• Aucun résultat trouvé

Isotropic and anisotropic collision-induced light scattering by gaseous sulfur hexafluoride at the frequency region of the ν<sub>1</sub> vibrational Raman line

N/A
N/A
Protected

Academic year: 2022

Partager "Isotropic and anisotropic collision-induced light scattering by gaseous sulfur hexafluoride at the frequency region of the ν<sub>1</sub> vibrational Raman line"

Copied!
9
0
0

Texte intégral

(1)

Isotropic and anisotropic collision-induced light scattering by gaseous sulfur hexafluoride at the frequency region of the ν 1 vibrational Raman line

Y. Le Duff, J.-L. Godet, T. Bancewicz, and K. Nowicka

Citation: The Journal of Chemical Physics 118, 11009 (2003); doi: 10.1063/1.1575733 View online: http://dx.doi.org/10.1063/1.1575733

View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/118/24?ver=pdfcov Published by the AIP Publishing

Articles you may be interested in

From light-scattering measurements to polarizability derivatives in vibrational Raman spectroscopy: The 2ν5 overtone of SF6

J. Chem. Phys. 138, 174308 (2013); 10.1063/1.4803160

Isotropic collision induced light scattering spectra from gaseous SF 6 J. Chem. Phys. 116, 5337 (2002); 10.1063/1.1463421

Isotropic collision-induced scattering by CF 4 in a Raman vibrational band J. Chem. Phys. 110, 11303 (1999); 10.1063/1.478004

Collision-induced depolarized scattering by CF 4 in a Raman vibrational band J. Chem. Phys. 108, 8084 (1998); 10.1063/1.476247

Contributions of multipolar polarizabilities to the anisotropic and isotropic light scattering induced by molecular interactions in gaseous CF 4

AIP Conf. Proc. 386, 519 (1997); 10.1063/1.51803

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(2)

Isotropic and anisotropic collision-induced light scattering by gaseous sulfur hexafluoride at the frequency region of the

1

vibrational Raman line

Y. Le Duff and J.-L. Godet

Laboratoire des Proprie´te´s Optiques des Mate´riaux et Applications, Universite´ d’Angers, 2 boulevard Lavoisier, 49045 Angers, France

T. Bancewicz and K. Nowicka

Nonlinear Optics Division, Faculty of Physics, Adam Mickiewicz University, Umultowska 85, 61-614 Poznan´, Poland

共Received 3 February 2003; accepted 27 March 2003兲

Experimental binary isotropic and anisotropic Stokes spectra of the collision-induced light scattered by gaseous sulfur hexafluoride are measured at the frequency region of the␯1 vibrational Raman line. They are compared to theoretical intensities due to dipole–multipole interactions. Taking into account the results of a previous study on the interaction-induced intensities in the Rayleigh wings of gaseous sulfur hexafluoride, an experimental value of the derivative of the dipole–octopole polarizability associated with the␯1 vibrational mode is provided and compared to the result of a recent ab initio calculation. © 2003 American Institute of Physics. 关DOI: 10.1063/1.1575733兴

I. INTRODUCTION

Collision-induced light scattering 共CILS兲 arises in the scattering from dense media through molecular interactions.

Most of the studies on this effect have concerned the Ray- leigh wings of gases or liquids.1–3 However, CILS contrib- utes also to the scattering intensities in the frequency region of vibrational Raman bands.4 – 8In this last case, the scatter- ing of collisional origin is generally less intense than the whole scattering generated by monomers all together, and within vibrational Raman bands CILS has been only mea- sured for few molecules.4,6 – 8 For the sulfur hexafluoride molecule, which belongs to the symmetry group Oh, the derivative of the polarizability tensor with respect to the nor- mal coordinate of the symmetric vibration ␯1 is isotropic in the absence of any molecular interaction. On the contrary, when collisional effects are considered, the collisional- induced component of this Raman derivative polarizability tensor has an anisotropic part4,5which induces a purely col- lisional anisotropic scattering spectrum. Therefore, the study of this ␯1 Raman band is a good opportunity to measure CILS at the frequency region of a vibrational Raman line. In previous work,4 the frequency-integrated CILS anisotropic intensity in the vicinity of the ␯1 line has been studied for gaseous SF6 at room temperature.

