• Nem Talált Eredményt

Ralph L. Amey Department of Chemistry, Occidental College, Los Angeles, California

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Ralph L. Amey Department of Chemistry, Occidental College, Los Angeles, California"

Copied!
20
0
0

Teljes szövegt

(1)

Ralph L. Amey

Department of Chemistry, Occidental College, Los Angeles, California

1. Introduction 183 2. Theory 184

2.1. Dielectric Constant 184 2.2. Clausius-Mossotti Function 185

2.3. Multipole Effects 189 2.4. Index of Refraction 190 2.5. Lorentz-Lorenz Function 192

3. Experimental 192 3.1. Literature Sources 192

3.2. Reference Liquids 193 3.3. Pure Liquids 194 3.4. Liquid Mixtures 198

4. Conclusion 200 References 200

L Introduction

With the continued success of theoreticians in refining techniques for calculating molecular properties [1], it is certainly apropos that the experimental data available for comparison with theory be examined critically. In the area of nondipolar liquids the dielectric data of present theoretical usefulness are particularly sparse. Nevertheless, several recent investigations are of interest in that they suggest—in a composite fashion—the direction which further profitable studies should take.

Recognizing the difficulties attendant upon a priori polarizability calculations, we note that the first calculation of the frequency-dependent polarizability of molecular hydrogen was not reported until 1964 [la].

This should not be too surprising, for the problem is not a straight- forward one even for isolated atoms. Furthermore, when the atom or molecule is immersed in an intermolecular force field, the prospects of maintaining a realistic model—which at the same time describes its dielectric properties in a mathematically tractable form—decrease

183

(2)

considerably. The polarization terms which exist for the isolated molecule will usually be modified in some manner by intermolecular forces. In addition, further polarization terms may arise due to the system of molecules in strong interaction.

Atoms and molecules are described satisfactorily on a mathematical basis only within the context of quantum mechanics. Thus when induced polarization terms are related to the frequencies and matrix elements for associated energy transitions in a molecule, both the fre- quencies and matrix elements will be altered—and hence a change in that polarization term—when the intermolecular force changes. In terms of perturbation theory this entails modifying the wave function for the isolated ground state molecule by the inclusion of modified excited state wave functions. To include such terms increases the mathematical problem considerably.

It becomes evident that the type and magnitude of intermolecular forces significant to a given system will probably be a function of the exact geometry and charge distribution of the interacting molecules.

This further complicates the problem, for it means that the treatment of dielectric behavior in transition regions—particularly near the critical point (see Section 4)—cannot be as general as one would like. Some efforts have been made to circumvent certain of these problems by combined variation-perturbation techniques [2, 3]. Such methods yield good values for the wave functions. However, the polarizabilities obtained from such wave functions will be accurate only to the extent to which the model is a realistic description of the molecular system. Calculations based on exact variational procedures are limited to extremely simple molecules [4].

The principal concern of this chapter is the use of electromagnetic data—especially the dielectric constant and refractive index—in the interpretation of molecular interactions and molecular structure. The development of the relations between experimentally measurable macroscopic quantities and the microscopic molecular quantities (e.g., polarizability and multipole moments) are briefly presented in Section 2.

The available experimental data on several nonpolar liquids and liquid mixtures are presented and discussed in Section 3. Quadrupole moment data for a number of molecules are summarized in Section 2.

2* Theory

2.1. DIELECTRIC CONSTANT

The dielectric constant e was first introduced as the ratio:

e = C/C0

(3)

where C is the capacitance of a parallel-plate condenser filled with a gas, liquid, or solid, and C0 is the vacuum or geometric capacitance of the same condenser. It was discovered experimentally that if a static or low frequency field of moderate intensity is applied, e is only a function of:

(1) the chemical composition and (2) such parameters (density, etc.) necessary to define the states of the material.

