• Aucun résultat trouvé

Orientational disorder in plastic molecular crystals - III. — Infrared and Raman spectroscopy of internal modes

N/A
N/A
Protected

Academic year: 2021

Partager "Orientational disorder in plastic molecular crystals - III. — Infrared and Raman spectroscopy of internal modes"

Copied!
16
0
0

Texte intégral

(1)

HAL Id: jpa-00209583

https://hal.archives-ouvertes.fr/jpa-00209583

Submitted on 1 Jan 1983

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Orientational disorder in plastic molecular crystals - III.

- Infrared and Raman spectroscopy of internal modes

M. Yvinec, R.M. Pick

To cite this version:

M. Yvinec, R.M. Pick. Orientational disorder in plastic molecular crystals - III. - Infrared and Raman spectroscopy of internal modes. Journal de Physique, 1983, 44 (2), pp.169-183.

�10.1051/jphys:01983004402016900�. �jpa-00209583�

(2)

Orientational disorder in plastic molecular crystals

III.

2014

Infrared and Raman spectroscopy of internal modes

M. Yvinec and R. M. Pick

Département de Recherches Physiques (*), Université P. et M. Curie, 4, place Jussieu, 75230 Paris Cedex 05, France

(Reçu le 23 juin 1982, accepté le 14 octobre 1982)

Résumé.

2014

La spectroscopie Raman et infrarouge des modes internes dans les cristaux moléculaires à désordre d’orientation est analysée dans le formalisme, précédemment développé [5, 6], des fonctions de base adaptées

aux

symétries du site et de la molécule. Dans certains

cas

la contribution rotationnelle

aux

profils Raman et infra-

rouge des modes internes peut être déconvoluée de la largeur purement vibrationnelle

ou

d’autres effets dus à des

couplages variés. La partie purement rotationnelle du profil spectral est ici analysée à l’aide d’un petit nombre

de fonctions d’autocorrélations rotationnelles indépendantes du point de

vue

de la symétrie. On montre de plus

que l’intensité intégrée des raies internes permet de déterminer expérimentalement les premiers coefficients indé-

pendants de la densité de probabilité d’orientation des molécules.

Abstract.

2014

The formalism of symmetry adapted functions for molecular orientations introduced previously [5, 6]

is here applied to the analysis of the Raman and infrared spectroscopy of internal modes in orientationally disor-

dered molecular crystals. In

some

favorable cases, the contribution to the Raman and infrared lineshapes arising

from the rotational dynamics of the molecules

can

be disentangled from other contributions arising from the

vibrational lifetime

or

various coupling effects. Here, the rotational lineshapes

are

analysed in terms of independent, symmetry adapted rotational self-correlation functions. Furthermore, it is shown that the integrated intensity of

internal modes provides

a

measurement of the first symmetry independent coefficients in the development of the

orientational probability density function.

Classification

Physics Abstracts

02.20

-

63.50

-

78.30

1. Introduction.

-

The initials ODIC (Orientational

Disorder In Crystals) are presently widely used to

name those high temperature phases of molecular

crystals in which the centres of mass of the molecules still form a regular crystal lattice while their orienta- tions display some degree of disorder. This structural disorder is usually related to a complex dynamics in

which the molecules can perform. large amplitude

reorientations as well as librational motions around

some preferred orientations. Recently many experi-

mental techniques such as NMR, X-ray or neutron scattering, optical infrared and Raman spectroscopy have been used to probe the rotational dynamics of

the molecules in the plastic phases [1, 3]. Here, we shall focus on the Raman and infrared spectroscopy of internal molecular modes. Indeed, in ODIC phases,

the internal molecular vibrations often give rise to

broadened infrared and Raman lineshapes which are

at least partly due to the rotational freedom of the molecules [2, 4]. This paper merely intends to use the

formalism of symmetry adapted functions (SAF), lar- gely developed in two previous papers [5, 6] to analyse

the rotational information which can be extracted from the Raman and infrared lineshapes of internal modes.

In fact, this paper is mainly the continuation of those two previous papers [5, 6] which hereafter will be referred to as (I) and (II) respectively. In (I), we intro-

duced a canonical basis for the functions of the mole- cular orientations. This basis is made of symmetry adapted functions (SAF) which transform according

to some irreducible representations of both the site group and the molecular symmetry group. In (I) this

basis was used to give a complete and unique deve- lopment of the static orientational probability density function, PO(Q), (in short, p.d.f) which takes into account, a priori, the simplifications arising from the

molecular and the site symmetries. In (II), this for-

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01983004402016900

(3)

malism was extended to describe the rotational dyna-

mics of molecules in terms of independent symmetry

adapted rotational correlation functions. This was

used to show how experimental data arising from

neutron scattering experiments could be analysed.

In this last paper of the series, we intend to apply the

same method to Raman and infrared spectroscopy.