In this paper, we present the anisotropic and isotropic CILS spectra measured for a pair of two SF6molecules up to 100 cm1 from the center of the␯1 Raman line. Scattering intensities are given on an absolute scale at room tempera- ture. Experimental data are compared with theoretical semi- classical computations. In our analysis, we consider the dipole-induced dipole共DID兲interaction and high-order mul- tipolar mechanisms.9,10 Taking into account a value previ- ously obtained11for the component E of the dipole–octopole polarizability tensor, we have deduced the value Eof the derivative of the dipole–octopole polarizability correspond- ing to the ␯1 mode of SF6. This result is compared with a

quantum-mechanical value of E⬘ recently calculated for the

1 mode of SF6.12

II. EXPERIMENT

Data were collected using a typical scattering setup de- scribed previously in detail.13 We used the green line (␭L

⫽514.5 nm) of an argon ion laser to illuminate the gaseous sample contained in a four-window high-pressure cell. The light scattered at 90° was analyzed using a double mono- chromator whose output signals were measured using either a low-noise photomultiplier or a charge-coupled device 共CCD兲 cooled at 140 K. Gas SF6 was provided by the Praxair Company with a purity of 99.995%. Gas densities were determined from pressure measurements performed to an accuracy of 1% and pressure volume temperature共PVT兲 data compiled in Ref. 14. Our scattering data were collected for a temperature of the sample stabilized at 294.5⫾1 K. The scattering intensities reported in this work correspond to the Stokes side of the ␯1 Raman band centered at about 775 cm1. The value of this frequency may change slightly according to the gas pressure. Two types of scattering inten- sities, I and I, were measured. They were obtained with the polarization of the laser beam parallel or perpendicular, respectively, to the scattering plane.15 Each intensity value was corrected to take into account the spectral sensitivity of our apparatus as well as the aperture of the scattering beam collected at the entrance of the monochromator. A typical spectrum, displayed in Fig. 1, shows the Stokes side of the

1 band of gaseous SF6 at 19 bars. A weak Raman band at the frequency region of the ␯3 mode is visible in the high- frequency part of the wing with a maximum intensity at about 950 cm1. In order to obtain binary intensities, which is the purpose of this work, we measured scattering intensi- ties at several gaseous densities ␳, up to 27 amagats. Then, for each frequency shift␯considered, we fitted total intensi- ties It(␯) to a power series of the form

11009

0021-9606/2003/118(24)/11009/8/$20.00 © 2003 American Institute of Physics

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(3)

It共␯兲⫽I0共␯兲⫹I1共␯兲␳⫹I2共␯兲␳2I3共␯兲␳3, 共1兲 where I2(␯) and I3(␯) are the two- and three-body correla- tion terms, respectively. I1(␯) stands for the monomers con- tribution, and I0(␯) is the intensity independent of the gas density. In this way, we obtained the parallel and perpendicu- lar pair intensities I,2(␯) and I,2(␯), respectively, accord- ing to the polarization of the laser. To illustrate this proce- dure, typical curves Irf (␳) are displayed in Fig. 2 for several frequency shifts ␯⫽10, 20, and 30 cm1. The re- duced intensity Ir is defined by

Ir⫽共ItI0I,

1

, 共2兲

where I,

1is the integrated perpendicular intensity of the␯1

Raman band. Pair intensities I2(␯) are deduced from the slopes of the curves Irf (␯) at the zero-density limit. Fi- nally, the binary anisotropic spectrum Iani(␯) and the binary isotropic spectrum Iiso(␯) were obtained from

Iani共␯兲⫽I,2共␯兲, 共3兲 Iiso共␯兲⫽I,2共␯兲⫺76I,2共␯兲. 共4兲 All the intensities were calibrated on an absolute scale, on the basis of a comparison with the intensity of the S0(1) rotational line of hydrogen used as an external reference.16 III. THEORY

A. General considerations

We consider a macroscopically isotropic system com- posed of N-like globular molecules in an active scattering

volume V illuminated by laser radiation of angular frequency

L⫽2␲␯L, polarized linearly in the direction e. We analyze the secondary purely collision-induced ␯1 vibration Raman electromagnetic radiation emitted by the system in response to that perturbation. At a point R distant from the center of the sample, the radiation scattered at␻s⫽2␲(L⫺␯1⫺␯) is measured on traversal of an analyzer with polarization n.

