153
- ' ) - & « „ - * .
we arrive at:
A„-Z
l"
bq(2)^aa(z){$}
l"
(A 1.98)
a„-V I"
lPlr
, •apr
la
+X«a(z)aa(z)/«X77^
°\12
l"
a
( ^ , -Ag)(Aq2
<7i
I" 92
ll
lP
lP
E
ly
lr
-&q) 1
1\
12
In the problem on a mass defect, aa is not a function of z, and q takes on 3N values. Therefore, on reducing the expression (A 1.98) to a common denominator, the numerator D(z) becomes a polynomial in z of the degree 3JV. In the problem on an impurity oscillator, aj^z) also has a denominator of the form (Aa - z) but differs in that the nominator D(z) is a polynomial of the degree 3N + 3, as we set out to prove. For an isotropic lattice, we take advantage of the transform given in Eqs. (A1.73) and (A1.74) to derive, from Eq. (A1.95), three equations in frequencies of local and quasilocal vibrations of an impurity oscillator:
co2pp(co2p)=
mQ mc
(Dn-(0n COn
co2=^iua
-
mc
a =x, y, z
(A 1.99)
The distinction of the equation (A1.99) from Eq. (A1.85) consists in the linear dependence of the right-hand side on cop. Let us analyze the solutions of Eqs. (A1.99) for the Debye spectrum (A1.65), (A1.67), and (A1.86). Fig. A1.4 illustrates that local vibrations occur at coa > coD and mc < m0. Quasilocal vibrations arise at wa < coD and mc < m0. It is noteworthy that the Debye approximation is not valid for frequencies close to coD and the graphic analysis given in Fig. A 1.4 is hence meaningful only for roots cop distant enoughfromG)D.
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
154
In the case of an isotropic lattice, Eq. (A 1.94) is readily solved for
"00°M =
>2p(co2)-i?ra2p(co2)
1
mcco-a
UQ(W):
~a^
(A1.100)
m^^_-^_a)2p^2yi;ro)2p^
Substituting Eq. (A 1.100) into (A 1.91), we find: (A1.101)
«8(a>)-«g{a>) = Xa(<»)Fa{) is the impurity bond susceptibility equal to:
X* ( » ) -
-m^
1
+
co2p(co2)-i7rco2p{co2)
^C
mcco2a
(A1.102)
m^w2-^a2pLlVixa2pLl\ co2
mC
The imaginary part of %a(co) proportional to the absorption spectrum for impurity bond vibrations is specified by the relation:
Im^M =
-
fel
7TCO p
^
.
(A1.103)
A
mc<*> a m
C
coa
In the frequency range for local vibrations, where p(co2)—><x>, the function Im x\°>) represents an infinitely narrow and high spike at co =cop {cop is the root of Eq. (A 1.99)). In the frequency range for quasilocal vibrations, the function (A 1.103) has a resonance character with a finite half-width. Introduce the spectral function normalized to unity: 00 2
a
s(a> )=-lmx (w),
js(co2)dG>2 = 1.
(A1.104)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
155
The normalization constant C can be found by calculating the zero moment of the distribution S(a?). Indeed, since the GF of an impurity bond amounts to -^(co), it follows from formulae (A 1.74), (A 1.67), and (A 1.102) that (
1
Mm oj2Cza{oj) =
\ = M0=-
m0
1^ mc
c-£
(A1.105)
m
and hence C = m = mcmo/(mc+m0). Assume Eq. (A 1.99) to have a single root cop. Then the function (A 1.104), with formally set p{co2) -> 0, should become Syo J=d\fo -ojpJ. Furthermore, in this limit,
*M-
mmQCO m
C
mcco^
coi
m02co 2 mc(m0 + mc)a)a dco1 {mc
s(co2-w2p),
col
(I)
=0)p
so that m0
m
C
a2pl
col
w
2 2 mQ(op 2
2
(A1.106)
mc{m0+mc)cOa 2
2
By approximately expanding the term vanishing at CO = eo in the denominator of Eq. (A 1.103), we obtain the Lorentz approximation for S(co2): S\fO j«—-. U
mc(m0 + mc )a)g7tp\(02p)l ni2. ^ = j -. r
[a2 ~ a>l j + h e ( m 0 + mC hl^p^l
r)/ml\
This yields the halfwidth of the distribution maximum for S(co2):
(A1.107)
156
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
Aa&.^^riwtpl).
(A1.108)
rriQ
Spectral line halfwidths measured over the scale of frequencies rather than squared frequencies are clearly related to: Aa^/j^Aco^^cOp.
Substituting here the
explicitly expressed quantity p\p>p) for the Debye spectrum (see Eqs. (A 1.65) and (A 1.67)), we are led to:
A«ty 2 *
2,n mc{m0+mc)(ac
^
=
W
4
K D.
mQ
~ ~ ~4 m0+.m~ cmcco t»o 8npc3
^
m
)
where p = m^V0 is the crystal density. In normalizing the expression (A1.103), we have already taken advantage of the fact that the quantity -xa(®) is the GF Gaa(co) of an impurity bond. It can be strictly defined in a standard manner:
GaP(t) = -U(t^c(t)-ua0(t\
uac{0)-4(Q)).
(A1.110)
Setting up the equations of motion for GF Ga\f) with the Hamiltonian of the problem (A 1.88), we can relate its frequency Fourier component with the GF Ggf (a») for the main subsystem of particles. Actually, this idea has already been implemented in terms of Fourier components of displacements u" {co) and UQ (CO) (see Eqs. (A1.89) - (A1.94) and (A1.100) - (A1.102)). Yet, after writing Eq. (A 1.94), we discussed the case of an isotropic lattice and applied the transform (A 1.73). Let us consider the result of a more general treatment in which the GF GQQ\G>)
is diagonal but not necessarily isotropic (as for Rayleigh surface
vibrations). In this case, the relationship between Ga^{co) and G^(
G°«(co) =
\
{
2
^cfG^{co)
mccoa [co -aa]lcoa-tnc(o that generalizes Eq. (A 1.102).
(AU11)
G%$(co)
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
157
Given two identical impurity particles bound to the atoms ;' = 0 and / = 1, and provided external forces F°(t) are absent, the system of equations (A 1.89) - (A 1.90) is generalized as follows:
- mcco2u£0 + ka (u^0 - u% ) = 0, ~mcco2u^n+ka(u^n-u^)
= 0,
(A1.112)
iP
Performing transformations like (A 1.91) - (A 1.94), we derive the system of equations: mcco2kp ^ 0 2~T~GOoH m^co
^W-^cSWW I G l{
"P
P
mcco
mC(Q2kp
P
mcco
-kp
°ap
= 0,
-*p
mC(o2-kp
a
= 0.
(A1.113) For a monoatomic lattice treated in the approximation of isotropic GFs, the equation is deduced for natural frequencies of the system: W
2
2
-COC
= mcco2 Re[G(0, a)-G(n, co)],
col ,
(A1.114)
G(n,«) = — Y — mNq(ozco {q)+ iOsignco where co2 = kjniQ . As pointed out above, the GF G(n, CO) decreases fast with increasing |n|. Importantly, the explicit form of the drop is much dependent on the form of the function w(q) not only at small values q, but also near the boundary of the first Brillouin zone. To estimate frequency displacements induced by crystalmediated (i.e., indirect) interaction between impurity particles, we restrict ourselves to the case of a linear monoatomic chain which enables, due to the simple explicit form of the dispersion law (co(q) = coD\sin(qa/2] with a denoting the lattice
Appendix 1. Local and resonance states for a system of bound harmonic oscillators
158
constant) and also due to the linearity of all displacements (Justifying a index-free notation in Eq. (A1.114)), the value G(n,co) to be conveniently calculated at co >
COQ:
nja
cos qna f nm JJ a>2-a>2Dsm2{qa/2) 1
cosnx nm JJ,^ _ J ^ 2 + I W 2
G(K,O>)= —
(-If
CO -COD/2-0}yjt3)
J*J~-co
-0)£)
2
w•V2 D
2
mco
dx CQSx
D
(A1.115) The parenthesized expression is less than unity and hence G(n, co) decreases exponentially with increasing n. The asymptotic (A 1.115) at co» coD appears as
-UL
G(n, a>)»
mco
l+L^D
If <°D 2
1 co
(A1.116)
4co2
Therefore, in the case co » coD, local vibration frequencies are approximately equal to:
co2*
u
\
mc
coD ±
2 m + m(j cor
mc
r
2 N"
m + m^ 4S2C)
col
(A1.117)
where coc =coc(m + mc)/m = k(m + mc)/mmc is the squared frequency of an oscillator with the force constant k and the reduced mass mmcl(m + mc). The second term in Eq. (A 1.117) accounts for a difference between the local vibration frequency for an isolated impurity particle and the value G>Q . The third term represents the contribution from the interaction of impurity particles separated by the distance na, the two signs corresponding to their in-phase and anti-phase vibrations.
Appendix 2 Thermally activated reorientations and tunnel relaxation of orientational states in a phonon field The rotational mobility of adsorbed molecules is caused by its rotational degree of freedom (resulting from the fact that the molecule is tightly bound to the substrate through the only atom) and by the coupling of molecular vibrations with surface atomic vibrations. The rotational motion intensity is strongly temperature-dependent and affects spectroscopic characteristics. As a result, the rotational mobility of surface hydroxyl groups was reliably detected.200"203 Consider reorientations of a diatomic surface group BC (see Fig. A2.1) connected to the substrate thermostat. By a reorientation is meant a transition of the atom C from one to another well of the azimuthal potential U(
Fig. A2.1. Schematic depiction of a diatomic surface group BC.
159
Appendix 2. Thermally activated reorientations and tunnel relaxation
160
Let the minimum of the potential U{cp) be characterized by the angular coordinate value cp = 0 and two nearest maxima be at the points cp = ± cpm. The probability densityfi^co,cp) for a particle to be located at a point with the angular coordinate cp and to have the angular velocity co under thermodynamic equilibrium with a thermostat is given by the Gibbs distribution:
j da -co
\d
where / is the moment of inertia relative to the rotation axis. During the time dt, the barrier can be surmounted at the point 0). The average number of the jumps over the barrier for the time dt is therefore equal to:
dW = dt]dcoaf(a,
^[M
r
k£_
1/2
2nl
(A2.2)
\d
If the reorientation barrier is U((pm) - £/(0) = AU(
y
^C^U"(0)
(A2.3)
-
The average reorientation frequency co can be defined as an average velocity with which a particle leaves a given potential well over two barriers with cp = + cpm: w = IdWIdt. With the cyclic frequency of rotational vibrations in the well defined by Eqs. (A2.2) and (A2.3) as w^[U"(0)/^'\ we arrive at: w = TC~XCO9 exp(- AUp lkBT) .
(A2.4)
Appendix 2. Thermally activated reorientations and tunnel relaxation
161
This expression offers a clear physical interpretation. The preexponential is equal to the frequency of particle's jumps onto the well walls, while the exponent represents the probability for the potential well Af/P to be surmounted on thermal activation. We demonstrate that the spectral function of valence harmonic vibrations of a diatomic group that effects rotational reorientations is broadened by w. The vector of atom C displacements relative to the atom B (see Fig. A2.1) may be represented as x(t)e(t), where x(t) is the change in the length of the valence bond oriented at the time / along the unit vector e(/). Characteristic periods of valence vibrations are much shorter than periods of changes in unit vector orientations. As a consequence, the GF of the displacements defined by Eq. (4.2.1) can be expressed approximately as: G<#(/)
« -Ld(t)([x(t),x(0)}(ea(t)eP(0))
= G„(t)(f
(t)^(0))
.
(A2.5)
Here we first calculate the commutator [x(t)jc(0)] averaged over an equilibrium Gibbs state ensemble with fixed orientations e so as to obtain the GF of a harmonic oscillator: Gxxit) = -(mo)oyl0(t)s'm
o)0t
(A2.6)
and then average it over the unit vectors e(f)Without regard for deformational and rotational vibrations of unit vectors e(t), the qualitative behavior of the time dependence of the correlation function for twodimensional reorientations is describable by the following relation: (ea(t)e^(0^
= ^Sap
exp(-w0, t > 0 .