For theoretical considerations it is more convenient to express the dielectric constant in terms of a set of vectors which satisfy Maxwell's equations [5]. The electric field strength is defined as a vector E which satisfies the condition curl E = 0, and the dielectric displacement vector D as one satisfying the condition div D = 4πρ} where p is a volume charge density. These in turn are related by a third vector quantity P, the electric polarization or induced dipole moment per unit volume:

D = E + 4TTP

In gases, most liquids, and isotropic solids (i.e., those with cubic sym- metry), the vectors E, D, and P are all parallel. Thus we may write P = χΕ and D = (1 + 47τχ)Ε, where χ is defined as the electric suscepti- bility. If now the dielectric constant or permittivity is defined as

it follows that D = eE. In the case of crystals with anisotropic symmetry this definition must be modified to allow χ and e to operate as tensors.

Also it is to be remembered that € is "constant" only if there is no phase difference between D and E. For this reason it has often been suggested [6] that the term "permittivity" is a better name for e than dielectric constant.

2.2. CLAUSIUS-MOSSOTTI FUNCTION

2.2.1. The Lorentz Field

An expression relating the dielectric constant to molecular properties of atoms with polarizability a is given by

^ - = ^ (2.1)

e + 2 n 3 v '

where n\V is the molar density, N0 is Avogadro's number, and where the experimentally determinable left-hand quantity is the Clausius- Mossotti (C-M) function [7, 8].

The principal assumptions made in the derivation of Eq. (2.1) are:

(4)

(1) m = OÎEL is the only induced moment present in the molecules and is the same for all molecules; and that the Lorentz field [9]

EL = [(c + 2)/3]E rather than E is to be used in calculating the polar- ization P.

(2) That the molecules can be considered as being located on a lattice of cubic symmetry.

(3) That a is a molecular constant.

Within the framework of Lorentz's assumptions the C-M function predicts a molecular constant independent of density and temperature.

This prediction is in fact a rather good approximation for many nonpolar gases and liquids. For many crystalline solids of cubic symmetry it is an excellent approximation. However, in most cases, at least one of the assumptions which have gone into the C-M function formulation proves to be too naive, and deviations—some admittedly small—will be observed.

2.2.2. Moderate-Density Fluids

In the case of gases, the regularly spaced atoms or molecules of a crystal lattice no longer exist; each particle's neighbor-distance is continually and rapidly changing, and the molecules have no fixed points of reference. The problem can thus be stated [10] in terms of the evaluation of the mean moment (ja^ · c0> parallel to the applied field E0

for molecule j :

<m, · e0> = aE0 + ot(Y ejk · e0^ (2.2) where oc is assumed constant and the field ejk is due to the other mole-

cules k. The resulting set of vector equations becomes rather intractable when considered generally due to the introduction of the total energy into the classical averaging process through the Boltzmann factor.

An evaluation was first made by Kirkwood [11] and Yvon [12] for nonpolar molecules by assuming moderate gas densities and small applied field E0, and expanding in terms of a power series in density or the polarizability. The interaction field ejk was calculated in terms of the moment m, = a(E0 + ejk), where

«* = -grad, [ ^ ]

= -grad, ^ [«E. - « grad, £ ^(*E0 - ···)] (2.3)

jK i ^ fc Ofcz

(5)

Upon successive substitution it is seen to be a power series in a. By proper arrangement of terms an expansion in pair and higher order interactions results. In order to evaluate Eq. (2.3) it is necessary to carry out the averaging process over configurations, and hence the name

"translational fluctuation effect" given by Kirkwood to the contribution due to these terms. As a density expansion

^(-f)=^*(^)+ c (^)'+··· <«>

where

In analogy with the virial expansion for a gas A, B, and C are often referred to as the first, second, and third dielectric virial coefficients [13].

Others have sought with varying degrees of success to summarize the classical statistical theory for nonpolar substances [14, 14a, 15, 16].

Each obtained an expression of the form

C M = i f i [ l + s ( i , r ) ] (2.5)

where S expresses the density and temperature dependence in various ways. Attempts were made to express in explicit form the interference between orientation and deformation effects as well as other polarization phenomena which might occur, such as surface polarization. These treatments predict a gradual increase in the C-M function with density with a maximum to be observed at higher densities. This is in qualitative agreement with experiments conducted on nonpolar molecules in the gas phase. In going to liquids and crystalline solids, Lorentz's considera- tions should have more validity as various fluctuations wash out [11].

The possibility that C, the third virial coefficient, might be negative should not be ignored [15], since then one would predict the further decrease in the C-M function experimentally observed at higher fluid densities.