Then the quantum-mechanical expression for the pair differ- ential scattering cross section has the form17

共5兲

where ⌬A(1) states for the interaction-induced Raman ␯1

vibration pair polarizability and 具 典denotes a canonical av- erage. Introducing the tensor共dyad兲

Wn e, 共6兲

the pair correlation function F(1)(t) of Eq.共5兲can be rewrit- ten

F1t兲⫽具关W:A(1)共0兲兴关W:A(1)t兲兴典, 共7兲

FIG. 1. Experimental Stokes spectrum of the1 Raman band of gaseous SF6at 294.5 K. Raman frequency shifts are given in cm1. Solid circles共䊉兲 indicate experimental anisotropic intensities I() in arbitrary unit ob- tained with the polarization of the laser beam parallel to the scattering plane.

FIG. 2. Experimental reduced CILS depolarized intensities Ir in cmsee text兲 vs density共in amagat兲for gaseous SF6 at 294.5 K and several fre- quency shifts (10, 20, and 30 cm1) measured from the center of the1

Raman line.

11010 J. Chem. Phys., Vol. 118, No. 24, 22 June 2003 Le Duffet al.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(4)

where the symbol: denotes the twofold scalar tensor product.

Using the reduction scheme of the symmetric second-rank tensor Ti j in its irreducible Cartesian form

Ti jTi j(0)Ti j(2), 共8兲 where

Ti j(0)⫽共13Tkk兲␦i j, Ti j(2)12Ti jTji兲⫺共13Tkk兲␦i j, 共9兲 we transform the W andA(1) tensors into their irreducible forms and we change their coupling schemes. Next we group the geometrical and molecular parts and we assume that for an isotropic fluid the mean value occurring in Eq.共7兲has to be a rotational invariant. Then for the pair correlation func- tion F(1)(t) we obtain

Fj1t兲⫽

0,2j 2 j1⫹1共W( j):W( j)兲具A( j) (1)

共0兲:⌬A( j)(1)t兲典. 共10兲 We note that our pair correlation function F(1)(t) for the isotropic ( j0) as well as the anisotropic part ( j⫽2) of the scatterred light is a product of the geometrical factor

j j⫽ 1

2 j⫹1共W( j):W( j)兲 共11兲 and the molecular factor

F(j j1)t兲⫽具⌬A( j)(1)共0兲:⌬A( j)(1)t兲典. 共12兲 Taking into account the form of the dyad W, using Eqs.共6兲, 共8兲, and共9兲, for the pair double-differential cross sections of the isotropic and anisotropic collision-induced Raman light scattering we finally obtain

⳵␻⳵2

iso

⫽共e"n2

3

s 4

c4

⫺⬁

dt eitF00(1)t兲, 共13兲

⳵␻⳵2

ani

⫽3⫹共e"n2

30

s 4

c4

⫺⬁

dt eitF22(1)t兲. 共14兲 For the A – B molecular pair the incremental pair Raman polarizability tensor⌬A␣␤(1)for the normal vibration␯1reads

A␣␤(1)⫽⳵⌬A␣␤

QA

1 QA1⫹⳵⌬A␣␤

QB

1 QB1, 共15兲 where ⌬A␣␤ is the pair Rayleigh excess polarizability and Qp1 is the normal coordinate for the mode␯1 and molecule p. Moreover, for the1 vibration of SF6, the Raman polar- izability tensor of an isolated molecule is isotropic and, con- sequently, monomer polarizabilities do not contribute to the depolarized Raman scattering considered in this work. The analytical expressions for the pair Rayleigh-multipolar first- order isotropic and anisotropic parts of the pair polarizability for octahedral molecules have been given elsewhere.11,18

B. Second-order isotropic Raman scattering

Now let us consider the second-order isotropic Raman

1 mode light scattering correlation function in detail. For the A – B pair the Rayleigh second-order DID incremental 共excess兲pair polarizability tensor reads19,20

(2)A␣␤⫽␣A

2BT␣␥ABT␥␤BA⫹␣B

2AT␣␥ABT␥␤BA. 共16兲 In the case of small vibrations of the nuclei, the polarizability can be expanded into a Taylor series

␣共Qp1兲⫽␣p共0兲⫹␣p(1)Qp1⫹¯, 共17兲 where

p(1)

⳵␣Qp1

0

. 共18兲

Then according to Eq.共15兲for the collision-induced Raman polarizability we obtain