(A2.7)
For the initial time t = 0, the above formula (A2.7) is identical with the result of averaging over random orientations in a surface plane. In the course of time, the "memory" of the initial orientation fades, the condition t » w'] (w"1 is the average period between reorientations) permitting an independent averaging over e"(t) and e^(0), and the correlation function (A2.7) tends to zero. We substitute (A2.6) and (A2.7) into (A2.5) to obtain the frequency dependence of the spectral function:
162
Appendix 2. Thermally activated reorientations and tunnel relaxation
(
r.
co2 )=--\mGaP(co) 7t 8a/3
00
= 2nma)Q —^— [dte-°>'smcotf sin cot 2mna)ft J (A2.8) cow
mn [co2 - co2 - w2 f + 4co2w2 It is easy to verify that Sa\a?) is normalized by the condition:
Js°0(fl>2) codco = ~Sap. 2m
(A2.9)
Since the average reorientation frequency w is far less than the cyclic frequency of valence vibrations a>o, expression (A2.8), with measured frequency values co close to coo, can approximately be rewritten as:
47rnico0((o-a)0y + wz (the normalization condition (A2.10) differs from (A2.9) in that it has no the cofactor co in the integrand but has an additional cofactor (2con)'] in the right-hand side). The spectral function thus has a Lorentz shape with a halfwidth at the half distribution height equal to the average reorientation frequency w. If expressed in spectroscopic units (cm"1), the halfwidth Av1/2 amounts to caf2nco (c0 designates the velocity of light in vacuo). Ryason and Russel measured the temperature dependence of the IR absorption band halfwidth for valence vibrations of hydroxyl groups on the silica surface.200 At T > 325 K, the least squares method permits a straight line to be drawn through experimental points of the dependence In AvV2 (T1), the equation of the line appearing as follows:200 ln(2Av, /2 ) = -(4.42±0.35)10 2 7 , " 1 +(2.79±0.08).
(A2.ll)
Hence, the activation energy, Aemt, amounts to 38±3 meV and the preexponential, 2AV]/2 (T-*oo) is equal to 16.3±1.3 cm"'. Compare this result with the theoretical estimate (A2.4). If zl^ot is identified with the reorientation barrier AU^, then AvU2
Appendix 2. Thermally activated reorientations and tunnel relaxation
163
(7"-»oo) = « / 2 ^ o =3(Af//2/) l/2 /2^c 0 * 73 cm'1 (with / = 1.4810'48 g-cm2 for a hydroxyl group), this value being nine times as large as the experimental one. At T = 300 K, we have from Eq. (A2.4) that AV\J2 ~ 16.8 cm ', whereas experiments yield Av,/ 2 »2 cm"'. Nevertheless, relation (A2.4) agrees well with experimental characteristics of the Brownian rotational motion of molecules in solid state, when molecular reorientation barriers are several times higher.159'204 It comes as no surprise, because the greater are values At/p, the more applicable is the classical description for reorientations. The criterion here may be represented by the inequality p = AUl/JM(l> = (2IAU9)y2/nA » 1 implying that the height of the barrier A£/p should include a great number of quanta fio&y for rotational vibrations. In this case, it is possible to disregard the discreteness of energy portions derived from the thermostat which are needed for a particle to fluctuationally surmount the barrier AVr For a hydroxyl group on the silica surface, the reduced barrier p is of the order of unity and hence a classical description is no longer applicable. We have already ascertained this by analyzing the empirical relation (A2.11). To refine the estimated value of the energy barrier AU^,200 it was taken into account205 that the activation energy A£-rot should be reckoned not from the well bottom (A£^ot would then be equal to AUf) but from the ground energy level for zero vibrations Ay/2. Since co9 is dependent on AU^, we are led to the following quadratic equation for the reorientation barrier AU9: AErol=AU
(A2.12)
With Aerot «38 meV and / * 1.48-10"40 g • cm2, we obtain AUV * 55 meV, in agreement with the quantum chemical calculation.206 However, the relation (A2.4) remains inapplicable, as the preexponential in it proves to be anomalously high. It is necessary to take proper account of the discreteness of energies transferred to a surface group from the substrate thermostat. If p ~ 1, then the first excited level with the energy ~ Z/ta^jl lies near the potential well top and the quantum transition to it, when activated by the interaction with the substrate phonon thermostat, will enable the atom C to pass freely over the barrier or under a low barrier by tunneling. In this case, the rate of transitions from the ground to the first excited level is expected to be a good estimate for an average reorientation frequency. The probability for a transition to occur between two states per unit time is determined by Fermi's golden rule and depends on the operator of interaction between the subsystem concerned and a thermostat. As orientational states are characterized by a low-energy spectrum, they will be substantially influenced by the
Appendix 2. Thermally activated reorientations and tunnel relaxation
164
interaction with low-frequency vibrational modes of a solid-state matrix. In the isotropic elastic continuum approximation, this interaction is expressible as:207 H
m = Tap (r)"aj3 +mcr-ii,
(A2.13)
where uap is the deformation tensor, Ta^y) represents the deformation potential,208 and r = r c - rB denotes the orientation vector for the group BC whose center-ofmass vibrations are described by the deformation vector u. In what follows, we shall derive a relation like Eq. (A2.13) and for now it is reasonable to elucidate the physical meaning of its terms. The first term describes the change in the energy of the interaction between the atom C and the substrate induced by the substrate deformation. Its order of magnitude may be estimated by the total energy ec of the bonds between the atom C and all nearest-neighboring atoms except the atom B; sc being of the same order as reorientation barriers AUr The second term allows for the d'Alembert force, m c u, in a noninertial system of reference connected with the center of mass for the group BC which undergoes acceleration ii as a solid vibrates. If we designate the energy binding the atom B to a solid-state matrix by eB and the interatomic distance by a, then we have |u| ~ £BlmBa and the second term in Eq. (A2.13) at r ~ a turns out to be of the order (jnclmB)sB. Thus, the ratio of the second to the first term is estimated as (mc/mB)-(£B/£c).lM With the atom C strongly bound not only to B but also to the other atoms of a solid-state matrix (i.e., when ec ~ EB) the above ratio is small in the parameter mclmB « 1, so that the dominant contribution to the interaction with phonons is provided by the deformation potential. Reorientation probabilities were calculated, with the deformation term only taken into consideration, in Refs. 209, 210. For a diatomic group BC, EQ ~ At/,, ~ 0.1 eV, whereas sB ~ 10 eV (a typical bond energy for ionic and covalent crystals). A strong binding of the atom C only to the atom B results in the dominant contribution from inertial forces.2" For OH groups, as an example, the second term in Eq. (A2.13) is more than 6 times as large as the first one. Calculate the probabilities for the transitions between orientational states | a) of the BC group which are dictated by the d'Alembert force. If the group BC is incorporated into the solid-state matrix bulk (as for instance, hydroxyl groups contained in many biological macromolecules212 and amorphous substances213), then the deformation vector u is describable by Eq. (A 1.43) and the second term in Eq. (A2.13) may, in terms of secondary quantization, be written as:
"in! = E ^ V qv
+Zgvbgy).
(A2.14)
Appendix 2. Thermally activated reorientations and tunnel relaxation
165
Here b^v and b^v are creation and destruction operators for phonons of the vth acoustic branch (one longitudinal with v = I and two transverse ones with v = t) with the wave vector q which obey the dispersion laws ft\,(q) =cv|q| (c; and c, are the longitudinal and transverse sound velocities). The operator xqv is defined by the relation ' Xqv
ftm^(q) 2pV
M2
re^(q),
(A2.15)
where p is the medium density, V is the volume of the main area, and e^(q) is the phonon polarization unit vector. Provided the group BC is located on a continuum surface, the deformation vector u is given by Eq. (2.35) for Rayleigh phonons and the interaction operator has the form (A2.14), with the summation performed only over the two-dimensional wave vector q (the index v is absent) and the operator Xq specified as follows:
Xq
AX 2pS
ir\r-~
(A2.16)
The dispersion law for Rayleigh phonons is specified by the relation cq,(q) = cp\q\ (cp is the propagation velocity for Rayleigh waves); ^= c/c, is the positive root of the following equation: (A2.17) the values A and K\ are found from formulae:
i+-
*F?_
*i
(A2.18)
Let us substitute relation (A2.14) into the expression for Fermi's golden rule: W^a=^Y,aXna\H%'a'f5{yEn-En'-^a'a)-
(A2.19)
166
Appendix 2. Thermally activated reorientations and tunnel relaxation
The appearing therein squared matrix element of //>nt' in the basis of phonon thermostat states {nqv} is easily calculable using matrix elements of the second quantization operators: 2
^qv},«|^it)|{«^},«1)
= £|(4vk)
"q A ' ,n.v-\ +(>V + 1 K' ,n.v+\\ •
qv
(A2.20) Therefore, expression (A2.19) assumes the form:
2^^!/ I
I Al2 X-1
ex
Pl-^^(qKvJ
n
(A2.21)
The summation over n\v retains only terms with «'q„ = « q „+ 1. Summing then over nqv, we arrive at the result: -pham
(«)=
\eP^-\Y
n=0
(A2.22)
-ptuon
«=o As a consequence, expression (A2.21) turns into the sum of two quantities, namely, the probabilities for transitions per unit time accompanied by phonon absorption, w
a'i-a ' a n ^ .,(+) a^a
em ss on
i ' > wcT<-a > w n ' c n
2K h2
expressed as:
("«'«) + - + - E ^ k ^ k ) ^a'a-^vCq)) qv
where
are
(A2.23)
Appendix 2. Thermally activated reorientations and tunnel relaxation
{"a'a) = W * k r ' a | / * B 7 , ) - 1 J ~ '
a
a'a = (*«' -^a),h
167
•
(A2.24)
We substitute relation (A2.15) or (A2.16) respectively for bulk or Rayleigh phonons into Eq. (A2.23) to obtain the same result:164
For bulk phonons, K = K\= 1 and for Rayleigh phonons, the values K and K\ are respectively specified by the equation K = Zrc /[8A(2 + y )] and by Eq. (A2.18). The average sound velocity c was introduced in Eq. (A 1.66). The probability for a transition to the first excited state of rotational vibrations to occur per unit time is readily estimated by formula (A2.25) if the harmonic approximation with the matrix element (ij| r =t)/2mccoiQ is invoked:
*sA2o
= *^£?f[exp(»a>io/M')-ll-1 . Anpc
(A2.26)
The comparison of formulae (A2.26) and (A2.4) proves to be informative. The temperature dependence of w is determined by the average number of phonons with the frequency equal to co\0. Such a temperature-dependent cofactor typical of the Bose-Einstein statistics was invoked previously to estimate the reorientation frequency for light bulk solid-state defects. At ficoio« kBT, this cofactor provides a linear dependence w(T) and at the inverse inequality changes to the Boltzmann factor e\p(-^Wio/kBT) with the activation energy ho)\^ (instead of bJJ9 in Eq. (A2.4)). The cofactor of the bracketed expression in Eq. (A2.26) depends on the parameters of a substrate and a diatomic group. Its value for the system OH/Si0 2 (w c = mH= 1.67-l(r24g, ^ 0 = 3.77-1013 s"1 (200 cm-1), p = 2.2 g/cm\ c = 5.6-105 cm/s, K = 1) amounts to 6.9 • 1011 s"1 (3.7 cm"1). At T = 300 K, the probability (A2.26) is found to be w = A A • 10 n s"1 (2.3 cm'1), which is consistent with experimentally measured IR-absorption spectral line halfwidths for valence vibrations of hydroxyl groups on the silica surface. It should be noted that unlike Eq. (A2.4), relation (A2.26) is dependent on a specific mechanism of the subsystem-thermostat interaction. A rigorous treatment of the IR-absorption spectral line broadening for valence vibrations of a reorienting group should include, in addition to reorientational
Appendix 2. Thermally activated reorientations and tunnel relaxation
168
Brownian motion, other mechanisms, as for instance, a decay of a local vibration into substrate phonons (see Chapter 4) or inhomogeneous broadening caused by static shifts of oscillator frequencies in random electric fields of a disordered dipole environment. A temperature dependence of a broadening arising from these additional effects should be considerably weaker than the exponential dependence in Eq. (A2.26) or (A2.4). The total broadening is therefore expressible as AcoU2(T) = const + w(T).