This implies to express Raman and infrared line-

shapes in terms of appropriate rotational correlation functions. This approach was first introduced by

Gordon [7] and since, it has been widely extended in the framework of molecular liquids [8, 9]. The SAF

formalism allows to extend those previous results to

the case of anisotropic crystalline sites. In summary, it allows to make precise which ones of the independent

rotational correlation functions are measured through

each spectral component of each internal molecular mode depending on both the site symmetry and the symmetry of the internal molecular mode.

The plan of this paper is as follows : part 2 is devoted

to generalities about Raman and infrared spectroscopy of internal modes in molecular crystals. It intends to

stress the basic assumptions which are to be made, if one wishes to extract rotational informations from

lineshapes which generally involve many other effects such as vibrational relaxation, vibration-rotation

coupling, Coriolis coupling, etc. Part 3 applies the

formalism of the symmetry adapted functions to the expression of the rotational correlation functions relevant for Raman and infrared scattering. In part 4,

we show that some information about the statistical distribution of molecular orientations can be obtained from the measurement of integrated intensity of

Raman internal lines. At last, we analyse in part 5 the rotational information coming out from Raman

and infrared lineshapes.

2. Generalities on Raman and infrared spectroscopy of internal modes. The various approximations.

-

We shall not, of course, develop here the general theory of interaction between light and matter but simply start from the basic formulae which express the Raman scattered intensity and the coefficient of infrared absorption in terms of appropriate correla-

tion functions.

The first one relates the Raman intensity to the

fluctuations of the macroscopic polarizability of the

material sample :

In (2 .1 a) :

R((o) is the ratio of the scattered intensity with a frequency shift W = Wd - Wi to the incident intensity,

kd is the length of the scattered wave vector, ei and ed are the polarization vectors of respectively

the incident light and the analysed scattered light,

E(t) is the total polarizability tensor of the scattering

volume at time t,

and the symbol > stands for an average over a thermal equilibrium distribution of the target states.

The second basic formula relates the infrared absorp-

tion to the fluctuations of the dipole moment :

with

IR(w) is the fraction of incident energy absorbed per unit thickness of the sample,

w and ei are the frequency and the polarization

vector of the incident light,

11(t) is the total dipole moment in the volume V of the sample at the time t.

In the case of molecular crystals, the total polariza- bility tensor and dipole moment of the sample can be

written as a sum of molecular contributions

FormVIae (2 .1 ) and (2.2) are still quite general in the

sense that they provide the Raman and infrared spec- tral intensity over the whole frequency range and must,

therefore, include the whole time dependence of the polarizability tensor EL(t) and dipole moment IIL(t).

More precisely, these quantities depend on the one

hand on the external (translational and reorientational) dynamics of the molecule which gives rise to the low frequency part of the spectra (either purely rotational

or including a « collision induced » translational con-

tribution). On the other hand, the molecular polariza- bility and dipole moment reflect the deformation of the molecule through its various internal vibrational modes. Each internal mode j’ gives rise (if this mode is, in the ODIC phase, sufficiently decoupled from all

the other internal modes) to a spectral line, the inten- sity and lineshape of which is the main subject of our study here. Usually the internal lines are well apart from the low frequency part of the spectra and we shall assume that there is no overlapping between spectral lines arising from different internal modes.

With such an hypothesis, the Raman and infrared

lineshapes corresponding to a given internal mode j’

arise from contributions to the polarizability tensor

and dipole moment which can be written :

where QLj’,,’(t) are the normal coordinates related to

the mode j’ of the molecule L,

(4)

n’ labels the different partner coordinates if j’ is a degenerate mode and ELj’,,’(t) and - are the

derivative tensors :

Inserting (2. 3) into (2. 2) and (2 .1 ) we finally get the basic formulae which give the Raman and infrared

lineshapes related to a molecular internal mode :

At this stage, it is clear that Raman and infrared lineshapes depend on complicated correlation functions which involve both the dynamics of the internal mode and the external dynamics of the molecules. In fact, the exter-

nal dynamics of the molecules appears twice in the formulae (2 , 5a) and (2.5b) : once, indirectly, through its

influence on the vibrational dynamics (the time evolution of the internal coordinate QLj’n,(t) may depend on

the relative orientation and position of the Lth molecule with respect to its neighbours) and once directly, through its coupling to the detection process (the derivative tensors ELj’,,’(t) and "Lj’,,’(t) depend on the mole-

cular orientations and positions). To extract some information about the rotational dynamics of the molecules

we have to disentangle which contribution to the lineshape is due to the rotational motions and which one

arises from non rotational effects such as purely vibrational relaxation or inhomogeneous broadening. We

are thus lead to make two main approximations.

1) The first one concerns the detection mechanism and is known as the Kastler Rousset hypothesis. It

assumes that, in the molecular axes, the individual Raman and infrared tensors ELj,,,, and °Lj’n’ have definite components (for instance cartesian component . and 7r’,) depending only on the ( j’, n’) considered vibra- tion. In particular, these components are unaffected by the dynamics of all the surrounding molecules and are

time independent. Thus, the components of these tensors in the crystal axes depend on time only through the

molecular reorientational motion : as usual, the orientation of the molecule L at time t is described through

the rotation OL(t) which brings the crystal axes (x, y, z) in coincidence with the molecular fixed axes (xL, yL, zL).