A␣␤(1)⫽关共2␣ABA(1)⫹␣B

2A(1)QA

⫹共2␣ABB(1)⫹␣A

2B(1)QBT␣␥ABT␥␤BA. 共19兲 Taking into account that T␣␥ABT␥␣BA6RAB6 and assuming that we deal with the gas composed of identical molecules, for the isotropic Raman second-order pair polarizability we ob- tain

(2)Aiso(1)(2)A␣␣共␯1

3 ⫽6␣2(1)

RAB6QA1QB1兲. 共20兲

TABLE I. Excess Rayleigh nonlinear origin anisotropic pair polarizabilityA2 M(NL)tensor for octahedral molecules.

Mechanism jA jB lA lB N A2 M(NL)

B0T6 0 4 2 4 6

10513 60 T6

(AB)(A)B0(B)Q4(4)2 MT6

(AB)(A)Q4(B)B0(0)2 M

B4T6 4 4 2 4 6

x

bxT6

(AB)(A)B4(B)Q4(x)2 M⫹共⫺1兲xT6

(AB)(A)Q4(B)B4(x)2 M

b4⫽⫺ 1

601155 b5⫽⫺

1

60105 b6⫽⫺

23113

60 b7⫽⫺ 1

12105 b8⫽⫺

1721 60

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(5)

The pair polarizability⌬Aiso(1)(t) depends on the transla- tional and vibrational variables of molecules. It is generally assumed that the vibrational variables Q are independent of the others and that the vibrational motions of two molecules are uncorrelated. The vibrational lifetime is very long with respect to the induction phenomena. Then we can write

QA共0兲QBt兲典⫽␦ABQ2cos1t, 共21兲 where␻1 is the vibrational frequency. Finally, for the isotro- pic Raman spectrum due to the second-order DID light scat- tering mechanism we obtain from Eq. 共12兲

F00(1)t兲⫽具(2)Aiso(1)共0兲(2)Aiso(1)t兲典

⫽72␣4共␣(1)2RAB6共0兲RAB6t兲典具共Q12. 共22兲 C. Contributions from nonlinear polarizabilities

In order to obtain contributions to the Raman spectrum from nonlinear origin polarizabilities, the Rayleigh form of the respective pair nonlinear polarizabilityA␣␤ is neces- sary关see Eq.共15兲兴. We arrived at the following formula for the excess pair Rayleigh polarizability due to nonlinear po- larization of the molecules of a pair:7

AK M(NL)关11兴

⫽共1⫹PABj

AlA,lB 共⫺1兲(KNlBx1)

2lA2!N2lB!

1/2

jNA lKB lxA

XXjAKNx

⫻兵TN (AB)Bj

A关共11兲KlAQl

BxK M. 23

For the␯1vibration of sulphur hexafluoride, the lowest- order contribution to this mechanism results from the dipole2-quadrupole hyperpolarizability tensor B␣,␤,␥ of one SF6 molecule combined with the permanent hexadecapole moment⌽ of the other SF6 molecule. For SF6 in the frame of the main reference axes, the B,, tensor has two inde- pendent Cartesian components Bz,z,zz and Bx,z,xz. For this tensor the following relations between its irreducible spheri- cal and Cartesian components hold:

B00关共11兲22兴⫽

65B0, B40关共11兲22兴⫽

157 2 ⌬B,

共24兲 B44关共11兲22兴⫽ ⌬B

2

6,

where B0Bz,z,zz2Bx,z,xz and ⌬B3Bz,z,zz4Bx,z,xz, whereas for the hexadecapole moment we have21

Q40⫽⌽, Q44

145 ⌽. 共25兲 Table I gives the form of the SF6 excess Rayleigh nonlinear origin anisotropic pair polarizabilityA2 M(NL) tensor derived using Eq.共23兲.

In our considerations of the light scattering correlation function, Eq. 共12兲, we assume that rotational and transla- tional motions of SF6 molecules are not coupled. Moreover, we assume that the molecules in the scattering volume are correlated radially but uncorelated orientationaly. We use the following formula to calculate the succesive components of the tensorial part of the correlation function F22(1)(t), Eq.