(A2.27)
It behaves like w(T) at high temperatures and becomes a constant at low temperatures, as illustrated in Fig. A2.2. InAVj^lniv
15
•
•
W / o •
\ o\L£^ *
2 • • 3 A A
0,5
\
^e A
1000/T Fig. A2.2. Temperature dependences measured for the halfwidths Avm of the 1R absorption bands for valence vibrations of OH(D) groups on Si0 2 surface (filled markers) and recalculated for the halfwidths w of three components of Lorentzian lines (empty markers) for OH (1) and (OD) groups of high concentration (2), and for (OD) groups of low concentration (3).2"2
Appendix 2. Thermally activated reorientations and tunnel relaxation
169
An additional factor implied by the summand "const" in Eq. (A2.27) is the multicomponent structure of the spectral line which results from the tunnel splitting of the vibrational levels approximating orientational states (see Fig. 4.4). Previously the temperature dependences &vm(T) for OH(D) groups on the Si0 2 surface were measured and converted into w(T) with regard for the relative shift of three components by 2.2 cm'1.202 The results are demonstrated in Fig. A2.2. The plotted curves correspond to the Bose-Einstein temperature factors in Eq. (A2.26) with the value «| 0 = 3.77 • 1013 s"' (fi(0\a ~ 25 meV). The inhomogeneous broadening exhibited in the low-temperature region was discussed in Refs. 61 and 121. At sufficiently low temperatures {kBT « fia)w), the rate of transitions to the first excited state of rotational vibrations (see Eq. (A2.26)) becomes exponentially small, so that the transitions between the lowest tunnel-split energy levels predominate in the reorientation process. To be precise, the term "reorientation" keeps its original meaning no more, because the process under discussion is of purely quantum nature and represents the orientational derealization of the atom C caused by the subbarrier tunneling in the interaction with the thermostat. The process of this kind was termed most accurately in the title of Ref. 214, namely, "tunneling relaxation in a phonon field". It has been shown in the above paper that in a certain low-temperature range, relaxation times are governed by two-phonon transitions. In other words, the transitions between tunnel-split levels are attributed to the interaction which is a quadratic function of atomic displacements in a phonon thermostat. An interaction appearing as expression (A2.13) is a linear function of the displacements u and gives rise to one-phonon transitions. It is therefore necessary to derive a more exact expression that involves, in addition to the terms of Eq. (A2.13), contributions quadratic in u. Let ry (j = 0, 1, 2, ...) and r c respectively denote the positions of heavy atoms bound to each other and the position of the light atom C strongly bound to the atom withy = 0 and weakly bound to the nearest n atoms withy = 1,..., n. In Fig. A2.1, the atom withy = 0 is represented by the atom B and atoms withy = 1,..., n are 0|,..., 0„ (n = 3). The Hamiltonian function for the nuclei of the atoms concerned may be written in the approximation of pah-wise interactions with the atom C:
H^.^pf^cMc-r^i^,.). J
(A2.28)
J
We now turn to the vectors of the inertia center, ?0 = (moro + mCrc)/{m0 and of the relative nuclear positions r = r c - r0. Then Eq. (A2.28) becomes:
+m
c) >
170
Appendix 2. Thermally activated reorientations and tunnel relaxation
mr2
, (mQ+mcfi
*-.*»,-r
//=-— +
+ £^^C0(r)
2
j*o
( m0 r\ + U ro r0-r+m0 + mc \
\ m0+mc
(A2.29)
~—rj|v J
Expanding the sum of interactions UCJ in r (see Eq. (1.56)) yielded, with the nearest n atoms positioned symmetrically, the hindered rotation potential U(
(A2.30)
Now we should make allowance for possible atomic displacements uy from equilibrium positions rj ' and perform an expansion both in uy and r together. In so doing, u0 is considered to mean the displacement of the inertia center from the equilibrium position ffi ':
U ? o — ^ — r,n.- = m0 + mc ) 2 ntn+mr
2>Q
ro"" - /
m0 m0 + mc
>8UCjWj)(..a y>0
dArf
UQ
-U
{
2! f* BArfdArf [ Ar
.=?(0)
(0)
a a ® 00'Ir rP
^—\ /MQ + rriQ
*-i
0j
(A2.31) J
2*-? •"
J
J
= 2>Q(Ary)
a .
™0
„, ^
(A2.32)
m0+mc
m0+mc
X
m
0 + mC
Appendix 2. Thermally activated reorientations and tunnel relaxation
171
Here 3>"^ are the force constants. The first term in Eq. (A2.31) does not contain u; and can be included in the potential C/Co(r). The second term gives, with regard to the equation of motion for u0,
(m0 + m c K - " 2 ° o y " y -
( A23 3)
7*0
the potential energy for the d'Alembert inertia force, mcr-\i0, as in Eq. (A2.13). The third term appears as a standard form of the potential energy for atomic displacements in the harmonic approximation. The subsequent expansion terms involve products of the different powers of the components ra and u" with the largest contribution to two-phonon interactions coming from the terms proportional to rau^uYj . At mc « m0, they are small in the parameter mc/m0. In the expansion (A2.32), the first term is merely a constant, while the second one renormalizes equilibrium atomic positions but gives no contribution to the interaction of the atom C with a thermostat (provided a symmetric disposition of atoms, the term linear in r vanishes). The third term contains small corrections to [/Co(r) and to force constants "r , and the contribution to a single-phonon interaction proportional to ra\u{f -ujj.
It is clear that in the continual
approximation the quantities of such a structure will account for the deformational interaction, Ta^r)uap in Eq. (A2.13). The subsequent expansion terms in Eq. (A2.32) will contribute also to two-phonon interactions proportional to One-phonon interactions are described in general by formula (A2.14). An analogous expression for two-phonon interactions appears as:
"2 "I X I^IS-
+
"V)(V-*-V).
CA2J4)
where B depends on the radius vector r. Substitute Eq. (A2.34) into formula (A2.19) (with Hfal instead of # v j and a' = 2, a= 1) that specifies rates of transitions between the two tunnel-split levels separated by the energy gap Ae = fi(Oi\. In the case Ae « kBT, characteristic phonon energies E„ are much larger than As and hence the term ^ca^ in the argument of ^-function (see Eq. (A2.19)) can be omitted.
172
Appendix 2. Thermally activated reorientations and tunnel relaxation
By calculating matrix elements of phonon operators in (A2.34) and summing over the occupation numbers «q„(as performed in Eqs. (A2.20) and (A2.21)), we get:
\
(
^
^
(
^ )
+
1 > ( M q ) - M q - ) ) . (A2.35)
q v,qV
To estimate the above expression, consider a particular case of the symmetric twowell potential in which the two-phonon interaction H>n{ can approximately be derived from expansion (A2.32): H
™ *^
c o s
4 t f ~«2)M -«f -«f).
(A2.36)
(The axis x goes through atoms withy = 1 , 2 and
(qv,qV)=M^e-(q)e-'(q,)COSJe-'^
_e^2)
, ihMJmAr,z ,2 xlx 3mQrN W:
x(2-e-^2_e^2)^^,M
(A2.37) COS (p.
Substituting Eq. (A2.37) into (A2.35) and taking into consideration that (l|cos q>\2) » 1, we obtain, in the Debye approximation at Ae/ks « T«TD: (2)
25-35-,(8)Al/Xr9
,A23g.
The same temperature dependence for symmetric wells was reported previously (for asymmetric wells, w2i °c T ).214 Compare Eq. (A2.38) with the contribution from one-phonon transitions for which we can apply formula (A2.25) with (/|| r = r2 :
Appendix 2. Thermally activated reorientations and tunnel relaxation
,,.(!).. yy-n
21
2„2
m
~
'A^
Cr 2
~ 2nh pd
3
173
(A2.39)
V "J
With the parameters As = 0.28 meV, AL^ = 50 meV, Ar, = 3r = 3A, c = 5.610 5 cm/s, p = 2.2 g/cm3, expressions (A2.38) and (A2.39) prove to be equal to 104 c"1 (10"7 cm"1) at T = T* ~ 30 K, while expressions (A2.26) and (A2.38) become equal at r= T** ~ 170 K. Thus, at temperatures T < T*, the one-phonon relaxation described by Eq. (A2.39) is dominant, in the temperature range T* < T < T**, the two-phonon relaxation (see Eq. (A2.38)) prevails, and at T > T**, rotational vibrations are excited and the relaxation is of thermoactivational nature as seen from Eq. (A2.26). It is noteworthy that the w\J (and especially w^]) is several orders of magnitude smaller than (o1\ for the atom C with a mass of a hydrogen atom. The dependence of (On = Ael/zon mc is given by the formula Ae*K{n)^22^p^r, id
K{n)=[>
" ^ [3/4-n = 3
(A2.40)
which implies that the value con for a heavy atom C may be several orders smaller and only then it may be equal to w^'. The relaxation times for a two-level subsystem are found which at w2i « a>z\ are equal to 2w2\ for diagonal elements of the density matrix (in the basis of eigenfunctions of the subsystem) and to w2\ for nondiagonal elements that decay oscillating with the frequency co^.61 As was shown before,214 at w2\ » o>n, the subsystem is characterized by quite different relaxation times. For surface atomic groups with a rotational degree of freedom, the inverse inequality, w2\ « a&i, holds and the relaxation times differ only by a factor of two. In this Appendix, we attempted to elucidate the basic features of the transition of a particle from one potential well to another in various special cases: at high reduced barriers p when a classical description is applicable, for p ~ 1 when the probability for a rotational vibration to occur can be regarded as an estimate of the reorientation frequency for not too low temperatures, and, finally, for the lowtemperature situation when subbarrier tunneling relaxation becomes dominant. However, the quantitative description of the processes considered cannot be taken as satisfactory, since it is rather fragmentary and, in addition, no general expression has been derived for an average reorientation frequency which would reduce to Eqs. (A2.4), (A2.26) or to (A2.38), (A2.39) in the above-listed special cases. As of now, an adequate approach has been developed which allows this quantity to be
174
Appendix 2. Thermally activated reorientations and tunnel relaxation
calculated in a wide range of parameter values.215 The subbarrier tunneling of a particle is described by the equations of motion dependent on imaginary time. Solutions of these equations specify possible trajectories for a particle interacting with a thermostat. Then the technique of Feynman integrals over trajectories is invoked. Extremum (action-minimizing) trajectories of a certain kind which correspond to a transition of a particle from one potential well into another are called instantons.216 In quantum chronodynamics, instantons are solutions of the field equations with a nontrivial topology and allow for gluon field fluctuations which cannot be taken into consideration by the perturbation theory. The points discussed in this appendix are of direct relevance to a description of the chemical reactions that involve surmounting the potential barrier. The barrier is surmounted due to thermal activation at high temperatures and due to tunneling at low temperatures, the latter fact accounting for an occurrence of low-temperature chemical reactions.216'217
Appendix 3 The spectral function of local vibration (low-temperature approximation involving collectivized high-frequency and low-frequency modes) In the Heitler-London approximation, with allowance made only for biquadratic anharmonic coupling between collectivized high-frequency and low-frequency modes of a lattice of adsorbed molecules (admolecular lattice), the total Hamiltonian (4.3.1) can be written as a sum of harmonic and anharmonic contributions: Htot-H + HA,
H=^HK,
(A3.1)
K
v
v
(A3.2)
"*=~W
IX.^^CJ^-I^+K,
,
(A3.3)
where the terms in the expression (A3.2) respectively represent high-frequency and low-frequency vibrations of the admolecular lattice, crystal phonons, and the harmonic coupling between phonons and low-frequency modes. In the above relation, quantum states of phonons of the admolecular lattice are characterized by the surface-parallel wave vector K, whereas the quantum numbers of substrate phonons are indicated by the couple of indices K and v. The latter accounts for the polarization of a quasi-particle and its motion in the surface-normal direction; it implicitly reflects both the atomic arrangement in the crystal unit cell and the position of the admolecular lattice, as a whole, relative to the crystal surface. The utility of introduction of the complex index vwas substantiated previously.138 As a rule, the density of states for molecular lattice vibrations is negligible as compared to that for crystal phonons. Therefore, the K-mode of a molecular lattice is coupled with the crystal phonons specified by the same wave vector K. Besides, the low-frequency collective mode Gfc of adsorbed molecules can be considered as a 175
176
Appendix 3. The spectral function of local vibration
resonance vibration with the renormalized frequency <»K and the inverse lifetime 7 7 , ^ ) . 1 3 8 1 4 6 The response of the system concerned to an external electromagnetic field is conveniently described in terms of double-time Green's function (GF) which can be introduced in a variety of representations.144'218"221 In what follows we will involve the representation in Matsubara's frequency space218 which is accepted in the theory of anharmonic crystals197 and provides a number of exact solutions in the case of a single adsorbed molecule.150152 In this approach, the spectral line shape for highfrequency vibrations can be determined as follows:184 L(a>) = M im G(o 0 aJ, co + /0), p = (kBT)~
(A3.4)
K
JG(a04,r)ehwTdT,
G(a04,co) = 1
(A3.5)
T2b(r)2b + (0)S(/?)\
G(a0alr) = - i -
'S-
M
(A3.6)
where X
S(fi)*S(ft,0),
5(r,r 0 ) = Texp
\dT'** AW
1
(A3.7)
T
o
A(r) =
eHTAe-HT,
Sp{e^*. (•••>„
=
Sp\e -ftH
-}
(A3.8)
and A is an arbitrary operator function. Some of lowest-order diagrams for the temperature GF (A3.6) are shown in Fig. A3.1. The dashed and the solid lines represent the GFs of high-frequency and low-frequency vibrations of a planar lattice in the harmonic approximation:
Appendix 3. The spectral function of local vibration
177
.Q.