In the crystal axes, the cartesian components (E Lj’n’(t) "b and II"Lj’n’(t)) of the derivative polarizability and dipole

moment derive from the molecular fixed axis components (e Jlfl and nj:n’) through the following rotational

relation :

.

...-.nJ. f." ...-.nJ. .. v

where M(OL(t)) is the usual rotation matrix associated with the three Euler angles corresponding to the rota-

tion QL in the crystal axes.

2) The second hypothesis, which we shall call the decoupling approximation, concerns the dynamics of

the crystal. It assumes that the rotational motions and the vibrational dynamics are statistically independent

so that the correlation functions appearing in (2. 5) can be written as a product of a purely rotational corre-

lation function by a vibrational one which leads to

This approximation is, of course, a very crude one and it is perhaps useful to make a little more precise what

it really implies for the dynamics of the molecules. In a molecular crystal, it is usually reasonable to consider

that the dynamics of the j’ internal mode of a given molecule couples only with the same mode of the neigh-

bouring molecules.’ Thus, for a given orientational configuration, we can write a potential energy for each set

{ QLj’.’ } of normal coordinates j’ under the following form

(5)

Wj’ is the frequency of the mode j’ for an isolated molecule and the coupling term V L iZ({ q }), which, a priori, depends on the orientation of all the molecules, may include a static part and a fluctuating one.

The decoupling approximation assumes that the fluctuations of the coupling terms are unimportant either

because their amplitude is negligible (in the weak coupling approximation) or because they are too fast compared

to the correlation time of vibrational coordinates (in the case of motional narrowing).

In part 4, we shall see that these two approximations (namely the Kastler Roussel hypothesis and the decoupling approximation) are sufficient to derive some information about the static aspect of the orientational disorder (measurement of the first coefficients of PO(Q)) from the integrated spectral intensity of internal modes.

In part 5, we shall furthermore assume that there is no coupling between the internal modes of different molecules, i.e.

in that case, the decoupling approximation simply assumes that the local molecular field VLL(T S2 }) does not depend much on the molecular orientations, so that the vibrational correlation function can be written as

Up to now, we have neglected the effect of Coriolis coupling which may be important if the mode j’ is

in the vector representation of the molecular group. Up to first order, Coriolis forces induce a coupling between

the (n, n’) degenerate coordinates of the mode j’, which is proportional to the angular velocity of the molecule.

As it will be clear later, the Raman and infrared detection process decouples the normal coordinate n and n’

and thus, Coriolis effect can simply be taken into account in the vibrational correlation function ovib

Before discussing these points, it is useful to simplify expressions (2.6) and (2.7). This is the purpose of part 3.

3. Application of group theory to the Raman and infrared spectroscopy of molecular crystals.

-

It is well-

known that Raman and infrared experiments yield different spectra according to the polarization of incident and scattered light. If we introduce cartesian components relative to the crystal axes, formulae (2.1) giving

the Raman and infrared intensity can be rewritten as

Of course, all the cartesian components Rab,cd(W) and IRab(w) of the Raman and infrared tensors are not inde-

pendent and the number of different spectra which can be obtained depends on the symmetry group of the

crystal. Our purpose here is to show that in the case of molecular crystals, the group theory analysis can be

made at the same time for both the site and the molecular symmetries. This approach allows to specify the

different independent informations resulting from Raman and infrared lineshapes corresponding to each internal molecular mode. In the following, we shall assume, for simplicity, that we are dealing with a molecular crystal

with one molecule per unit cell so that there is no distinction to be made between the site symmetry group and the crystal symmetry group.

a) First, it is convenient to define the irreducible components of a tensor with respect to the crystal and

molecular symmetry group. Let us assume that we have a tensorial object T defined for instance by its cartesian

components Tab in the crystal axes. It is well-known that this tensor can be split into its spherical components

(6)

T7 which, under the effect of a rotation of the axis system, transforms according to the irreducible representation Di of the rotation group. This splitting involves a projection operator i such that

(see notes (1) and (’)).

Now, through a unitary transformation a applied to the spherical components TJJ’, we can define new

tensorial components Tf which transform according to the irreducible representations of the crystal group

This unitary transformation a has been used in (I) to define, from the spherical harmonics YJJ’, the site symmetry harmonics Yf. As in this paper, the Greek letter A is a shortened notation for a composite index

r is the name of an irreducible representation of the group S,

11 labels the independent r subspaces if this representation is included more than once in the I manifold,

p numbers the different partners for a degenerate representation.