Jansen and Mazur [17] considered the evaluation of the C-M function when including (1) the change in effective polarization with varying distances between molecules (i.e., density effects), and (2) the effect of statistical fluctuations in (a) the polarizability of the molecules and (b) the induced dipole moments. By performing a perturbation calculation of the mean moment in terms of the wave functions of the free molecules the initial contribution to a was found to come from the third-order

(6)

perturbation. A crude calculation suggested a positive contribution comparable to that attributed to fluctuation effects. This was then statistically averaged over a canonical ensemble to yield an expression of the form

C-M = ^ â [ l + * ( - f , r ) ] (2.6) where ä is an average polarizability due to density effects and can be

expanded in powers of oc0 , the polarizability of an isolated atom; R includes the fluctuation effects of both m and ä.

Unfortunately from the viewpoint of liquid behavior, Jansen and Mazur assumed conditions of relatively low density. For simplicity exchange forces were ignored and all excited energy levels were replaced by a group of practically identical energy levels. A further conclusion drawn was that the polarizability for a system of isotropic harmonic oscillators interacting through induced dipole forces is a constant in any order of approximation. However, they showed that the polarizability does not remain constant when a more realistic model is used. Similar calculations by Jansen [18], Jansen and Solem [19], and Yaris and Kirtman [20], also predict a positive contribution to the C-M function. In all of these quantum-mechanical treatments, however, perturbation or variational wave functions for a pair of atoms have been constructed from products of single-molecule wave functions, and they are therefore most appropriate in the attractive region of molecular separations for which overlap or exchange effects can be neglected.

2.2.3. High Density Fluids

No quantitatively reliable theory of interaction effects at short range has yet been developed, although several interesting calculations are available. Ten Seldam and deGroot [21], by employing as a model an isolated compressed atom, assume neither fluctuations in a nor attractive forces between atoms, and calculate a decrease in a at high pressures.

Dick and Overhauser [22] calculate a decreased polarizability in alkali halide crystals as a result of overlap effects. Amey's calculations [23] of the effect of changes in boundary conditions on the polarizability of a particle in near-harmonic potential fields suggest that placing a bound on the permissible region of the particle decreases the particle's polariza- bility. This is equivalent to using a harder potential function at small distances of separation. Jansen and Mazur's employment of large intermolecular distances, corresponding to lower densities, is equivalent to the use of a softer potential than that of the harmonic oscillator. Such a behavior is suggested in Fig. 1, where the harmonic oscillator (HO)

(7)

/ / / / I I /

1 '

/ II

\

\1

\\ \\

1/ /

• w

1 ^ 1 /

\f 1 / Û

> a , _ . . . - it n

/ L J /

- 1

Intermolecular distance

FIG. 1. Effective polarizability (a), harmonic oscillator potential (HO), and Lennard- Jones potential (L-J) as a function of intermolecular distance. Plotted in arbitrary units.

potential, Lennard-Jones (L-J) potential, and the effective polarizability (a) are plotted in arbitrary units against interatomic distance. Calculations of a0 from the liquid density at the boiling point plus dielectric data measured near the boiling point suggest that a turns over at greater interatomic distances than the L-J potential does. It is possible to infer from this that a more realistic model about the critical point might utilize a potential harder than the harmonic oscillator at small intermo- lecular distances and softer at large separations. The electron cloud configuration will change in high density regions, so the accompanying wave functions must be changed accordingly to maintain orthogonality conditions.

2.3. MULTIPOLE EFFECTS

For molecules with less than spherical symmetry there may be a significant contribution to the C-M function from quadrupole and higher order multipole terms [24, 25]. An excellent review of the subject through 1959 is given by Buckingham [26]. Such contributions can

(8)

result from the moments induced in nondipolar molecules by permanent multipole moments of a neighbor. These have a net effect when an external field makes orientations of low energy favorable. Because they arise from pair interactions there will be a contribution for molecules with axial symmetry to the second virial coefficient B of Eq. (2.3) of the form:

ΒΘ = (47r7V0)W(l + §k*)(R-*\ßkT (2.7)

where Θ is the quadrupole moment, ä is the mean polarizability, and

<i?-8>0 is defined by

<#-8>0 = j R-s exp[-W(R)lkT]R* dR

W(R) is a radial intermolecular pair potential and k is the anisotropy factor. It is also seen that B will now be temperature-dependent. Jansen [18] has calculated that polarizability and fluctuation effects give an increase of 0.37 % in the C-M function for nitrogen at a density of 4 moles/liter, and the quadrupole effect yields a further increase of 0.08 % at 316°K. The total of 0.45% is only slightly larger than that observed experimentally by Johnston et al. [27]. With the stronger quadrupole moment present in carbon dioxide, Jansen estimates a contribution to the C-M function of about 1%. Experimental results are consistent with this [28-31]. Quadrupole moments for several molecules are noted in Table I.

Possible octupole-induced dipole interactions have been suggested as capable of explaining the increase in C-M function for gaseous methane at high pressures [26]. A value for the octupole moment of methane calculated from dielectric virial coefficients was given as

ω = 1.7-3.4 X 10~34 esu cm3.

A calculation of ω by Parr et al. using a one-center wave function with Slater orbitals yielded a value of 2.6 X 10- 3 4 esu cm3 [40]. Kielich [41]

has reported ω = 5-8 X 10~34 esu cm3 when calculated indirectly from second virial coefficients.

2.4. INDEX OF REFRACTION

The refractive index n may be defined in terms of the phase velocity v- of an electromagnetic wave of a certain frequency (v) in the medium and the corresponding velocity C in vacuum:

(9)

TABLE I QUADRUPOLE MOMENTS

Ô x 102e esu cm2

±1.7 1.8 4.4 4.59 4.6 5.2 5.3 2.6 3.92 4

<0.3 0.52 0.6 0.63 0.95

<1.0

±0.80 1.0 1.64 1.8 1.80 1.90

<0.3

Method Pressure broadening Pressure broadening Microwave absorption Dielectric virial coefficient P.V. virial coefficient

Microwave absorption (corrected) Dielectric constant

Microwave absorption Dielectric virial coefficient Dielectric virial coefficient Microwave line broadening Dielectric virial coefficient Pressure-induced broadening Diamagnetic susceptibility Viscosity

Microwave absorption Pressure broadening Microwave absorption Pressure-induced absorption Pressure broadening Viscosity

Dielectric virial coefficient Microwave absorption

Ref.

[32]

[31]

[33]

[34]

[30]

[28]

[29]

[28]

[34]

[29]

[31]

[34]

[35]

[36]

[37]

[28]

[32]

[28]

[38]

[39]

[37]

[34]

[28]

When the movement of charges in the medium is sufficiently slow relative to the velocity C, the refractive index is related to the product of dielectric constant € and the magnetic permeability μ as

n{v) = [*(,) μ{νψ*

In most nonpolar fluids μ(ν) & 1; for diamagnetic substances μ differs from 1 by less than 10~5. Hence at optical frequencies c = n* if v is not chosen near a dispersion region (i.e., no energy loss). Under ordinary laboratory conditions [42] using light of wavelength 5893 A (sodium D

"line") the measured refractive index nm is related to the absolute value n by n = 1.00027^m .

(10)

2.5. LORENTZ-LORENZ FUNCTION

The relation between refractive index and the polarizability a at optical frequencies was derived by Lorentz [43] and Lorenz [44] and is of the form

^2 - 1 / V\ 4TTN0

L

"

L =

^ + ^ U

=

- 3 ^

a (2

'

8)

When the refractive index determined at optical frequencies is extrapolated to infinite wavelength (^oo)> the Lorentz-Lorenz (L-L) function for electronic polarization is given by

The determination of η^ by extrapolation from nondispersive regions in the visible or ultraviolet is seldom accurate although often calculated by Cauchy-type relations [5, p. 253]. Extrapolation from nondispersive infrared regions are to be desired, but until recently infrared data were much more difficult to obtain than those in the visible or ultraviolet [45,46].

Of particular interest is a suggestion by Buckingham that dielectric constant and refractive index data may be used together to gain valuable information concerning molecular quadrupole and higher order moments [26].