_0.
Q a
b
Fig. A3.1. Some lowest-order diagrams for the temperature GF (A3.6). The dashed and solid lines correspond to the GF for high-frequency and resonance low-frequency vibrations of a molecular planar lattice in the harmonic approximation (see Eq. (A3.9) and (A3.10)). Each vertex is associated with the factor -y/N, the integration and summation being performed over each vertex coordinates r, from 0 to /?, and over all internal wave vectors K. At phdK »1 , the main contribution is provided by a-type diagrams.184
Gk0)(r) = /£flK(r)«£(0)\ ={[n(QK)+l^{r)+n(QK)9(-r)}e-hn^,
^ W ^ M ^ C O ) )
(A3.9)
= jdcoVlK{a>){[n(a>)+l]d{T)+n(cD)0(-T)}e-hco\ (A3.10)
Appendix 3. The spectral function of local vibration
178
where «(«) = |e / ? *' u -l]" 1
(A3.11)
is the factor of Bose-Einstein statistics and
»K(»)=T
~Tffir~/ff
(A312)
[CO - C0K - P K {C0)[ + [«77K \fD)f
is the resonance function of low-frequency vibrations,138'40
fc^-EM2^-**.,)
~PMA^^.
(A3.13)
The real and imaginary parts of the pole in expression (A3.12) define the renormalized frequency <5K and the inverse lifetime ? 7 K ( ^ K ) = 2 ^ 7 K ( ^ K ) °f the resonance vibration. Let us take advantage of the inequality / ? f i f i K » l permitting neglect of the terms of the order exp(-/#?QK). Then the trace taken over high-frequency and lowfrequency modes, Sp{...}, is reduced to that for low-frequency modes, with all high-frequency vibrations considered only for the ground state. The resulting temperature GF (A3.6) takes the form: G{aQ4,r)
= (a 0 (r)5(r)a 0 + (0)) , fi > r > 0 .
(A3.14)
This expression refers to diagrams without closed high-frequency loops (Fig. A3.1 a). Thus, provided the inequality 0hClK »1 is valid, the GF defined above can be written in the form involving no high-frequency mode operators: G(aoflJ.r) = e-m°T(Sm(r))0,
(A3.15)
Appendix 3. The spectral function of local vibration 00
X
»=/>
0
T
\
*n-\
0
0,
179
K,...K„_,
A
AT
exp<j
IL
/KK' JJ /RK'
00
(A3.16) *KK'(r) =(*(*•))
K
=-~exp[h(QK - Q K ' ) r ) ^l * +
K(r)AKrK'(r).
(A3.17)
Representing average value (A3.15) as (%)(r))0=/(r)
(A3.18)
and differentiating the left and the right sides of expression (A3.18) with respect to T, we arrive at:
Eq. (A3.19) corresponds to compact diagrams in Fig. A3.la. The expression obtained allows the line shape to be determined in the approximation of the small anharmonic coefficient /, as well as in the low-temperature limit. Importantly, the three-time GF
gK,K'k>T'>T")=-
j^-n\\ ( W ))0
(A3 20)
-
satisfies the Dyson-like equation with the assumption of non-interacting highfrequency collective vibrations (QK=^o):
180
Appendix 3. The spectral function of local vibration
hy (A3.21) (the resonance GF gK(0)(r) m the harmonic approximation is determined by the expression (A3.10)). Equation (A3.21) takes into account intermolecular interactions of low-frequency resonance modes and thus generalizes the corresponding equations derived before.150152 To invoke the perturbation theory for a small anharmonic coupling coefficient, we use the Wick theorem for the coupling of the creation and annihilation operators of low-frequency modes in expression (A3.19). Retaining the terms of the orders y and y2, we are led to the following expressions for the shift AQ and the width 2T of the high-frequency vibration spectral line:184 00
AQ = - ^ - £ jdcn(a))mK(o}),
2r
(A3.22)
= 2 ^2 ]T pft)«(«)k« + ^ 0 - % ) + l]^K+K'(^)9?K'(^ + ^ 0 - ^ K X N
K,K'. • 0 0
(A3.23) where n(co) and S R K ^ ) a r e respectively determined by relations (A3.11) and (A3.12). It should be noted that expressions (A3.22) and (A3.23) describe a single Lorentz-like spectral line of local vibrations, with its position and width dictated by dispersion laws and lifetimes for resonance low-frequency modes. The analytical description of high-frequency line shapes becomes possible in the low-temperature limit, i.e., at n(coK) « exp{-(37jcoK}«l, which represents an experimentally important case. In this situation, the Wick coupling for the operators of low-frequency modes in expression (A3.19) involves only the terms in which the annihilation operator is to the left of the creation operator in all but one operator pair. Then Eq. (A3.19) can be written as:
Appendix 3. The spectral function of local vibration
dFJr) _ dr ~
hy N
w
181
00
'
YJ jdamK(<»Ufo)Y,(-flrT
\dTi>T-n)
«=0
*1
0
(A3.24)
"IJ-I
x jdr2,rl - r 2 ) . . \drn(pK{co,Tn_x
-T„),
where
PK(«,r) = - ^ ] r pfl»'SRK.(a>')exp[»(n0 - % - K , K
+
ffl
-'»')f]
-co
(A3.25)
K'
On integrating Eq. (A3.24) over the variables r h r2,..., r„, we obtain:
K
-CO
—CO
(A3.26) CO
Then expression (A3.26) in the Markov approximation (7i77K(ft>)r, hyr» reduced as follows:184 dF(r)_ dr
"K:)
hy
N jf i _ ( y / N ) ^ [ « K K ' + ixfyc. W)+Tk'foc'))]"1
1) can be
(A3.27) '
K'
where «KK' = ^ 0 ~ f i K - K '
+fi,
K -«K'-
(A3.28)
As seen, the spectral line of high-frequency local vibrations is of the Lorentz-like shape:
182
Appendix 3. The spectral function of local vibration
L(CO)=
—Imn
CO-QQ-W
with the shift AQ=Re^and the width 2T=-2 \mW.
hdr
(A3.29)
Bibliography 1. K. Burke, D. C. Langreth, M. Persson, and Z.-Y. Zhang, Phys. Rev. B 47, 15869(1993). {6, 105, 111,112} 2. W. Steele, Chem. Rev. 93, 2355 (1993). {6, 51} 3. D. Marx, H. Wiechert, Adv. Chem. Phys. 95, 213 (1996). {6, 51} 4. A. Patrykiejew, S. Sokolowski, K. Binder, Surface Science Reports 37, 207 (2000). {6,51} 5. R. D. Diehl, M. F. Toney, and S. C. Fain, Jr., Phys. Rev. Lett. 48, 177 (1982). {6,8,51} 6. R. D. Diehl and S. C. Fain, Jr., Phys. Rev. B 26, 4785 (1982). {6, 8} 7. R. Wang, S.-K. Wang, H. Taub, J. C. Newton and H. Shechter, Phys. Rev. B 35,5841(1987). {6,8} 8. S.-K. Wang, J. C. Newton, R. Wang, H. Taub, J. R. Dennison and H. Shechter, Phys. Rev. B 39, 10331 (1989). {6, 8} 9. M. F. Toney and S. C. Fain, Jr., Phys. Rev. B 36, 1248 (1987). {6, 8} 10. K. Morishige, Mol. Phys. 78, 1203 (1993). {6, 8} 11. A. Terlain, Y. Larher, F. Angerand, G. Parette, H. Lauter, I.C. Bassignana, Mol. Phys. 58, 799 (1986). {6, 8} 12. Y. P. Joshi, D. J. Tildesley, J. S. Ayres, R. K. Thomas, Mol. Phys. 65, 991 (1988). {6,8} 1 3 . 0 . Berg and G. E. Ewing, Surf. Sci. 220, 207 (1989). {6, 8, 9, 71} 14. O. Berg, R. Disselkamp and G. Ewing, Surf. Sci. 277, 8 (1992). {6, 8, 9, 71} 15. J. Heidberg, E. Kampshoff, O. Schonekas, H. Stein and H. Weiss, Ber. BunsenGes. Phys. Chem. 94, 112, 118, 127(1990). {6,8,9,71} 16. J. Heidberg, E. Kampshoff, R. KUhnemuth and O. Schonekas,. Surf. Sci. 272, 306(1992). {6, 8,9,71} 17. J. Heidberg, E. Kampshoff, R. KUhnemuth and O. Schonekas,. Surf. Sci. 269/270, 120(1992). {6, 8,9,71} 18. J. Schimmelpfennig, S. Folsch, and M. Henzler, Verh. Deutsch Phys. Ges. 4, 192(1990). {6,8,71} 19. G. Lange, J. P. Toennies, R. Vollmer, and H. Weiss, J. Chem. Phys. 98, 10096 (1993). {6,8,71} 20. J. Heidberg, E. Kampshoff, R. Kuhnemuth, O. Schonekas, M. Suhren, J. Electron. Spect. Related Phenom., 54/55, 945 (1990). {9, 71} 21. W. Chen, W.L. Schaich, Surf. Sci. 220, L733 (1989). {9, 71}
* Pages on which the references appear in the book are indicated in braces. 183
184
Bibliography
22. G. Lange, D. Schmicker, J. P. Toennies, R. Vollmer, and H. Weiss, J. Chem. Phys. 103,2308(1995). {9,71} 23. J. Schimmelpfennig, S. Folsch, M. Henzler, Surf. Sci. 250, 198 (1991). {9, 71} 24. A. Vigiani, G. Cardini, V. Schettino, J. Chem. Phys. 106, 5693 (1997). {9, 71} 25. S. Picaud, S. Briquez, A. Lakhlifi, C. Girardet, J. Chem. Phys. 102, 7229 (1995). {9,71} 26. C. Girardet, S. Picaud, P.N.M. Hoang, Europhys. Lett. 25, 131 (1994). {9, 71} 27. J. Heidberg, K. Stahmer, H. Stein, H. Weiss, M. Folman, Z. Phys. Chem. 755, 223 (1987). {9} 28. J. Heidberg, E. Kampshoff, R. Kuhnemuth, M. Suhren, and H. Weiss, Surf. Sci. 269/270, 128 (1992). {9, 29, 42, 67, 71} 29. R. Disselkamp, H.-C. Chang, G. E. Ewing, Surf. Sci. 240, 193 (1990). {9} 30. D. J. Dai, G. E. Ewing, J. Electron Spectrosc. Relat. Phenom. 64/65, 101 (1993). {9} 31. S. Picaud, P. N. M. Hoang, C. Girardet, A. Meredith, A. J. Stone, Surf. Sci. 294, 149(1993). {9} 32. D. Schmicker, J. P. Toennies, R. Vollmer, H. Weiss, J. Chem. Phys. 95, 9412 (1991). {9} 33. H. Weiss, Surf. Sci. 331-333, 1453 (1995). {9} 34. J. Heidberg, H. Henseler, Surf. Sci. 427, 439 (1999). {9} 35. J. Heidberg, M. Hustedt, J. Oppermann, P. Paszkiewicz, Surf. Sci. 352-354, 447 (1996). {9} 36. J. Heidberg, M. Hustedt, J. Oppermann, P. Paszkiewicz, Z. Phys. Chem. 215, 669(2001). {9} 37. J. Heidberg, D. Meine, B. Redlich, J. Electron Spectrosc. Relat. Phenom. 64/65, 599(1993). {9} 38. V. Panella, J. Suzanne, P. N. M. Hoang, C. Girardet, J. Phys. I (France) 4, 905 (1994). {9} 39. J. Heidberg, M. Kandel, D. Meine, U. Wildt, Surf. Sci. 331-333, 1467 (1995). {9} 40. J. Heidberg, E. Kampshoff, R. Kuhnemuth, O. Schonekas, H. Stein, H. Weiss, Surf. Sci. Lett. 226, L43 (1990). {9} 41. O. Berg, L. Quatrocci, S. K. Dunn, G. E. Ewing, J. Electron. Spect. Related Phenom. 54/55,981 (1990). {9} 42. H. H. Richardson, C. Baumann, G. E. Ewing, Surf. Sci. 785, 15 (1987). {9} 43. H.-C. Chang, H. H. Richardson, G. E. Ewing, J. Chem. Phys. 89, 7561 (1988). {9} 44. C. Noda, G. E. Ewing, Surf. Sci., 240, 181 (1990). {9} 45. R. Disselkamp, H.-C. Chang, G. E. Ewing, Surf. Sci. 240, 193 (1990). {9} 46. V. M. Rozenbaum, Zh. Eksp. Teor. Fiz. 99, 1836 (1991) [Sov. Phys. JETP 72, 1028 (1991)]. {12, 14, 18, 20, 26, 47, 59}