Thus, from (3. 5) and (3. 3) we can write the decomposition of the tensor T into irreducible tensorial compo- nents of the crystal group under the form

In the same way, if t is a molecular tensor with cartesian components, fA’b’ in the molecular axis, we can

define its spherical components t"

and, through the unitary transformation introduced in (I), we can define the irreducible components t1’ with

respect to the molecular symmetry group

In full analogy with (3. 6a), the Greek letter h’ is a composite index relative to the molecular group

From (3. 5b) and (3. 3b) the decomposition of t into molecular irreducible tensorial components writes

b) The analysis of Raman and infrared spectra into independent components with respect to the crystal

group is now straightforward. Indeed, the crystal polarizability tensor E and the dipole moment II can be split

into crystal irreducible tensorial components El (1

=

0, 1

=

2) and IIIa (I

=

1), which from (3 .1 ) and (3 . 2) leads to

(1) We follow, here, the definition of irreducible spherical

tensorial components given by Rose [10]. This implies that if

a

tensorial object T has the spherical component TML in

a

given (x, y, z) axis system, the spherical component T ML

of T in

a new

axis system (x’, y’, z’) obtained from the old

one (x, y, z) through the rotation Q

=

(a, fl, y) (defined in (x, y, z))

are

where D"’(0)

=

D7m’(ex, f3, y)

are

the Wigner function

as

given e.g. by Rose [10].

(1) The formula (3. 3) should in fact include

an

auxiliary

index to label the independent sets of spherical components which belong to the same I representation of the rotation group.

In this article, however,

we are

only dealing with dipole

moments and polarizability tensors which

are

first rank tensor and second rank symmetric tensors respectively. A

first rank tensor projects only

once on

the manifold I

=

1,

and

a

second rank symmetric tensor has two sets of spherical

components,

one

for 1

=

0 and

one

for 1

=

2.

(7)

for Raman intensity (with h, 12

=

0, 2) and to

for infrared absorption (with /1

=

12

=

I).

The correlation functions of irreducible ..tensorial components appearing on the right hand side of (3. 9a

and b) have to reflect the symmetry of the crystal, so that they must be all zero except when, the two implied components À1

I =

(F 1, p 1, p 1) and ).2

=

(T 2, /"21 P2) are contravariant which means that they are equivalent

partners (p 1 = p2) in conjugate representation (T

1 =

F*, 2 or T

1 =

F2 if the representations are real), in which

case the result is independent of the degeneracy index (pl

=

p2)

Thus, one independent spectral component arises for each possible association of two conjugate representations

of the crystal group.

Let us take, for instance, the well-known example of a cubic crystal (symmetry group Oh). The isotropic part (I

=

0 component) of the polarizability tensor is of course in the A1g representation of Oh, while the anisotropic

part (1 = 2 component) splits into the two representations Eg + F2g. This, altogether, give three independent

Raman spectra RA , RE and RF2.’ In the case of a liquid, the spherical components (1

=

0 and I

=

2) are already

irreducible components, which yield only two independent Raman spectra :

c) If now we turn towards the molecular aspect of the crystal, we are led to make explicit the crystal irre-

ducible components (ELj’Il"t(t) and 17 Lj’?i’,l ’(t)) of the individual molecular derivative tensors E Lj’?i’ and "Lj’Il"

This can be done easily by first going through the spherical components of these tensors in the crystals axis

Then, making use of the Kastler Rousset approximation, we can derive the spherical components in the crystal

axes from the components in molecular axes as a function of the rotation QL(t) which characterize the orienta- tion of the L molecule with respect to the crystal axes

In the molecular axes, the spherical components are related to the molecular components through :

Thus, using the definition given in (I) for the canonical basis of symmetry adapted function for molecular orien- tations

we obtain from (3.13, 14, 15)

(8)

At last, the molecular irreducible components Gj’n’ ,r and 1tj’n’ ,r are quite simple. Indeed, for a given vibrational

mode j’, the derivative tensors Ej’n’ and 1tj’n’ have only irreducible components which belong to the representation 0393’j of the mode j’ in the molecular group. Thus, if a

=

(r’, 11’, p’) we can write

The constants g j, and 7r ’l characterize the Raman and infrared activity of the vibrational mode j’.

(03BC’)

For a given j’ mode, the number of infrared constants

nr(/1;) is equal to the number of r J, representations includ-

ed in the I

=

1 tensorial subspace (if zero, the mode j’ is simply inactive for infrared absorption). In the same way,

,

the number of Raman constants eil is equal to the number of r;. representations present in the 1

=

0 and 1

=

2 subspaces.

For instance, table III of paper (I) shows that for a tetrahedral (Td) molecule, the Raman active modes are

the modes of symmetry A1, E and F2. The derivative polarizability for an Al mode has one isotropic component (I

=

0), while the tensors corresponding to the mode E or F2 have one anisotropic (I

=

2) component. Only the F2

modes are infrared active through one I

=

1 component. In the case of a C3v molecule, the A, mode has two

Raman constants, one for 1

=

0 and one for 1

=

2, while the E modes are Raman active through two independent

l

=

2 constants. The A1 and E modes are also infrared active with one constant each.