共12兲:

具兵TN共0兲Aj

A共0兲Bj

B共0兲兴xL

䉺兵TNtCj

AtDj

Bt兲兴x⬘其L

⫽共⫺1兲LNjAjBxxXL2 XN j

AjB

2 具关TN共0兲䉺TNt兲兴典

⫻具关Aj

A共0兲䉺Cj

At兲兴典具关Bj

B共0兲䉺Dj

Bt兲兴典

⫽共⫺1兲LNjAjBxxXL2 XN j

AjB 2

2N兲! 2N

⫻共

jA

jA兲共

jB

jBSNtRj

AtRj

Bt兲, 共26兲 where the tilde denotes the polarizability or the permanent multipole moment in its molecular frame reference system.

For the nonlinear polarizabilities discussed here, the respec- tive scalar tensor products appearing in Eq. 共26兲 are of the form

00兲⫽6

5B02, 共44兲⫽1 5⌬B2, 共⌽˜䉺⌽˜兲⫽12

7 ⌽2,

共27兲 共

4䉺⌽˜兲⫽2

353B⌽, 共䉺⌽˜兲⫽6

7

7 E⌽, 共4兲⫽

35EB.

The autocorrelation functions of the isotropic and anisotropic parts of the Raman pair polarizability are provided in Table II for successive multipolar and nonlinear induction opera- tors. In Table II, the F function for the Raman Stokes side of the˜1 normal vibration refers to.22

Fb

1 2

1⫺e[hc˜1/kBT], 共28兲 where b

1⫽关h/(82c˜1)兴1/2 is the zero-point vibrational amplitude of the␯1 mode (h being Planck’s constant, c the light velocity, and˜1 the normal-mode frequency in cm1).

Moreover, in Table II, E and ⌽ stand for the independent component of the tensor E and the hexadecapole moment⌽, whereas their normal vibration derivative is defined as Z

⫽⳵Z/⳵Q1

p. It is noteworthy that the normal vibration deriva- tive is related to a corresponding bond length R derivative.

For SF6with six equivalent bonds this relation has the form22

11012 J. Chem. Phys., Vol. 118, No. 24, 22 June 2003 Le Duffet al.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(6)

⳵␣Q1

molecule

⫽ 1

6mF

⳵␣R

molecule

, 共29兲

which may be extended to any multipolarizability, hyperpo- larizability, or multipole moment.

IV. RESULTS AND DISCUSSION

Experimental anisotropic and isotropic CILS spectra of SF6in the vicinity of the␯1Raman line共pointed hereafter as

‘‘Raman spectra’’兲are reported in Figs. 3 and 4, respectively.

Experimental intensities are given on an absolute scale for

the Stokes side of the␯1 lineup to␯⫽100 cm1, where␯is the frequency shift measured from the center of the ␯1 line.

They are compared with the theoretical double-differential cross sections, taking into account translational and ro- totranslational contributions computed with the SF6intermo- lecular potential Z of Zarkova.23According to Eq.共5兲these theoretical intensities are proportional to the Fourier trans- form of the autocorrelation functions provided in Table II. In Fig. 4, the isotropic translational contribution共ITC兲is calcu- lated from the second-order DID mechanism by using clas- sical trajectories of two molecules in interaction.24,25In Fig.

3, the anisotropic translational contribution共ATC兲to the Ra- man spectrum is deduced from the experimental anisotropic Rayleigh spectrum.11Indeed, considering this Rayleigh spec-

TABLE II. The coefficientsLL N, j1, j2

involved in the Raman isotropic and depolarized light scattering correlation functions FLL(t)⫽兺N, j1, j2LLN, j1, j2SN(t)Rj1(t)Rj2(t)F for successive multipolar and nonlinear induction op- erators.

Raman

CIS N j1 j2

Isotropic

00 N, j1, j2

Depolarized

22 N, j1, j2

First-order DID 2 0 0 0 48(␣␣(1))2

Second-order DID 4 0 0 216(2(1))2 108(2(1))2 DO 4 0 4 2249 ((1)E)2(E(1))2 2209 ((1)E)2(E(1))2

OO 6 4 4 110

3 (EE(1))2 2266021 (EE(1))2 B0 6 0 4 0 158435 关(B0(1))2(B0(␯1))2

B4 6 4 4 0 8815(B(␯1)⌽⫹⌬B(␯1))2

OO–B4 6 4 4 0 10567 EE(1)(B(1)⫹⌬B(1))

FIG. 3. Two-body anisotropic scattering spectra on an absolute intensity scale (cm6) for the Stokes wing of the1Raman band of SF6at 294.5 K.