3* Experimental

3.1. LITERATURE SOURCES

The dielectric constants of more than 800 liquids reported in the literature prior to 1951 are tabulated by Maryott and Smith [47]. Other data may be found in the several excellent treatises dealing with dielectrics [5, 10, 14a, 26, 48-52]. Perhaps most useful are the annual publications of the National Research Council [53] which summarize each year's literature in the field of dielectrics.

The experimental data reported in this section have been obtained subsequent to 1950, and hence supplement the tables of Maryott and Smith. No attempt has been made to report all available data on a given liquid. Those which we feel to be most reliable are listed herein. The tabulation may be considered fairly exhaustive with respect to variety of substances. Perhaps one of the most significant observations to be made concerning this section is the lack of reliable data on nonpolar liquids and

(11)

the need for further measurements of high accuracy on nonpolar liquids of symmetries other than those already reported.

3.2. REFERENCE LIQUIDS

Aside from their obvious value as model systems of high symmetry, certain liquids have received considerable attention as reference sub- stances. These accurately determined dielectric constants may be used for the calibration or testing of equipment.

Among nonpolar fluids, those most often used as reference liquids are benzene, cyclohexane, and carbon tetrachloride. Values of the static or zero-frequency dielectric constant are listed for a variety of temperatures for these liquids in Table II. Although benzene has often been the

TABLE II

REFERENCE LIQUIDS

T ( ° K ) 282 293 293 293 293 298 303 283 293 293 293 293 298 298 298 298 303 303 303 293 293 293 293 293

*o 2.05

2.0230 ± 0.0002 2.0250

2.021 2.021 2.0173 2.010 2.300

2.2832 ± 0.0002 2.2836

2.285 2.2821 2.281 ± 0.006 2.2726 2.2730

2.2747 (100 Mc) 2.2615

2.270 2.263

2.2368 ± 0.0002 2.2365

2.233 2.249 2.249

Ref.

[54]

[55]

[56]

[57]

[58]°

[59]

[60]

[61]

[55]

[56]

[61]

[62]

[63]

[64]

[59]

[65]

[66]

[61]

[60]

[55]

[62]

[67]

[57]

[58]°

a Data are given for other temperatures in the reference cited.

(12)

favored reference liquid of investigators, the difficulty associated with the removal of any final traces of water has led to a variety of values for e0 . Due to its ease of purification, cyclohexane has been suggested [56] as a better reference liquid.

The review by Bauer et ah [42] should be consulted as a source of refractive index data on reference liquids.

3.3. PURE LIQUIDS

In Tables III to VI are listed most of the dielectric data reported in the literature from 1950 to the time of this writing. Unfortunately, a significant portion of the electromagnetic properties is reported at points far removed from a melting point, boiling point, or critical point.

Thus some of the most interesting aspects of present dielectric theory cannot yet be properly tested.

T A B L E III

PROPERTIES OF MOLECULES W I T H SPHERICAL SYMMETRY

Wave- T e0 n C - M L-L length Ref.

(°K) (cm3/mole) (cm3/mole) (A) He 1.2

2.17b 4.215 Ar 83.85b

87.27c 83.8 85.54 150.01 125.00' 150.01' Kr 116.05c

119.80d Xe 161.35c

165.05d 1.0569

1.510 1.504

1.664 1.657 1.880 1.874

1.233

1.2312 ± 0.0002 1.1073 ± 0.0002 1.0122 ± 0.0002 1.0606

0.509 0.516 4.097(1.4152)e 4.126 (1.3985)e

6.222 (2.440)«

6.239 (2.4130)«

9.653 (3.084)«

9.685 (3.06)«

4.17

4.202 (1.394)«

4.118(0.6798)«

4.102(0.07942)«

4.122(0.3870)"

5893 5893 5893

[68]«

[69]«

[69]«

[70]*

[71?

[72]°

[72]«

[72]«

[72]«

[70]«

[70]«

« Data are given for other temperatures and densities in the reference cited.

b A-point.

c Normal melting point.

d Normal boiling point.

e Density in gm/ml.

'Gas.