,

Finally, inserting (3.16) into (3.9) and making use of the decoupling approximation (3.7), we obtain the

independent Raman and infrared spectral components under the following form :

In the formalism, the rotational contribution to the Raman and infrared lineshapes of internal mode appears in terms of correlations between functions of the molecular orientations, ð ’V" (Q L( t»), which are just the symmetry

adapted functions introduced in (I). This formulation is going to be used throughout the next two parts where it appears to be very useful especially in the case of high molecular and site symmetry.

4. The integrated intensity of Raman internal line : measurement of the p.df. P o(Q).

-

One of the most

important features in the description of plastic phases is the statistical distribution of molecular orientations.

It can range between two extreme cases : a) an isotropic distribution of molecular orientations and b) a distri-

bution of the molecules among a finite set of possible orientations. As usual, we call p.d.f. (probability density function) the function PO(Q) which describes the probability density, for a molecule, to have the orientation Q.

In (I), we used the canonical basis { AÅ/’(Q) }, adapted to the site and molecular symmetry, to write down an

explicit development of the p.d.f. P o(Q)

This development allows, a priori, the symmetry of the molecule as well as the site symmetry to be taken into

account and yields a reduced set of relevant and non redundant coefficients (A ’0’6), the values of which characte-

rize the amount of anisotropy present in the statistical distribution of molecular orientations.

(9)

Let us simply recall that those coefficients are divided into two classes :

(i) Coefficients of the first kind : A for which the indices h§ and AO+ belong to the identity represen- tation (To , and To+) of respectively the site and the molecular group.

(ii) Coefficients of the second kind : A "0-’0- ; they exist only when both the site (S) and molecular (A)

groups contain improper rotations, in which case the indices Ào and Ao belong respectively to the second representation (r 0- and to ) of S and fl which induces the identity representation of the subgroups Sr (of S)

and Ar (of m) which contain only pure rotations.

The purpose of this part is to show how the measurement of the integrated intensity of Raman internal lines

provides a numerical estimate of some of the first kind coefficients in the development (4.1) of PO(O). Indeed,

if we neglect the variation of the scattered wave number, kd, over the frequency range of a given internal line,

we obtain, by integrating (3.18a)

Now, we shall assume that the instantaneous correlation between internal modes of different molecules vanishes so that

where n(wj’) is the Bose factor for the mean frequency (oil of the mode j’ in the plastic phase. The relation (4. 3)

is obviously valid if there is no coupling between the internal modes of different molecules. However, this appro- ximation can be justified in much more general cases. Indeed, since the dephasing processes arising from the

fluctuations of the coupling potential written in (2. 8) and (2. 9) play no role in the correlation values at t

=

0, the relation (4.3) is a valid approximation as long as the dispersion of the vibrational frequencies wj, arising

from the coupling is not too large so that the Bose factor n(wj’) can be considered as a constant over the whole Wj’ frequency range.

where ,1’1

=

(r JI, ,u i, n’) A2

=

(r JI, 11;, n’)

N is the number of scattering molecules.

In 4. 4 the thermal average on the right hand side involves only one molecule and thus depends only on the p.d.£ Po(Q)

Introducing the development (4.1) of PO(Q) into (4. 5), we get

where the projection coefficients can be expressed in terms of the usual Clebsh-

Gordan coefficients (lm ) 1 11 12 m1 m2)

(10)

Finally, inserting (4. 6) into (4. 4), we obtain the integrated intensity of each independent spectral component of an internal mode as a linear combination of the coefficient A ’l" of PO(Q)

(4. 9) is still a very complicated expression, but it can be simplified with the help of the two following remarks : First, the indices À’1 and À2 refer to the same representation rJ, of the molecular group; this implies that,

in the generalized Clebsh coefficient

,

1’ refers to the identity representation of the

molecular group. Therefore, in any case, the integrated Raman intensity depends only on the coefficient of the first kind A ).o+fó+ of Po(S2).

Second, there is a more important selection rule. Indeed, the polarizability tensors are second rank symmetric

tensors which implies that the two indices 11 and 12 in (4. 9) are restricted to the value 0 or 2. As the product of the

symmetry adapted functions A).l’l(Q) A).;2(Q) projects only on the subset of function {A).’;" (Q) with ) li - 12 I I h + /2 }, formula (4.9) involves only the coefficients A ).,;.’ of Po(S2) with 1 4. Therefore,

Raman scattering can provide information only on the coefficients of PO(Q) of the first kind with 1 4.

The above treatment applies as well to the integrated infrared absorption lines. However the last selection rule implies that infrared absorption involves only the coefficients A ).0; ).ó+ of PO(Q) with 1 2, which is not

very interesting since in many cases this subset includes only the first coefficient Ao, the value of which is fixed

by the normalization of PO(Q) :

In principle, in the general case, (4.9) is not very useful because neither the constants

eil Y’ characterizing

the Raman activity of the mode j’ nor their relative values are known. However, in many cases, two or more 03BC’

independent spectra depend on the same constants Ej, ( i) which allows to eliminate them. Let us, for instance consider the example of tetrahedral (TJ molecules in a cubic (0h) site. We know from (I) that, in this case, the

p.d.f. p oCQ) has only one non trivial coefficient A4 for 1 :( 4 :

Each vibrational mode E or F2 of the Td molecule is Raman active through a traceless Raman tensor deter-

mined by one constant e(l

=

2) which gives rise to two independent spectral components REg and RF2g in the

crystal group. For each mode, the constant (1

=

2) is eliminated by forming the ratio

g g

If we take the usual definition of the spectral components RE and RF2g, and apply (4.8) and (4.9) to this

case we obtain :

for E vibrational modes and

for F2 vibrational modes.