Frequency shifts are measured in cm1from the center of the1line. Our experimental datasolid circles共䊉兲兴are given with error bars. The partial and total theoretical spectra are computed for the Z potentialRef. 23,

4.549 Å3 Ref. 32,7.38 Å2 Ref. 33, E3.1 Å5 Ref. 11, and E46 Å4: anisotropic translational contribution ATC: - - - -, dipole- induced octopoleDIO: – – –, octopole-induced octopoleOIO: — - —, nonlinear-origin contributionNLC: , and total . We also provide the dipole-induced dipoleDIDcontribution— - - —due to the free dimers in the 10– 100 cm1frequency range.

FIG. 4. Two-body isotropic scattering spectra on an absolute intensity scale (cm6) for the Stokes wing of the1 Raman band of SF6at 294.5 K. Fre- quency shifts are measured in cm1 from the center of the1 line. Our experimental data关solid circles共䊉兲兴are given with error bars. The partial and total theoretical spectra are computed for the Z potentialRef. 23,

4.549 Å3 Ref. 32,7.38 Å2 Ref. 33, E3.1 Å5 Ref. 11, and E46 Å4: dipole-induced dipoleDID: - - - -, dipole-induced octopole DIO: – – –, octopole-induced octopoleOIO: — - —, and total .

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(7)

trum, it is possible to subtract the theoretical rototranslational intensities Irot/aniRay leigh(␯) from the experimental intensities Iexp/aniRayleigh(␯) to get the translational intensities IATCRay leigh(␯):

IATCRay leigh共␯兲⫽Iexp/aniRay leigh共␯兲⫺Irot/aniRay leigh共␯兲. 共30兲 On the other hand, according to Table II and the autocorre- lation functions given in Ref. 11, we can write

IATCRaman共␯兲⫽ 2

2

⳵␣Q1

2

LL1

4IATCRay leigh, 31

where␯ is the frequency shift measured from the center of either the Rayleigh line for the Rayleigh spectrum or the Raman line for the Raman spectrum. The computation of the rototranslational dipole-induced octopole 共DIO兲, octopole- induced octopole 共OIO兲, and nonlinear共NLC兲 contributions has been previously described.13To provide them on an ab- solute intensity scale, several values for polarizabilities and multipoles must be known. As far as the nonlinear contribu- tion given in Fig. 3 is concerned, we used the values of ⌽,

B, B0, and of their R derivatives computed by Maroulis.12 The value chosen for the dipole–octopole polarizability E (E⫽3.1 Å5) is the one proposed in our previous paper on the Rayleigh spectra of SF6 for the aforementioned potential.11The Evalue (E46 Å4) is obtained by fitting the theoretical anisotropic Raman spectrum to the experi- mental data in the 30– 70 cm1 frequency range in which the multipolar contributions play the leading role. Because of the competition between E and E⬘ in the coefficients in- volved in the different correlation functions given in Table II, several values of the pair (E,E⬘) may correspond to similar theoretical intensities. The ensemble of such solutions forms an area, as shown in Fig. 5 for the potential Z in the plane sector defined by E苸关2.2, 4.0 Å5and E⬘苸关43, 49 Å4兴. This area has been obtained through set inversion26 by con- sidering both the Rayleigh isotropic spectrum 共used in Ref.

11 for the measurement of E) and Raman anisotropic spec- trum. For any point (E,E⬘) in this area, theoretical intensi- ties lie inside the error bars of all experimental data in the frequency ranges for which multipolar effects may be con- sidered as predominant (30– 85 cm1 for the isotropic Ray- leigh spectrum and 30– 70 cm1 for the anisotropic Raman spectrum兲. The pair of values (E⫽3.1 Å5,E⬘⫽46 Å4) used in Figs. 3 and 4 is set at the center of the area of solutions provided in Fig. 5 for the potential Z. As far as the isotropic Raman spectrum reported in Fig. 4 is concerned, it can be noticed that in this case theoretical spectrum is far below experimental data. Moreover, as follows from Fig. 5 the value of E⬘ cannot exceed 48.2 Å4 for the potential Z. The choice of this maximum value would not be sufficient to obtain theoretical isotropic intensities as high as experimen- tal ones. Similar discrepancies have been observed in the case of CF4.8 They are connected to the surprisingly low values reported in Fig. 6 for the depolarization ratio ␩()

I,2(␯)/I,2(␯) of SF6. Even at low frequencies, the depo- larization ratio is lower than the value expected for the in- tensities generated by the dipole–octopole mechanism alone (␩⫽22/63⬇0.35).27 This shows that the experimental iso- tropic intensities are greater than those produced by multipo- lar mechanisms. Therefore, we may conclude that the inten-

sities Iiso(␯) result not only from the dipole–multipole mechanisms but also from another contribution.