Bibliography
185
47. B. C. Kohin, J. Chem. Phys. 33, 882 (1960). {12, 27, 28, 45} 48. D. A. Goodings and M. Henkelman, Can. J. Phys. 49, 2898 (1971). {12, 27} 49. V. A. Slusarev, Yu. A. Freiman, I. N. Krupskii, and I. A. Burakhovich, Phys. Stat. Sol. B 54, 745 (1972). {12, 27, 28} 50. J. C. Raich and N. S. Gillis, J. Chem. Phys. 66, 846 (1977). {12} 51. F. Malder, G. van Dijk, and A. van der Avoird, Mol. Phys. 39, 407 (1980). {12} 52. V. M. Rozenbaum and S. H. Lin, J. Chem. Phys. 112, 9083 (2000). {12, 27, 121, 124} 53. V. M. Rozenbaum, Ukr. Fiz. Zh. 36, 302 (1991); J. Electron. Spectr. 56, 373 (1991). {13, 14, 18} 54. P. I. Belobrov. R. S. Gekht, and V. A. Ignatchenko, Zh. Eksp. Teor. Fiz. 84, 1097 (1983) [Sov. Phys. JETP 57, 636 (1983)]. {14, 66} 55. J. M. Luttinger and L. Tisza, Phys. Rev. 70, 954 (1946); 72, 257 (1947). {14, 40} 56. V. M. Rozenbaum and V. M. Ogenko, Fiz. Tverd. Tela (Leningrad) 26, 1448 (1984) [Sov. Phys. Solid State 26, 877 (1984)]. {14, 16, 25, 46, 60, 64, 66} 57. V. E. Klimenko, V. V. Kukhtin, V. M. Ogenko, and V. M. Rozenbaum, Ukr. Fiz. Zh.35, 1426(1990). {16} 58. V. E. Klimenko, V. V. Kukhtin, V. M. Ogenko, and V. M. Rozenbaum, Phys. Lett. A 150, 213 (1990). {16} 59. V. M. Rozenbaum, E. V. Artamonova, and V. M. Ogenko, Ukr. Fiz. Zh. 33, 625(1988). {16-18} 60. J. G. Brankov, D. M. Danchev, Physica A 144, 128 (1987). {19} 61. V. M. Ogenko, V. M. Rozenbaum, and A. A. Chuiko, Theory of Vibrations and Reorientations of Surface Groups of Atoms (in Russian), Naukova dumka, Kiev, 1991. {19. 23, 25, 52, 54, 59, 87, 95, 97, 169, 173} 62. D. J. Scalapino, Y. Imry, and P. Pincus, Phys. Rev. B 11, 2042 (1975). {21, 25, 26} 63. M. Takahashi, J. Phys. Soc. Jpn. 50, 1854 (1981). {21, 25, 26} 64. P. I. Belobrov, V. A. Voevodin, and V. A. Ignatchenko, Zh. Eksp. Teor. Fiz. 88, 889 (1985) [Sov. Phys. JETP 61, 522 (1985)]. {21} 65. F. J . Dyson, Commun. Math. Phys. 12, 91 (1969). {21} 66. Yu. M. Malozovskii and V. M. Rozenbaum, Zh. Eksp. Teor. Fiz. 98, 265 (1990) [Sov. Phys. JETP 71, 147 (1990)]. {21, 23, 25, 47} 67. Yu. M. Malozovsky and V. M. Rozenbaum, Physica A 175, 127 (1991). {21, 23-26, 47} 68. A. Z. Patashinskii and V. L. Pokrovskii, Fluctuation Theory of Phase Transitions Pergamon Press, Oxford, 1979. {23} 69. N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966). {23}
186
Bibliography
70. V. M. Rozenbaum, Pis'ma Zh. Eksp. Teor. Fiz. 63, 623 (1996) [JETP Lett. 63, 662 (1996)]; Zh. Eksp. Teor. Fiz. I l l , 669 (1997) [JETP 84, 368 (1997)]. {24, 40,47,66, 125} 71. J. Villain, R. Bidaux, J. P. Carton, and R. Conte, J. Phys. (Paris) 41, 1263, (1980). {24} 72. E. F. Shender, Zh. Eksp. Teor. Fiz. 83, 326 (1982) [Sov. Phys. JETP 56, 178 (1982)]. {24} 73. D. M. Danchev, Physica A 163, 835 (1990). {24} 74. A. F. Sadreev, Phys. Lett. A 115, 193 (1986). {24} 75. V. M. Rozenbaum and V. M. Ogenko, Pis'ma Zh. Eksp. Teor. Fiz. 35, 151 (1982) [JETP Letters 35, 184 (1982)]. {25, 47, 49} 76. S. Romano, Nuovo Cimento D 9, 409 (1987). {26, 47} 77. J. A. Goychuk, V. V. Kukhtin, and E. G. Petrov, Phys. Stat. Sol. B 149, 55 (1988). {26} 78. I. Amdur, E. A. Mason, and J. E. Jordan, J. Chem. Phys. 27, 527 (1957). {27, 29} 79. V. E. Klymenko and V. M. Rozenbaum, J. Chem. Phys. 110, 5978 (1999). {27, 38} 80. I. G. Kaplan, Theory of Molecular Interactions (Elsevier, Amsterdam, 1986). {29} 81. V. M. Rozenbaum, Phys. Lett. A 176, 249 (1993). {29, 31, 62, 66, 67, 71, 121} 82. P. N. M. Hoang, S. Picaud, and C. Girardet, J. Chem. Phys. 105, 8453 (1996). {29} 83. O. G. Mouritsen and A. J. Berlinsky, Phys. Rev. Lett. 48, 181 (1982). {33, 51} 84. A. B. Harris, O. G. Mouritsen, and A. J. Berlinsky, Can. J. Phys. 62, 915 (1984). {33} 85. S. Tang, S. D. Mahanti, and R. K. Kalia, Phys. Rev. B 32, 3148 (1985). {36, 46} 86. H. Y. Choi, A. B. Harris, and E. J. Mele, Phys. Rev. B 40, 3766 (1989). {36, 46} 87. H. Shiba, Solid State Commun. 41, 511 (1982). {36} 88. C. R. Fuselier, N. S. Gillis, and J. C. Raich, So). State Comm. 25, 747 (1978). {36} 89. A. B. Harris and A. J. Berlinsky, Can. J. Phys. 57, 1852 (1979). {37} 90. S. F. O'Shea and M. L. Klein, Chem. Phys. Lett. 66, 381 (1979). {37, 50} 91. D. Marx, O. Opitz, P. Nielaba, and K. Binder, Phys. Rev. Lett. 70, 2908 (1993). {37} 92. F. Y. Hansen and L. W. Bruch, Phys. Rev. B 51, 2515 (1995). {37} 93. B. Bussery and P. E. S. Wormer, J. Chem. Phys. 99, 1230 (1993). {39} 94. A. Wada, H. Kanamori, and S. Iwata, J. Chem. Phys. 109, 9434 (1998). {39}
Bibliography
187
95. V. M. Rozenbaum, A. M. Mebel, and S. H. Lin, Mol. Phys. 99, 1883 (2001). {39} 96. O. Nagai and T. Nakamura, Prog. Theor. Phys. 24, 432 (1960). {40} 97. J. Felsteiner, Phys. Rev. Lett. 15, 1025 (1965). {40} 98. H. Miyagi and T. Nakamura, Prog. Theor. Phys. 37, 641 (1967). {40} 99. J. Heidberg, M. Grunwald, M. Hustedt, F. Traeger, Surf. Sci. 368, 126 (1996). {42,45,62, 121, 123} 100. J. Ashkin and E. Teller, Phys. Rev. 64, 178 (1943). {44} 101. R. J. Baxter, Exactly Solved Models in Statistical Mechanics (Academic Press, London, 1982). {44,48} 102. R. V. Ditzian, J. R. Banavar, G. S. Grest, and L. P. Kadanoff, Phys. Rev. B 22, 2542 (1980). {44} 103. P. J. Wojtowics, Phys. Rev. 116, 32 (1959). {46} 104. T. Tsuzuki, Progr. Theor. Phys. 57, 812 (1977). {46} 105. J. V. Jose, L. P. Kadanoff, S. Kirkpatrick, and D. R. Nelson, Phys. Rev. B 16, 1217(1977). {47} 106. S. Elitzur, R. B. Pearson, and J. Shigemitsu, Phys. Rev. B 19, 3698 (1979). {47} 107. V. L. Berezinskii, Zh. Eksp. Teor. Fiz. 59, 907 (1970); 61, 1144 (1971) [Sov. Phys. JETP 32, 493 (1970); 34, 610 (1971)]. {47} 108. J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 (1973). {47} 109. V. M. Rozenbaum, Phys. Rev. B 53, 6240 (1996). {47, 51, 52, 60, 114, 119} 110. A. N. Morozov, L. G. Grechko, V. M. Ogenko, and V. M. Rozenbaum, Fiz. Nizk. Temp. 18, 1018 (1992) [Sov. J. Low Temp. Phys. 18, 715 (1992); React. Kinet. Catal. Lett. 50, 249 (1993). {47, 51} 111. S.Romano, Phys. Scripta 50, 326(1994). {47} 112. T. T. Chung and J. G. Dash, Surf. Sci. 66, 559 (1977). {51} 113. J. Eckert, W. D. Ellenson, J. B. Hastings and L. Passell, Phys. Rev. Lett. 43, 1329(1979). {51} 114. S. E. Roosevelt and L. W. Bruch, Phys. Rev. B 41, 12236 (1990). {51} 115. C. Graham, J. Pierrus, and R. E. Raab, Mol. Phys. 67, 939 (1989). {51} 116. R. D. Etters, V. Chandrasekharan, E. Uzan,and K. Kobashi, Phys. Rev. B 33, 8615(1986). {51} 117. C. S. Murthy, K. Singer, M. L. Klein, and I. R. McDonald, Mol. Phys. 41, 1387 (1980). {51} 118. Z.-X. Cai, Phys. Rev. B 43, 6163(1991). {51} 119. P. Tarazona and E. Chason, Phys. Rev. B 39, 7157 (1989). {51} 120. F. Y. Wu, Rev. Mod. Phys. 54, 235 (1982). {51} 121. V. M. Rozenbaum, V. M. Ogenko, A. A. Chuiko, Usp. Fiz. Nauk 161, No 10, 79 (1991) [Sov. Phys. Usp. 34, 883 (1991)]. {52, 54, 59, 92, 95, 104, 114, 169}
188
Bibliography
122. V. M. Rozenbaum, Zh. Eksp. Teor. Fiz. 107, 536 (1995) [JETP 80, 289 (1995)]. {52, 56, 60} 123. J. Heidberg, M. Hustedt, E. Kampshoff, V. M. Rozenbaum, Surf. Sci. 427-428, 431 (1999). {58,71,75-77} 124. V. M. Rozenbaum, Phys. Rev. B 51, 1290 (1995). {62} 125. S. Prakash and C. L. Henley, Phys. Rev. B 42, 6574 (1990). {65} 126. I. Lyuksyutov, A. Naumovets and V. Pokrovsky, Two-Dimensional Crystals, Academic Press, New York, 1992. {66} 127. I. G. Chernysh, I. I. Karpov, G. P. Prikhod'ko and V. M. Shai, Physicochemical Properties of Graphite and its Compounds, Naukova Dumka, Kiev, 1990, in Russian. {67} 128. M. L. Dekhtyar and V. M. Rozenbaum, Surf. Sci. 330, 234 (1995). {70, 73} 129. H. Yamada and W. B. Person, J. Chem. Phys. 41, 2478 (1964). {71, 76} 130. W. Kohn and K. M. Lau, Solid State Commun. 18, 553 (1976). {72} 131. T. Yamaguchi, S. Yoshida, and A. Kinbara, J. Opt. Soc. Amer. 64, 1563 (1974). {72} 132. K. Norland, A. Ames, and T. Taylor, Photogr. Sci. Eng. 14, 295 (1970). {73, 74} 133. C. Reich, Photogr. Sci. Eng. 18, 335 (1974). {73} 134. G. G. Dyadyusha and A. D. Kachkovskii, Ukr. Khim. Zh. 41, 1176 (1975) [Sov. Progr. Chem. 41, No 11, 52 (1975)]. {73} 135. G. G. Dyadyusha, V. M. Rozenbaum, and M. L. Dekhtyar, Zh. Eksp. Teor. Fiz. 100, 1051 (1991) [Sov. Phys. JETP 73, 581 (1991)].{73, 83, 105} 136. M. L. Dekhtyar, Dyes and Pigments 28, 261 (1995). {74} 137. E. Kampshoff, PhD. Thesis, Hannover, 1992. {75} 138. V. M. Rozenbaum, Surf. Sci. 398, 28 (1998). {80, 106, 175, 176, 178} 139. I. M. Lifshitz, JETP 17, 1017, 1076 (1947) and 18, 293 (1948); Nuovo Cimento 4,716(1956). {83, 149, 152} 140. V. M. Rozenbaum and S. H. Lin, J. Chem. Phys. 110, 5919 (1999). {84, 107, 109, 111, 116, 178} 141. S. P. Lewis, M. V. Pykhtin, E. J. Mele, A. M. Rappe, J. Chem. Phys. 108, 1157 (1998). {86} 142. M. V. Pykhtin, S. P. Lewis, E. J. Mele, A. M. Rappe, Phys. Rev. Lett. 81, 5940 (1998). {86} 143. B. N J. Persson, E. Tosatti, D. Fuhrmann, G. Witte, Ch. Woll, Phys. Rev. B 59, 11777(1999). {86, 124} 144. D. N. Zubarev, Usp. Fiz. Nauk 71, 71 (1960) [Sov. Phys. Usp. 3, 320 (I960)]. {86, 89} 145. K. Blum, Density Matrix Theory and Application, Plenum Press, New York (1981). {87}