(11)

This method has recently been used to determine, in the plastic phase of potassium perchlorate, KCI04,

the unique coefficient A A,gA as a function of the temperature [11]. It has also be used, at room temperature, in adamantane C 1 oH 16 [ 12], where it was found that the value obtained for this unique coefficient was- the same

for various internal modes and also agreed with its value determined by neutron scattering [ 13].

In the case of a linear molecule, the p.d.f. is only a function of the polar and azimuthal angles (0, CP) of the C°°

molecular axis and can be developed on the site symmetry harmonics Si (cf. appendix A)

with

In the case of a site with cubic symmetry, the first non trivial term arises for 1

=

4, and the above technique applied to the Z + stretching mode of a linear molecule gives [ 14] :

This technique has been used to measure the a4 coefficient for the high-temperature phase of KCN [ 15] and the

obtained value is again in agreement with the corresponding neutron measurements [ 16].

5. The rotational contribution to Raman and infrared lineshapes.

-

As already stated in part 2, in general,

the Raman and infrared detection processes involve, the vibrational and orientational dynamics of the molecules in a complicated mixed way. However, things become simpler if, besides the Kastler Rousset hypothesis and the decoupling approximations, we can further assume that there is no coupling between the internal modes of different molecules (3). Indeed, under these conditions, inserting (2.11) into (3.18a) we find the independent

Raman spectra under the form

where

with

ø j!b(t) being defined by (2.11).

Of course, an equivalent formula holds for infrared lineshapes, if we replace the Raman constants by the infrared ones In other words, the Raman and infrared lineshapes

appear as a convolution product between a vibrational contribution and a purely rotational one.

Owing to the decoupling hypothesis between the internal modes of different molecules, the rotational

bandshapes (4.3) are given in terms of the

«

self-correlation functions At/Àí*(QL(O») 12 .(QL(t)) > which

involve only one molecule. Thus, Raman scattering and infrared absorption act as incoherent detection pro- cesses, which, like incoherent neutron scattering, only reflect the individual aspect of the reorientational dyna-

mics. In paper (II), this individual rotational dynamics has been described through the s.o.d.f (self-orientational density function) P(Ql Q2’ t) which is the joint probability for a molecule to have the orientation Q, at a time

(3) Note that this approximation is stronger than that leading to (4. 3). Indeed, the latter, which refers to

a

coupling

at t = 0, implies the neglecting of the time fluctuations of

the second term of the r.h.s. of (2.8). The approximation

used here neglects this second term

as a

whole.

(12)

tl and the orientation Q2 at the later time t2

=

t

1

+ t (cf. (II)). Of course, this function can be developed on

the canonical SAF basis (4) :

and it is obvious to identify the coefficients B of this development with the self-orientational correlation func- tions appearing in (5.3); indeed :

Furthermore, it was shown in (II) that, owing to the site and molecular symmetry, the development (5.4)

includes only a small number of non trivial and independent coefficients, which implies that there is also a small number of independent symmetry adapted self-correlation functions. To make things more precise we recall

here the conclusions obtained in (II) :

If we set

(with the usual meaning of these notations),

(i) One independent coefficient of the first kind arises for each possible association of mutually contravariant

representations for both the site group and the molecular group

(ii) Furthermore, when both the site and the. molecular groups include improper rotations, coefficients of the second kind arise from the product of« anticontravariant » representations for both groups :

In (5.3), the molecular indices À; and A’ refer to the same real representation rj, of the mode j’. Therefore,

each independent component J j’t1lf: of the rotational spectrum is a linear combination of the Fourier transform

self-correlation coefficients BFF" (W) which belongs to the correct symmetry of the molecu- lar and site groups :

(4) We have found convenient to take, in (5.4),

a

defini-

tion slightly different from the

one

used in (II), where

,&A2A(Q 12 )* appears instead of A-12 12’(Q2 ). As

a

result, the

selection rules given in (5.7) and (5.8)

are

not absolutely

identical to those given in that paper.

where Vj’ is the degeneracy of the j’ vibrational mode.

This result is very similar to the result obtained in (II)

where the incoherent part of neutron scattering data

was analysed with the same formalism. However,

Raman scattering (resp. infrared absorption) measures

(13)

only the independent self-correlation functions for

11 and 12 equal to 0 or 2 (resp. 11

=

12

=

1), while

neutron scattering implies no selection rules on the I values (5, 6).