(13)

The properties of group O elements with spherical symmetry are found in Table III. The refractive index data of Jones and Smith [71]

and Abbiss et al. [72]—after correction for dispersion—are in agreement with the dielectric data of Amey and Cole [70]. The measurements at the boiling point consistently yield a C-M or L-L function which is smaller than that obtained with data from compressed gas data [27].

This is in qualitative agreement with the theory discussed in Section 2.2.3.

Lee et al. [68] have made their measurements on liquid helium over the pressure range 0.2 to 29.5 atm, and the temperature range 0.14° to 1.0°K. The data by Chase and Maxwell [69] cover the range 1.4° to 4.2°K which includes both the λ-point and the normal boiling point.

Data at these temperatures are listed in Table III.

Stewart [73] reports dielectric and density data for parahydrogen from 24° to 100°K and 2 to 326 atm. The calculated C-M function can

TABLE IV

PROPERTIES OF MOLECULES WITH LINEAR SYMMETRY

Ref.

[74]

[75]

F2

Cl2

Br2

pH2

CS2

T (°K)

86°

203 213 223 233 243 273 283 293 303 32 31 293 166 303

*o 1.43 2.142 2.113 2.083 2.054 2.024 3.255 3.198 3.139 3.080 1.25962*

2.634 2.937

C-M (cm3/mole)

4.30 11.94 11.86 11.79 11.73 11.61 21.52 21.43 21.32 21.21 2.0250

21.25

[75]

[73]«

1.082d [76]

[57]

[58?

1.689* [77]

a Normal boiling point.

6 Density is 7.92892 x 105 gm/cm3.

c Data are given for other temperatures in the reference cited.

d At 5460 A.

* At 1330 cm-1.

(14)

be represented by a quadratic function of density in this region. The maximum observed is 0.2 % above the initial C-M value. There is some suggestion of a slight temperature effect at densities greater than 0.06 gm/cm3, but present data are insufficient to confirm this. An extrapolated zero-density molar polarizability is calculated to be 2.02451 cm3. In Table IV is included high-density data measured near the critical point (33.3°K).

Belonogov and Gorbunkov [76] have measured the refractive index of liquid parahydrogen and normal hydrogen over the temperature range 20° to 31°K and at pressures of 1 to 9 atm at wavelengths of 4360, 5460, and 5790 A. They conclude that at equal densities the refractive indices of pH2 and nH2 differ by about 6 X 10~4 whereas at equal temperatures the difference in refractive indices is about 1 X 10~3. They also find that the L-L function is valid for both liquids in the region studied with an accuracy of 5 X 10~5.

The data on liquid halogens by Fröhlich and Jost [75] suggest a density effect at least for chlorine and bromine. The remaining halogen data reported in Table IV consist of one point for fluorine. Once again very few of the measurements are even near a transition temperature.

Phillippe and Piette [58] measured the dielectric constant of carbon disulfide over part of its liquid range as a function of temperature.

Their value near the melting point is included in Table IV, as is also a room temperature measurement reported by Timmermans et al [57].

In Table V are found some measured and derived properties of molecules with tetrahedral symmetry. Decreases in C-M and L-L function have been observed for liquid methane at the boiling point and at the melting point by Amey and Cole [70], and Abbiss et al. [72]. This is qualitatively consistent with theory discussed in Section 2.2.3.

Waxier and Weir [80] have measured the absolute refractive index of benzene and of carbon tetrachloride as a function of pressure, tem- perature, and wavelength. Extrapolation of the data for carbon tetra- chloride to infinite wavelength yields values for the refractive index at zero frequency n^ (see Sec. 2.5). The calculated L-L functions along with their corresponding densities are listed in Table V for the tem- perature 327.50°K. Other temperatures reported by Waxier and Weir are 297.96° and 307.66°K. Pressures ranged from 1 to 1059.2 bars at 327.50°K. The normal boiling point and critical temperature for carbon tetrachloride are 349.81° and 556.2°K, respectively.

Miscellaneous dielectric data for liquid carbon tetrafluoride, silicon tetrachloride, stannic tetrachloride, and titanium tetrachloride are also found in Table V.