Bibliography
189
146. V. M. Rozenbaum, Zh. Eksp. Teor. Fiz. 102, 1381 (1992) [Sov. Phys. JETP 75, 748 (1992)]. {89, 94, 101, 102, 176} 147. C. B. Harris, R. M. Shelby, and P. A. Cornelius, Phys. Rev. Lett. 38, 1415 (1977). {89, 97, 115} 148. R. M. Shelby, C. B. Harris, and P. A. Cornelius, J. Chem. Phys. 70, 34 (1979). {89,90,98, 115} 149. A. A. Makarov and V. V. Tyakht, Poverkhnost', No. 8, 18 (1988). {89, 91, 92} 150. A. I. Volokitin, Surf. Sci. 224, 359 (1989). {89, 91, 92, 104, 176, 180} 151. B. N. J. Persson and R. Ryberg, Phys. Rev. B 32, 3586 (1985). {90, 99, 103, 104, 115} 152. D. C. Langreth and M. Persson, Phys. Rev. B 43, 1353 (1991). {91, 176, 180} 153. P. Nozieres and C. T. De Dominicis, Phys. Rev. 178, 1097 (1969). {91} 154. V. M. Rozenbaum, Opt. Spektrosk. 71, 453 (1991) [Opt. Spectrosc (USSR) 71, 263 (1991)]. {92, 104} 155. R. K. Tomas, Proc. Roy. Soc. London A 325, 133 (1971). {92, 93} 156. A. V. logansen, Opt. Spektrosk. 63, 1186 (1987) [Opt. Spectrosc (USSR) 63, 701 (1987)]. {93, 104} 157. A. V. logansen and M. Sh. Rozenberg, Zh. Strukt. Khim. 30 (1), 92 (1989) [J. Struct. Chem. (USSR) 30, 76 (1989)]. {93, 104} 158. I. V. Kuz'menko, V. E. Klimenko, and V. M. Rozenbaum, Opt. Spektrosk. 88, 237 (2000) [Opt. Spectrosc 88, 201 (2000)]. {94} 159. A. V. Rakov, Proc. Phys. Inst. Acad. Sci. USSR 27, 111 (1964). {94, 163} 160. H. A. Kramers, Physica 7, 284 (1940). {94, 104} 161. Yu. I. Dakhnovskil, A. A. Ovchinnikov, and M. B. Semenov, Zh. Eksp. Teor.Fiz. 92, 955 (1987) [Sov. Phys. JETP 65, 541 (1987)]. {94} 162. W. Hayes, O. D. Jones, R. J. Elliott, and C. T. Sennett, Proc. Internal. Conf. Lattice Dynamics, Copenhagen 1963 (1965) p. 475. {95} 163. A. A. Maradudin, Solid State Phys. 18, 273 (1966); 19, 1 (1966). {95} 164. V. M. Rozenbaum and V. M. Ogenko, Fiz. Tverd. Tela (Leningrad) 26, 3711 (1984) [Sov. Phys. Solid State 26, 2236 (1984)]. {97, 98, 103, 164} 165. W. Pauli, Festschrift fur A. Sommerfeld, Leipzig (1928) p. 30. {99} 166. B. Hellsing, J. Chem. Phys. 83, 1371 (1985). {103} 167. V. M. Rozenbaum, Opt. Spektrosk. 69, 342 (1990) [Opt. Spectrosc (USSR) 69, 206(1990)]. {104} 168. Z. Y. Zhang and D. C. Langreth, Phys. Rev. Lett. 59, 2211 (1987). {105} 169. W-K. Liu, M. Hayashi, J-C. Lin, H.-C. Chang, S.-H. Lin, and J.-K. Wang, J. Chem. Phys. 106, 5920 (1997). {105, 111,114} 170. R. P. Chin, J. Y. Huang, Y. R. Shen, T. J. Chuang, and H. Seki, Phys. Rev. B 52,5985(1995). {105} 171. J-C. Lin, K.-H. Chen, H.-C. Chang, C.-S. Tsai, C.-E. Lin, and J.-K. Wang, J. Chem. Phys. 105, 3975 (1996). {105, 111, 113}
190
Bibliography
172. M. A. Ivanov, L. B. Kvashina, and M. A. Krivoglaz, Fiz. Tverd. Tela 7, 2047 (1965) [Sov. Phys.-Sol. State 7, 1652 (1966)]. {105} 173. M. A. Ivanov, M. A. Krivoglaz, D. N. Mirlin, and I. I. Reshina, Fiz. Tverd. Tela 8, 192 (1966) [Sov. Phys.-Sol. State 8, 150(1966)]. {105} 174. B. N. J. Persson, F. M. Hoffman, and R. Ryberg, Phys. Rev. B 34, 2266 (1986). {106, 116} 175. V. M. Rozenbaum, Fiz. Tverd. Tela 40, 152 (1998) [Phys. Sol. State 40, 136 (1998)]. {106} 176. B. N. J. Persson and R. Ryberg, Phys. Rev. B 40, 10273 (1989). {111-114} 177. Z. Ye and P. Piercy, Phys. Rev. B 47, 9797 (1993). {111} 178. P. Jakob and B. N. J. Persson, Phys. Rev. B 56, 10644 (1997). {111} 179. P. Jakob, J. Chem. Phys. 108, 5035 (1998). {111} 180. R. Honke, P. Jakob, Y. J. Chabal, A. Dvorak, S. Tausendpfund, W. Stigler, P. Pavone, A. P. Mayer, and U. Schroder, Phys. Rev. B 59, 10996 (1999). {111} 181. H. Ueba, Progress in Surf. Sci. 55, 115 (1997). {111} 182. P. Dumas, Y. J. Chabal, and G. S. Higashi, Phys. Rev. Lett. 65, 1124 (1990). {113} 183. G. D. Mahan and A. A. Lucas, J. Chem. Phys. 68, 1344 (1978). {115} 184. V. M. Rozenbaum and I. V. Kuzmenko, Surf. Rev. Lett. 5, 965 (1998). {116, 120, 176, 177, 180, 181} 185. W. Erley and B. N. J. Persson, Surf. Sci. 218,494(1989). {117, 121, 123} 186. V. M. Rozenbaum and S. H. Lin, Surf. Sci. 452, 67 (2000). {117, 120} 187. I. V. Kuzmenko and V. M. Rozenbaum, Proceedings of the XIV International School-Seminar on Spectroscopy of Molecules and Crystals, Odessa, Ukraine (June 7-12) 1999, p. 200. {121, 123} 188. I. M. Lifshitz, JETP 12, 117 (1942). {149} 189. E. W. Montroil and R. B. Potts, Phys. Rev. 100, 525 (1955). {149} 190. E. W. Montroil and R. B. Potts, Phys. Rev. 102, 72 (1956). {149} 191. A. A. Maradudin, P. Mazur, and E. W. Montroil, Rev. Mod. Phys. 30, 175 (1958). {149} 192. R. L. MOssbauer, Z. Phys. 151, 124 (1958). {149} 193. Yu. M. Kagan and Ya. A. Iosilevskii, Zh. Eksp. Teor Fiz. 42, 259 (1962) [Sov. Phys. JETP 15, 182 (1962)]. {149} 194. R. Brout and W. M. Visscher, Phys. Rev. Lett. 9, 54 (1962). {149} 195. S. Takeno, Progr. Theor. Phys. (Kyoto) 29, 191 (1963). {149} 196. A. A. Maradudin, Theoretical and experimental aspects of the effects of point defects and disorder on the vibrations in crystals, Solid State Phys. 18, 273 (1966); 19, 1(1966). {149} 197. H. Bottger, Principles of the Theory of Lattice Dynamics (Academie-Verlag, Berlin, 1983). {149, 176} 198. M. Wagner, Phys. Rev. 131, 2520 (1963). {149}
Bibliography
191
199. M. Wagner, Phys. Rev. 133, A750 (1964). {149} 200. P. R. Ryasson and B. G. Russel, J. Phys. Chem. 79, 1276 (1975). {159, 162, 163} 201. T. Bernstein, H. Ernst, and D. Freude, Z. phyz. Chem. 262, 1123 (1981). {159} 202. V. M. Rozenbaum and V. M. Ogenko, Khim. Fiz. (in Russian) No 7, 972 (1983). {159, 168, 169} 203. V. M. Rozenbaum, V. M. Ogenko, and A. A. Chuiko, Teor. Eksp. Khim. 20, 219 (1984) [Theor. Exp. Chem. (USSR) 20, 208 (1984)]. {163} 204. D. F. Baisa, T. A. Gavrilko, V. Ya. Panchenko, Ukr. Fiz. Zh. 24, 317 (1979). {163} 205. V. M. Rozenbaum and V. M. Ogenko, Zh. Fiz. Khim. (in Russian) 55, 1885 (1981). {163} 206. A. G. Guzikevich, Teor. Eksp. Khim. 21, 513 (1985) [Theor. Exp. Chem. (USSR) 21, 489 (1985)]. {163} 207. V. M. Kontorovich, Usp. Fiz. Nauk 142, No 2, 265 (1984). {164} 208. J. H. P. Colpa, Physica56, 185 (1971). {164} 209. P. W. Anderson, B. I. Halperin, and C. M. Varma, Phil. Mag. 25, 1 (1972). {164} 210. B. P. Smolyakov and E. P. Khaimovich, Usp. Fiz. Nauk 136, No 2, 317 (1982). {164} 211. N. D. Sokolov, Zh. Eksp. Teor Fiz. 23, 392 (1952). {164} 212. M. V. Volkenshtein, Molecular Biophysics (in Russian), Nauka, Moscow, 1981.(164} 213. W. A. Phillips, Phil. Mag. 43, 747 (1981). {164} 214. Yu. Kagan and L. A. Maksimov, Zh. Eksp. Teor Fiz. 79, 1363 (1980) [Sov. Phys. JETP 52, 688 (1980)]. {169, 172, 173} 215. Yu. I. Dakhnovskii, A. A. Ovchinnikov, and M. B. Semenov, Zh. Eksp. Teor Fiz. 92,955(1987). {174} 216. A. I. Vainshtein, V. I. Zakharov, and V. A. Novikov, Usp. Fiz. Nauk 136, No 4, 553(1982). {174} 217. A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys. 59, 1 (1987). {174} 218. T. Matsubara, Progr. Theor. Phys. 14, 342 (1954). {176} 219. L. P. Kadanoff and G. Baym, Quantum Statistical Mechanics (Benjamin, NewYork, 1962). {176} 220. L. V. Keldysh, Zh. Eksp. Teor. Fiz. 47, 1515 (1964) [Sov. Phys. JETP 20, 1018 (1965)]. {176} 221. D. C. Langreth, Linear and Nonlinear Electron Transport in Solids, eds. J. T. Devreese and V. E. van Doren (Plenum New York, 1976). {176}