Under the previous approximations (Kastler

Rousset i.e. rigid infrared and Raman tensors, decou-

pling approximation and absence of vibrational

coupling between various molecules), the usefulness of (5.1) for measuring the various correlations func- tions BÀI;í Àj;2(t) contained in a given band profile

is limited by our knowledge of the vibrational contri- bution C,ib(W) (Eq. (5 . 2)). In the case of long enough

residence times, a sufficiently good approximation might be to assume that the main vibrational contri- bution consists in an inhomogeneous broadening.

The latter can be estimated from the band shape (or

the site splitting) in the low temperature ordered

phase,. This proved to be a reasonable hypothesis in

the case of the plastic phase of neopentane [17] for

instance. A limiting case of this approximation is of

course C%?(m)

=

6((o - w j’) which implies the neglect

of both inhomogeneous broadening and vibrational relaxation effects.

Assuming the possibility of the deconvolution process implied by (5. 1), the previous analysis of

infrared and Raman data is especially useful when the linear combination (5.9) reduces to one single term.

Then each independent spectral component, once deconvoluted from the vibrational linewidth, is just proportional to one independent self-correlation function. This is for example the case when Td mole-

(5) It is to be noted that the independent rotational self- correlation function B03BBl103BB03BB 03BB03BB2(t)

1

involves the selection rules

i

03B4n1n2 and ðn’ln2’

Thus, the Raman and infrared detection processes involve each normal component n’ of

a

degenerate mode indepen- dently from its partners

even

when these partners coordi-

nates

are

dynamically coupled through e.g. Coriolis effect.

(6) The above analysis (5. 7) is valid for any type of site symmetry and, in particular is in full agreement with the results given by Steele [9] for the rotational lineshapes in

molecular liquids. Indeed, in the

case

of

a

liquid the site indices Àl and À2

are

just the spherical indices ml and m2 and if

we

keep dealing with the spherical components of the molecular derivative tensors

we

get the isotropic (1

=

0)

and anisotropic (l

=

1 (IR), 1

=

2 (Raman)) components of spectral lineshapes under the following form :

where 6Q,(t)

=

Q,(O) - 1 Q,(t) is the reorientation under- gone by the molecule at time t.

cules reside in an Oh site. From table I in (II), we know

that there are four independent correlation functions for 11

=

12

=

2, namely : B + EgE B-, B+ F2gE and B+22g22. The two independent spectral components

(Eg and F2g) of a vibrational mode of symmetry E

are respectively proportional to B I EgE and B + 22gi,

while the two components of an F2 vibrational mode

give B + EgF2 and B+ 22g22· A precise analysis of these

four independent functions may be a first step for the

building of realistic reorientation models in plastic crystals.

The selectiveness of Raman and infrared scattering

may be particularly useful in some cases where it allows to clearly distinguish the contributions arising

from different kinds of molecular motions. Let us for

example consider the stretching (E +) mode of the linear CN- ions in NaCN [14] (cf. appendix A). This

mode gives rise, in the cubic site of the CN- ion,

to three independent spectra A 1 g, Eg, F2g. The A 1 g

spectrum derives from the (1

=

0) component of the Raman tensor and is independent of rotational motion ; it thus gives an estimate of the effect of vibrational relaxation of the Raman lineshapes. The

E spectrum measures the correlation S Eg, 1 (0)

S2."(t) > of the function

which, in the case of NaCN, appears to be extremal for the six preferred [100] orientations of the CN- ions.

Thus, the Eg spectrum does not see the small ampli-

tude librations of CN- ions around the [100] orien-

tations and can be used to evaluate the dynamics of the large amplitude reorientations between preferred

orientations. On the other hand, the F2g spectrum

involves both the reorientation and librational mo-

tions of the CN- ions, and if their residence time in a

preferred orientation is long enough, it is dominated,

at large enough frequency, by the librations.

6. Concluding remarks.

-

The object of this paper has been to show how it is possible, through infrared

and Raman spectroscopy of the internal modes of the molecules of an ODIC single crystal, to obtain infor- mation on the orientation of these molecules. More

precisely, we have discussed how some independent

coefficients on the p.d.f. P oCQ) could be derived from the integrated intensity of these modes, and how,

in a similar way, some independent coefficients of the s.o.d.f. P(ol, Q2, t) could be obtained from the study

of the band profiles. We have shown that the obtention of this information relies on a certain number of

approximations : the Kastler Rousset hypothesis of

Raman and infrared tensors rigidly fixed to the mole-

cules is one of them. More important is the role played by the coupling between the vibrations of different

molecules, and the influence of the molecular field

on the frequency of the internal modes of a single

molecule; we have given the conditions under which

(14)

some coefficients of P(Q) and P(Qll Q2, t) could be obtained, and shown that the restrictions in the latter

case are more stringent than in the former. The problem

of extracting rotational information, in the case when

the vibrational coupling between neighbouring mole-

cules cannot be neglected, has not been discussed here,

and has not yet received a proprer answer in plastic crystals, even when the vibrational-rotational cou-

pling is neglected. Further studies are clearly needed

to clarify this important aspect of the problem. Finally,

in all this paper, the population life time TB (18) of

the mode j’ has been considered to be infinite; more precisely we have assumed that Arm T

1

> 1, where Arm is the HWHM due to both the rotational and the

inhomogeneous broadening caused by the local

molecular field. This should not be an important restriction, as this condition seems to be very frequently

fulfilled.