Several effects may be mentioned in order to explain this behavior. The first one is connected to the inaccuracy of the potentials available for globular molecules such as CF4 or SF6. In Table III, we provide the pairs of values (E,E) obtained by using different SF6 potentials. The use of a dif- ferent potential slightly affects the values of E and Eas may be noticed in Table III, but cannot allow any increase in the theoretical isotropic intensities without a simultaneous 共and here forbidden兲 increase in the theoretical anisotropic intensities in the midfrequency range used to obtain the val- ues of E and E⬘. On the other hand, there are SF6potentials 共e.g., the MMSV potential given in Ref. 28兲which induce an enhancement of the spectral line shape of the multipolar con- tributions beyond 100 cm1. Nevertheless, this enhancement does not concern the frequency range of our Raman experi- mental data and cannot alone explain the lack of the theoret- ical intensities beyond 70 cm1 for both anisotropic and iso- tropic Raman spectra.

The second effect refers to short-range interactions such as the overlap and exchange mechanisms, molecular frame distortion, and the effect of the nonpointlike size of the mol- ecule. They play a role at high frequencies, but are not well known for SF6. On the one hand, in the anisotropic case, the extrapolation of the translational contribution from the Ray- leigh anisotropic spectrum is a way to take them into ac- count. Comparing the deduced Raman translational curve 共ATC兲of Eq.共31兲with the pure DID contribution共computed

FIG. 5. Area of solutions for the dipole–octopole polarizability E and its R derivative E. The possible values are inside the trapezium. These results are obtained from a set inversion analysisRef. 26using the comparison of experimental data with isotropic Rayleigh spectrum共from 30 to 85 cm1) and anisotropic Raman spectrumfrom 30 to 70 cm1) calculated with the potential of ZarkovaRef. 23. The cross () shows the values used in Figs.

3 and 4 (E3.1 Å5and E46 Å4).

11014 J. Chem. Phys., Vol. 118, No. 24, 22 June 2003 Le Duffet al.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

(8)

from classical trajectories兲, we found that they are close be- low ␯⫽30 cm1for the potential Z, differences are less than 15% whatever the frequency is兲. This shows that DID mechanism is predominant in this frequency region. Beyond 30 cm1, the ATC intensities become lower than the com- puted DID intensities, due to short-range effects. Taking into account the ATC instead of DID intensities leads to a slight enhancement of the measured Evalueless than 10%兲. On the other hand, in the Raman isotropic case, the lack of the- oretical intensity is not limited to the highest frequencies and concerns the whole frequency range of the spectrum. More- over, consideration of all translational contributions yields scattering intensities generally lower than the DID ones. This behavior cannot explain the difference observed between theoretical and experimental data for isotropic intensities.

The splitting of the Q(J) lines of the1 band due to the vibrational–rotational coupling could be the third effect to consider. The consequence of this effect decreases when the density or frequency increases and may be regarded as neg- ligible for our binary isotropic spectrum, in the frequency range scanned. Similarly, the influence of the nonlinear con- tributions, which, in the linear aproximation of TN, exists in the anisotropic spectrum and not in the isotropic spectrum, can be considered. In the case of the forbidden SF6 vibra- tional line ␯3symmetry species: F1u) the dipole moment belongs to the F1u representation too, so the off-diagonal matrix elements of the dipole moment are nonzero for this symmetry and then contribute to the Raman nonlinear origin polarizability. In this case the nonlinear origin light scatter- ing mechanism is of the the same order as multipolar contri- butions in computations of the intensity of the ␯3 Raman line, as suggested by Samson and Ben-Reuven.29 However, in our case the nonlinear origin contributions to the Raman spectrum of the␯1band of SF6are relatively small共see Fig.