Subject index Arrhenius-type temperature dependence [146, 159, 160], {94, 103, 104} Ashkin-Teller model [52, 100-102], {12,44} Biquadratic anharmonic coupling [1, 138, 140, 146-152, 154, 158, 174, 184], {79, 90,116,175} Berezinskii-Kosterlitz-Thouless phase [61, 66-68, 70, 73, 105-109, 121], {47} Bose-Einstein statistics [138-140, 146-150, 164], {79, 116, 167, 178} Bravais lattice (arbitrary) [46. 61,121], {20. 52, 59, 62], Brillouin zone [52, 53, 56-59, 61, 66, 67, 70, 81, 109, 121, 122, 124, 138], {15-17, 20,34,60,65,80,82, 123, 137} Complex lattices [109. 122, 124], {52} Cubic anharmonic coupling [1, 140, 167-173], {79, 104, 105, 107, 110, 116} d'Alambert force [61, 164], {97, 164} Davydov splitting [13-17, 20-45, 52, 81, 99, 122-124, 128, 186, 187], {3, 8-11, 52, 55-59, 61, 62, 67, 71, 73, 121-123} Debye approximation [61, 164, 143, 197], {97, 98, 124, 141, 147, 148, 153, 156} Deformation potential [164, 208-210], {164} Dipole-dipole interactions [1, 52-61, 64, 66, 67, 70, 73-77, 81, 109-111, 121-124, 128], {3, 11, 13-26,53-72, 115} Dispersion interactions [2-4, 47-52, 79, 80, 92-95], {3, 12, 26-29, 38} Dispersion laws [52, 59,61,81, 109, 121, 124], {16, 17,24,52,66, 114, 121, 125} Double-angle technique [83-86], {12, 31-39} Dyson equation [61, 144, 146, 150, 152, 153, 197,218-221], {101, 179} Electric field (influence) [57, 58], {16} Erley-Persson formulae [52, 185-187], {117, 121, 123} Exactly solvable statistical problems [75, 100-102, 146-150, 152, 153], {22, 25, 47, 48,89-91} Exchange dephasing model [138, 140, 146-152, 174-176, 181-187], {4, 79, 89-91, 97, 105, 115, 175} Fermi's golden rule [61], {163} Field functional formalism [74], {24, 25} Finite lifetimes [138, 140-143, 146, 152, 164], {78, 85, 101, 176} Fluctuation-dissipation theorem [61], {149} Fresnel formulae [123, 137], {57, 59} Full width at half maximum (FWHM) [52, 61, 99, 121, 138, 140-143, 146-152, 154-159, 163-187, 193-205], {78, 85, 102-105, 110-123, 156, 160, 162, 167, 168, 180-182} * Brackets and braces indicate bibliography references and pages of the book where the subject items are defined, considered, or used.
192
Subject index
193
Gaitler-London approximation [61, 146], {90, 108, 175} Generating-function method [146, 149], {90, 91} Goldstone mode [53, 61,70, 109, 121], {14,24,47, 125} Green's function method [1, 61, 138, 140-144, 146, 150, 152, 153, 175, 196-199, 218-221], {4, 79, 86-89, 91, 98, 109, 129-133, 139, 142, 145, 151, 176, 177} Ground states [46, 52, 54-61, 64, 77, 81, 85-87, 89, 121, 122, 124, 125], {3, 13-20, 34-43, 54, 55, 63-66} Harmonic coupling [138, 140-143, 146, 149-152, 181-187], {4,78, 127, 175} H-bond molecular complexes [146, 154, 155, 158], {79, 91-94, 104} Heat-capacity method [3, 4, 112], {51} Herringbone structures [83, 84, 86, 88-92, 128], {36, 37, 71, 74} High-frequency vibrations [61, 138, 140, 146-152, 154-158, 164, 184-187], {4, 78, 89,90,104, 120, 175} Hindered-rotation potential [52, 56, 61, 110, 121, 146, 159, 164, 200, 202-206], {25,39,46,79,95, 159} Homogeneous broadening [138, 140-143, 146], {78} Honeycomb lattice [109, 122, 124, 125], {62-67} Infrared light absorption [1-4, 13-17, 20-45, 81, 99, 121-124, 127, 128, 137, 138, 140-144, 146-152, 154-159, 162-164, 167-187, 189-191, 193-206], {9-11, 5659, 68-73, 77-79, 92, 93, 101, 114, 118-120, 122, 123, 139, 154, 155, 162, 176} Inhomogeneous broadening [61, 121], {78, 169} Integral intensity [81, 122, 123, 137, 197], {57, 58, 68} Interactions of nonpolar molecules [47-52, 78-80, 83-86, 88-98, 116, 117], {12, 2734,40,43,49, 124, 125} Interchain interactions [46, 59, 61-63, 66, 67, 110, 121, 128], {18, 21, 26, 69, 7073} Lateral intermolecular interactions [52, 79, 138, 140, 143, 174, 175], {3, 78-86, 106} Lattice-sublattice relations [109, 122, 124], {59, 60} Lattinger-Tisza method [54, 55, 60, 96-98], {14, 19, 40} Local vibrations [61, 138, 140, 146-152, 154-158, 164, 184-187, 196-199], {4, 78, 110, 144, 149} Long-polymethine-chain approximation [128, 134-136], {73, 74} Long-range order in dipole systems [61, 66, 67, 70, 109, 121], {21-24, 47, 66} Long-range-order parameter [56, 57, 61, 66-74, 76, 101, 109-111, 121, 126], {2224, 44-47} Low-energy electron diffraction (LEED) method [2-4, 5, 6, 9], {6, 8} Low-frequency vibrations [138, 140-143, 146, 184-187], {4, 78, 80-86, 90, 91, 120, 123-125, 175} Macro vortex states [64], {21} Markov approximation [146, 149], {4, 79, 86-89, 98, 105, 181}
194
Subject index
Matsubara's frequency space [197, 218], {5, 79, 176} Modulated orientational structures [86], {36} Molecular dynamics method [82, 85, 92], {29, 37} Monoatomic simple lattice [61], {135-141} Monolayers BTCC/AgBr [128, 132, 133], {73} CO/MgO(100) [38, 39], {9} CO/NaCl(100) [20, 27-32, 41-45, 52, 81, 82, 99, 128], {4, 9, 28, 29, 31, 42, 44, 45,62,69,70,71,79,121-125} CO/Ni(100) [1, 152, 177], {6, 111} CO/Ni(lll)[151], {6, 104} CO/Pt(lll)[l, 152, 176, 177], {6, 112} CO/Ru(001)[178, 179], {6, 111} CO2/CsF(100) [34], {9} C02/Graphite [10], {6, 8, 38} CO2/MgO(100) [26, 37, 38], {9} CO2/NaCl(100) [13-19, 20-26, 40, 123], {3, 6-9, 38, 69, 71, 73-76} CS2/Graphite [11], {6,8,38} C 2 N 2 /Graphite[ll], {6, 8, 38} H(D)/C(111) [140, 169-171], {4, 79, 113} H/Si(lll) [140, 180-182], {4, 113, 114} NH3/NaCl(100) [35, 36], {9} N2O/NaCl(100) [33], {9} N2/Graphite [5-8, 83, 84, 88-92, 112, 113], {6, 8, 36-38, 51} 02/Graphite [9], {6-8, 38, 39} OH(D)/Si02 [61, 121, 164, 167, 200-203, 205, 206], {4, 104, 105, 159, 162, 164, 167, 168} SO2/CsF(100) [34], {9} Monte Carlo method [3, 76, 83,85,90,91, 110, 111, 118], {26,37,47,51} Morse potential [61, 140, 176], {113} N2 dimer [79, 94, 95], {39} Neutron-diffraction method [2-4, 7, 8, 113],, {6, 8, 51} 0 2 dimer [79,93,95], {38, 39} Ordering due to disorder [70-72], {24,40,42} Orientational vibrations [52, 53, 61, 121], {14, 22, 24, 125} Pauli equation [146, 165], {4, 79, 98, 99, 105} Phase diagram for a square lattice [52, 85], {35, 41, 44} Phase diagram for a triangular lattice [86, 89], {37} Planar antiferromagnet [87], {36} Polarization vectors [122, 123, 128], {68, 73}
Subject index
195
Quadrupole interactions [2-4, 47-52, 79, 80, 82-86, 88-98, 114-118], {3, 12, 26-29, 36-45,48-51,125,126} Quasicontinuous spectrum [61, 138-140, 146, 188-191], {78, 83-86, 145-149} Quasi-dipole form [52], {11,12, 26-39} Quasinormal orientations [52], {3, 31, 69} Rates of chemical reactions [146, 160, 161, 166, 215-217], {94, 174} Rayleigh phonons [61, 164], {165, 167} Rectangular lattice [77, 59,61, 121], {19,26} Reduced density matrix [61, 145, 146], {87} Renormalization group analysis [68, 119], {51} Reorientation barriers [61, 121, 146, 159, 164, 204], {79, 94} Repulsive interactions [2-4, 47-52, 78, 92-95], {3, 12, 26-29} Resonance (quasilocal) vibrations [138, 140-143, 146, 149, 150-152, 174-187, 193197], {4, 78, 80-86, 116, 120, 123, 144, 148, 154} Resonance function [138,140], {85, 110, 178} Rhombic lattice [60, 61], {19, 38} Rotational degrees of freedom [61, 146, 164, 121], {78, 79, 95, 159} Screening effect [61, 130, 131], {72,73} Self-consistent field approximation [56, 61-64, 66, 67, 85, 86, 110, 121], {21, 25, 46, 47} Spectral function [61, 121], {91, 101, 139, 154, 161, 162} Spectral line shape [61, 121, 138, 140-143, 146-152, 154-157, 164, 174-187, 196205], {3, 78, 93, 95} Spherical model [73, 101], {24} Square lattice [52-54, 56-59, 61-64, 66-68, 70, 75, 76, 79, 85, 99, 107-109], {15-17, 24, 25, 34, 35,38-51, 60, 61, 121-126} Statistical operator [61, 145, 146], {87} Subbarrier tunneling [61, 161], {79, 174} fert-butanol [146, 156, 157], {93} Thermally activated reorientations [61, 121, 146, 159, 164, 200-204], {79, 94, 103, 159-169} Thermodynamic fluctuations [46, 61, 66-68, 70-73, 105-109, 121], {21-24, 66} Torsional vibrations [61, 121], {94} Triangular lattice [53, 56, 59, 61, 66, 67, 73, 83, 84, 86, 89-92, 109, 110, 121], {16, 22, 23, 25, 36-38, 48, 51, 62, 64, 65, 114-120} Tunnel relaxation in a phonon field [61,214], {169, 173} Tunnel splitting [61, 121, 146], {95, 96, 169} Two bound oscillators [61], {133-135} X-ray diffraction method [2-4, 10], {6, 8}
Author index Amdur, I. [78] Ames, A [132] Anderson, P.W. [209] Angerand, F. [11] Artamonova, E.V. [59] Ashkin, J. [100] Ayres, J.S. [12] Baisa, D.F. [204] Banavar, J.R. [102] Bassignana, I.C. [11] Baumann, C. [42] Baxter, R.J. [101] Baym,G. [219] Belobrov, P.I. [54, 64] Berezinskii, V.L. [107] Berg, O. [13, 14, 41] Berlinsky, A.J. [83, 84, 89] Bernstein, T. [201] Bidaux, R. [71] Binder, K. [4, 91] Blum, K. [145] BOttger, H. [197] Brankov, J.G. [60] Briquez, S. [25] Brout, R[194] Bruch, L. W. [92, 114] Burakhovich, I. A. [49] Burke, K [1] Bussery, B [93] Cai,Z.-X. [118] Cardini, G [24] Carton, J.P. [71] Chabal, Y.J. [180, 182] Chakravarty, S. [217] Chandrasekharan, V [116] Chang, H.-C. [29, 43, 45, 169, 171]
Chen, K.-H. [171] Chen, W. [21] Chernysh, I.G. [127] Chin, R.P. [ 170] Choi, H.Y. [86] Chuang, T. J. [170] Chuiko,A.A. [61, 121,203] Chung, T.T. [112] Colpa, J.H.P. [208] Conte, R. [71] Cornelius, P. A. [147, 148] Dai, D.J. [30] Dakhnovskii, Yu.l. [161, 215] Danchev, D.M. [60, 73] Dash, J.G. [112] DeDominicis, C.T. [153] Dekhtyar, M.L. [128, 135, 136] Dennison, J.R. [8] Diehl, R.D. [5, 6] Disselkamp, R. [14, 29, 45] Ditzian, R.V. [102] Dorsey,A.T. [217] Dumas, P. [182] Dunn, S.K. [41] Dvorak, A. [180] Dyadyusha, G.G. [134, 135] Dyson, F.J. [65] Eckert, J. [113] Elitzur, S. [106] Ellenson, W.D. [113] Elliott, R.J. [162] Erley, W. [185] Ernst, H. [201] Etters,R.D. [116] Ewing, G.E. [13, 14, 29, 30, 42-45] Fain, Jr., S.C. [5, 6, 9]
* Brackets indicate corresponding bibliography mentioned see in Bibliography).