The same type of analysis as the one presented in

this paper had been done in (II) for neutron scattering data, and it would be easily extended to the analysis

of data arising from various other experimental techniques. In each case (see (II)), simplifying assumptions will be necessary in order to extract the

same information, but these approximations differ

from one technique to the other. Thus the simulta-

neous use of various techniques is necessary in order to ascertain the validity of the obtained results.

Appendix A.

-

THE CASE OF LINEAR MOLECULES IN A PLASTIC CRYSTAL.

-

The symmetry group of a linear molecule is either D ooh or Coov depending on the

existence of an inversion center. The irreducible

representations of these groups are given [19] through

the following character tables I and II.

Table I.

-

Character table for the Coov group.

From (I) and (II), we know that, in the case of a linear molecule, the SAF writes

where £

=

(T, U, p) refers to the site symmetry.

(In the case of a liquid, the SAF are just the Wigner

function D ’T’ (Q).)

The development of the orientational p.d.f., Po(Q)

then takes the form :

where (0, 0) are the polar and azimuthal angles of the Coo molecular axis (cf. (II), appendix B) while the development of the s.o.d.f., P(Oll Q2, t) is

The selection rule arising from the Coo molecular symmetry implies that ml + m2

=

0, which insures that

P(Ql’ Q2’ t) does not depend on each angle q1 and q2 but only on the difference cp

1 -

q2.

In particular, in the case of a molecular liquid we find :

with C1 ’(t)

=

Bm,m’ - m I m’ (t).

As can be seen from A. 3 (and A. 4b), P(Q l’ Q2’ t) does not depend, a priori, only on the polar and azimuthal

angles of the molecular axis at time 0 and t, but on the full three Euler angles, as it involves AÀ;"(Q) with m # 0

and Dmm’(Q). We shall find, in fact, that through Raman and infrared spectroscopy, correlation functions involv-

ing m

=

0 and «m

=

1 » are detectable : in this specific case, light scattering and light absorption by internal

modes turn out to be somewhat superior to incoherent neutron spectroscopy which can give information on m

=

0 correlation functions.

One can show [ 18] that the internal vibrational modes are either stretching modes, (which belong to the Z + representation of Coov, or the Lg+ and Eu representations of D ooh) or bending modes (belonging to the n(Coov),

Tig or 1ru(Dooh) representations), The stretching modes Z + and Eg are Raman active with Raman tensors which

project on eg and E2O ; the Z + and Lu+ are infrared active, with infrared components p?. But for the bending modes,

the 7T and ng modes have anisotropic Raman components E2 while the 7r and 1ru modes have infrared components

Ill.

(15)

Table II.

-

Character table for the D ooh group.

Direct applications of (5.9) and of its infrared counter part then show that the stretching modes provide

information on the m

=

0 coefficients of (A. 3) and (A. 4b) while the bending modes allow us to detect the m

=

1 ones. Consequently the stretching modes do not give us informations on the third Euler angle, which

is normal, as a stretching mode leaves the molecule with the Coo symmetry. On the other hand, the bending modes

are sensitive to the rotational dynamics of the bending plane, which means that the third Euler angle is a relevant quantity which can be measured.

Let us take the example of a linear molecule sitting on a cubic site. A stretching mode gives rise, in Raman scattering, to an Alg(l

=

0) spectrum, and to two 1

=

2 spectra, Eg and a F2g, both sensitive to the reorientation processes. Their lineshapes are respectively proportional to the Fourier transform of the following correlation functions

where, for instance

If the stretching mode is infrared active, there is only one independent absorption spectrum, of symmetry Flu, proportional to the Fourier transform of

where

A Raman active bending mode, on the other hand, will give rise to only two 1

=

2 spectra, respectively proportional to the Fourier transform of

while an infrared active one gives rise to one I

=

1 absorption spectrum proportional to the Fourier transform of

Références

Documents relatifs

Nuclear magnetic relaxation Tl longitudinal relaxation times measurement allows the evaluation of correlation times related to molecular reorientation.. If the nucleus under

It must be stressed that in the case of symmetry groups containing improper rotations, the development of Po([Q]) involves basis functions which are not in the unity

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Using the frequency red-shift provided by the stimulated Raman effect in high pressure gaseous hydrogen, the tunability of a Nd : YAG pumped dye laser has been

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

The following chapter addresses mechanisms of internal appetite control, starting by discussing intuitive eating, a recently introduced concept that describes eating in response

By using the selection rules for the CaF2 structure and a combination of the neutron scattering data and theoretical phonon dispersion curves, we were able to

La diffusion Raman se calcule alors par la théorie des perturbations du premier ordre dépendant du temps, en prenant pour hamiltonien perturbateur les