3 for values computed by Maroulis12兲. We deal with the di- agonal 共ground-state兲 matrix elements of the hexadecapole moment and the nonlinear hyperpolarizability B as well as their off-diagonal matrix elements of the totally symmetric representation A1g. For this symmetry the hexadecapole is the first nonzero electric moment. Then this high-order mo- ment coupled with dipole2–quadrupole hyperpolarizability B gives a small contribution. Its influence on the measurement of E⬘ may be neglected and cannot explain the discrepancies observed.

Finally, the discrepancies between the experimental and theoretical isotropic spectra are probably more connected to the far wings of the ␯1 Q branch. Although the integrated intensity of this Q branch is proportional to the density, the dependence on the density of the spectral intensities in the wings may be different. Indeed, the intensity in the far wings of allowed spectral lines comes mainly from short-time in- teractions in binary collisions and is therefore quadratic in the density. Thus, the wing intensities of the ␯1 line gener- ated by the monomers may contribute to the experimental pair isotropic spectrum Iiso(␯) measured in this work using the virial expansion described above in Eq.共1兲.

Consequently, the frame of our analysis is very different from the one described in our previous work about the Ray- leigh wings of SF6.11In the Rayleigh case, the strong trans- lational contribution IATCRay leigh to the anisotropic spectrum predominates in the 0 – 100 cm1 frequency range scanned.11 It is thus more advisable to deduce the measured value of E from the Rayleigh isotropic spectrum for which the DID contribution is of second order and lower than mul- tipolar contributions.11On the contrary, because of the prob- able influence of the far wings of the␯1 line on the isotropic Raman spectrum, the latter is not adequate to measure the influence of multipolar mechanisms. Fortunately, in the Ra- man case, the translational contribution to the anisotropic spectrum is relatively less important than the one found in the Rayleigh case. In Fig. 3, the dipole–octopole contribu- tion becomes predominant beyond 30 cm1 where it never reaches 10% of the total intensity in the 0 – 220 cm1 fre- quency range of the anisotropic Rayleigh spectrum. Conse-

FIG. 6. Experimental pair depolarization ratio()I,2()/I⬜,2() vs fre- quency shifts in cm1 for the Stokes side of the 1 Raman band of gaseous SF6 at 294.5 K. The theoretical depolarization ratio DIO()

22/7 associated to the dipole-induced octopole mechanism is represented by a dashed horizontal line– – –.

TABLE III. Dipole–octopole polarizability E of SF6and its R derivative E for the1 mode of SF6. Experimental values are deduced from the com- parison of experimental data with theoretical spectra computed by using several potentials.

Theory E (Å5) E4)

Ab initioRef. 12 4.2 22.6

ExperimentCILS E (Å5) E4)

Potential Ref. 11 This work

28-7 LJ共Ref. 31兲 2.7⫾0.6 42⫾3

MMSVRef. 28 2.60.6 402

HFDRef. 25 3.80.8 552

ZRef. 23 3.10.7 462

Best/recommended 3.01.01.5 4710

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 193.52.40.1 On: Tue, 03 May 2016

Références

Documents relatifs

In order to illustrate the non-linear impact of the injec- tion magnitude on the spatial distribution of the stratospheric aerosol, we plotted the sulfate aerosol mass mixing

Summary of the main theories which predict the spectrum of depolarized light scattered from simple liquids Am frequency shift mi + m in ( cm-' ] ; J molecular moment of inertia ; z

As a contribution to the special issue in Biogeosciences on “Human impacts on carbon fluxes in Asian river sys- tems”, we report a data set obtained in the three branches (M � y

Furthermore, It should be noted that a similar trend was observed by comparing the bleaching of 1 and the tetracarboxyphenyl zinc porphyrin (Zn-TCPP), indicating that the

The study of collision induced light scattering 共 CILS 兲 intensities has been demonstrated to be a useful method to investigate collision induced polarizabilities in gaseous samples.

The blueprint for the presented study has four major steps following the footprints of our previous works: quantum chemistry (QC) ab initio calculations of the hyperpolariz-

rate, the lack of trees in the southeast African mountains during glacial extremes is unlikely the result of drought, because our record indicates that climate conditions in SE

In order to confirm that the 1 33 S seasonality we are ob- serving in Montreal aerosols does not reflect the signatures of primary mineral dust particles transported to the city,