(pages of the book on which the references are 196
Author index
Felsteiner, J. [97] Fisher, M.P.A. [217] Folman, M. [27] Folsch, S. [18,23] Freiman, Yu.A. [49] Freude, D. [201] Fuhrmann, D. [143] Fuselier, C.R. [88] Garg, A.J. [217] Gavrilko, T.A. [204] Gekht, R.S. [54] Gillis, N.S. [50, 88] Girardet,C. [25,26, 31,38, 82] Goodings, D.A. [48] Goychuk, J.A. [77] Graham, C. [115] Grechko,L.G. [110] Grest, G.S. [102] Grunwald, M. [99] Guzikevich, A.G. [206] Halperin, B.I. [209] Hansen, F.Y. [92] Harris, A.B. [84, 86, 89] Harris, C.B. [147, 148] Hastings, J.B. [113] Hayashi, M. [169] Heidberg, J. [15-17, 20, 27, 28, 3437,39,40,99, 123] Hellsing, B. [166] Henkelman, M. [48] Henley, C.L. [125] Henseler, H. [34] Henzler.M. [18,23] Higashi, G.S. [182] Highes, W. [162] Hoang, P.N.M. [26, 31, 38, 82] Hoffman, F.M. [ 174] Honke,R. [180] Huang, J.Y. [170] Hustedt, M. [35, 36, 99, 123] Ignatchenko, V.A. [54, 64]
197
Imry, Y. [62] Iogansen, A.V. [156, 157] Iosilevskii, Ya.A. [193] Ivanov, M.A. [172, 173] Iwata, S. [94] Jakob, P. [178-180] Jones, O.D. [ 162] Jordan, J.E. [78] Jose, J.V. [105] Joshi, Y.P. [12] Kachkovskii, A.D. [134] Kadanoff, L.P. [102, 105,219] Kagan, Yu.M. [193,214] Kalia, R.K. [85] Kampshoff, E. [15-17, 20, 28, 40, 123, 137] Kanamori, H. [94] Kandel, M. [39] Kaplan, I.G. [80] Karpov, I.I. [127] Keldysh, L.V. [220] Khaimovich, E.P. [210] Kinbara,A. [131] Kirkpatrick, S. [105] Klein, M.L. [90, 117] Klimenko, V.E. [57, 58, 158] Klymenko, V.E. [79] Kobashi, K. [116] Kohin, B.C. [47] Kohn, W. [130] Kontorovich, V.M. [207] Kosterlitz, J.M. [108] Kramers, H.A. [160] Krivoglaz, M.A. [172, 173] Krupskii, I.N. [49] Kuhnemuth, R. [16, 17, 20, 28, 40] Kukhtin, V.V. [57, 58, 77] Kuzmenko, I.V. [158, 184, 187] Kvashina, L.B. [172] Lakhlifi, A. [25] Lange.G. [19,22]
198
Langreth, D.C. [1, 168, 152, 221] Larher,Y. [11] Lau, K.M. [130] Lauter, H. [11] Leggett, A.J. [217] Lewis, S.P. [141, 142] Lifshitz, I.M. [139, 188] Lin,C.-E. [171] Lin, J.-C. [169, 171] Lin, S.-H. [52, 95, 140, 169, 186] Liu, W.-K. [169] Lucas, A.A. [ 183] Luttinger, J.M. [55] Lyuksyutov, I. [126] Mahan,G.D. [183] Mahanti, S.D. [85] Makarov, A.A. [149] Maksimov, L.A. [214] Malder, F. [51] Malozovskii, Yu.M. [66, 67] Maradudin, A.A. [163, 191, 196] Marx, D. [3, 91] Mason, E.A. [78] Matsubara, T. [218] Mayer, A.P. [180] Mazur, P. [191] McDonald, I.R. [117] Mebel, A.M. [95] Meine, D. [37, 39] Mele, E.J. [86, 141, 142] Meredith, A. [31] Mermin, N.D. [69] Mirlin, D.N. [173] Miyagi, H. [98] Montroll, E.W. [189-191] Morishige, K. [10] Morozov.A.N. [110] MOssbauer, R.L. [192] Mouritsen, O.G. [83, 84] Murthy,C.S. [117] Nagai, O. [96]
Author index
Nakamura, T. [96, 98] Naumovets, A. [126] Nelson, D.R. [105] Newton, J.C. [7, 8] Nielaba, P. [91] Noda, C. [44] Norland, K.[ 132] Novikov.V.A. [216] Nozieres, P. [153] Ogenko, V.M. [56-59, 61, 75, 110, 121,164,202,203,205] Opitz, O. [91] Oppermann, J. [35, 36] O'Shea, S.F [90] Ovchinnikov, A.A. [161, 215] Panchenko, V.Ya. [204] Panella, V. [38] Parette,G. [11] Passed, L. [113] Paszkiewicz, P. [35, 36] Patashinskii, A.Z. [68] Patrykiejew, A. [4] Pauli, W. [165] Pearson, R.B. [106] Person, W.B. [129] Persson, B.N.J. [143, 151, 174, 176, 178, 185] Persson, M. [1, 152] Petrov, E.G. [77] Phillips, W.A [213] Picaud,S. [25,26,31,82] Piercy,P. [177] Pierrus, J. [115] Pincus, P. [62] Pokrovskii, V.L. [68] Potts, R.B. [189, 190] Prakash, S. [125] Prikhod'ko, G.P. [127] Pykhtin, M.V. [141, 142] Quatrocci, L. [41] Raab, R.E. [115]
Author index
Raich, J.C. [50, 88] Rakov.A.V. [159] Rappe,A.M. [141, 142] Redlich, B. [37] Reich, C. [133] Reshina, I.I. [173] Richardson, H.H. [42,43] Romano, S. [76, 111] Roosevelt, S.E. [114] Rozenbaum, V.M. [46, 52, 53, 56-59, 61,66,67,70,75,79,81,95, 109, 110, 121-124, 128,135,138, 140, 146, 154, 158, 164, 167, 175, 184, 186, 187,202,203] Rozenberg, M.Sh. [157] Russel, B.G. [200] Ryasson, P.R. [200] Ryberg,R. [151, 174, 176] Sadreev, A.F. [74] Scalapino, D.J. [62] Schaich, W.L. [21] Schettino, V. [24] Schimmelpfennig, J. [18, 23] Schmicker, D. [22, 32] SchSnekas, O. [15-17, 20, 40] Schroder, U.[ 180] Seki,H. [170] Semenov, M.B. [161,215] Sennett, C.T. [162] Shai, V.M. [127] Shechter, H. [7, 8] Shelby, R.M. [147, 148] Shen,Y.R. [170] Shender, E.F. [72] Shiba, H. [87] Shigemitsu, J.E. [106] Singer, K. [117] Slusarev, V.A. [49] Smolyakov, B.P. [210] Sokolov,N.D. [211] Sokolowski, S. [4]
199
Stahmer, K. [27] Steele, W. [2] Stein, H. [15,27,40] Stigler, W. [180] Stone, A.J. [31] Suhren, M. [20, 28] Suzanne, J. [38] Takahashi, M. [63] Takeno, S. [195] Tang, S [85] Tarazona, P. [119] Taub, H. [7, 8] Tausendpfund, S. [180] Taylor, T. [132] Teller, E. [100] Terlain,A. [11] Thomas, R.K. [12] Thouless, D.J. [108] Tildesley, D.J. [12] Tisza, L. [55] Toennies, J.P. [19,22,32] Tomas, R.K. [155] Toney, M.F. [5, 9] Tosatti, E. [143] Traeger, F. [99] Tsai, C.-S. [171] Tsuzuki, T. [104] Tyakht, V.V. [149] Ueba, H. [181] Uzan, E. [116] Vainshtein, A.I. [216] van der Avoird, A. [51 ] vanDijk, G. [51] Varma, CM. [209] Vigiani, A. [24] Villain, J. [71] Visscher, W.M. [194] Voevodin, V.A. [64] Volkenshtein, M.V. [212] Vollmer, R. [19,22,32] Volokitin, A.I. [150]
200
Wada, A. [94] Wagner, H. [69] Wagner, M. [198, 199] Wang, J.-K. [169, 171] Wang, R. [7, 8] Wang, S.-K. [7, 8] Weiss, H. [15, 19, 22, 27, 28, 32, 40] Wiechert, H. [3] Wildt, U. [39] Witte,G. [143] Wojtowics, P.J. [103]
Author index Woll, Ch. [143] Wormer, P.E.S. [93] Wu, F.Y. [120] Yamada.H. [129] Yamaguchi, T. [131] Ye, Z. [177] Yoshida, S. [131] Zakharov, V.I. [216] Zhang, Z.-Y. [1,168] Zubarev, D.N. [144] Zwerger, W. [217]
Spectroscopy and Dynamics of Orientationally Structured Ad so rbates ^7>VT£~> Z^S
his book provides a detailed and rigorous
presentation of the spectroscopy and dynamics of
orientationally structured adsorbates. It is intended largely for specialists and graduate students in solid state theory and surface physics. To make the book readable also for beginners in surface science, a lucid style is used and a wealth of references on orientational surface structures and vibrational excitations in them is offered. The book is supplemented with two indices (alphabetical listing of subjects and authors, as well as cross-references) which will enable the reader to easily access the information both on principal concepts involved and on specific adsorbate compositions.
m
www. worldscientific. com 5110 he
j