(15)

TABLE V

PROPERTIES OF MOLECULES WITH TETRAHEDRAL SYMMETRY

T (°K) CH4 89.15°

111.65*

91.17-180 91.17 185.11 CF4 96.04-100

92.36 158.41 220.40 SiCl4 293 SnCl4 289

291 313 TiCl4 293 CC14 327.50

*0

1.672 1.622

2.2377 ±0.0002 2.881 ±0.0005

2.9031 2.8269 2.814±0.0004

n

1.2930 ±0.0002 1.1596±0.0002

1.2373 ±0.0002 1.1943 ±0.0002 1.1130 ±0.0002

C-M (cm3/mole) 6.456 (0.4550)*

6.510(0.4238)*

33.49 45.06

41.37

L-L (cm3/mole)

6.50d

6.501 (0.4513)*

6.416 (0.2573)*

7.10d 7.071 (1.872)*

7.106(1.542)*

25.86(1.5268)*

Ref.

[70?

[72]

[72?

[72]b [72]

[72?

[72]»

[72]b [78]

[78]

[79]

[78]

[80]b 25.83 (1.5718)*

25.79(1.6020)*

25.75(1.6392)*

25.72(1.6714)*

a Normal melting point.

b Data are given for other temperatures and densities in the reference cited.

c Normal boiling point.

d Wavelength = 5893 A.

* Density in gm/ml.

In Table VI are listed the dielectric properties of three liquids whose molecular symmetries are not included in any of the preceding tables.

Wirth's measurements on diborane [81] cover the entire liquid range.

Because its two hydrogen-bridged three-centered bonds create an ethylene-like structure, the interesting possibility arises of comparing its quadrupole moment with that of ethylene. Further measurements of the refractive index as well as dielectric measurements over a range of pressures are needed. Accurate density determinations would also be valuable.

(16)

TABLE VI

PROPERTIES OF MOLECULES WITH MISCELLANEOUS SYMMETRY

T C-M L-L (°K) €0 (cm3/mole) n (cm3/mole) Symmetry Ref.

Diborane 108.26- 2.3721-

B2He 180.66 0.0027653Γ Ethylenic [81]

SFe 223 1.81 16.5 Octahedral [82]

N204 291 2.42 26.5 1.394 15.2 Planar« [83]

° See text discussion.

It is surprising to note that the single measurement by Berg [82] on sulfur hexafluoride is the only item of dielectric interest available on the liquid. Due to its octahedral symmetry the first permanent electric moment is a hexadecapole moment. Thus, further measurements on this near-spherical molecule could be of considerable theoretical interest.

Liquid dinitrogen tetroxide presents the interesting case where a temperature-dependent equilibrium exists between itself and the monomer, nitrogen dioxide. However, the equilibrium in the liquid phase is strongly in the direction of N204 . At the melting point (262.0°K) there is only 0.01% N 02 present, and at the boiling point (294.3 ΓΚ) this has risen to only 0.1%. A further point of interest is the planar nature of the isomer reported present [84] in liquid N204 . The study by Addison et al. [83] measured the dielectric constant and determined

^oo at 291 °K. The discrepancy between the C-M and L-L functions was attributed largely to the proposed presence of monomer and ion pairs. This liquid warrants a more-detailed dielectric study.

3.4. LIQUID MIXTURES

Few dielectric properties have been measured for mixtures of nonpolar liquids. Of these, most have been binary mixtures. Fewer still follow the elementary additivity behavior suggested by Debye's expression for the molar polarization [85]

where xi is the mole fraction of component i. Decroocq and Jüngers [86]

have shown that the dielectric constant of mixtures of liquids that

(17)

presented in solution their original physical and chemical attributes can be accurately accounted for by the Onsager-Kirkwood theory. Kielich [87]

has reported a theoretical study of molecular refraction of dense binary mixtures. His treatment predicts a deviation from additivity of the refraction for nonpolar mixtures of anisotropically polarizable molecules.

Sackmann and Kehlen [78] have reported measurements made on six binary systems of carbon tetrachloride, silicon tetrachloride, titanium tetrachloride, and tin tetrachloride. Some of these are listed in Table VII.

T A B L E V I I

PROPERTIES OF N O N P O L A R L I Q U I D MIXTURES'1

TiCl4/CCl4

SnCl4/CCl4

SiCl4/CCl4

SnCl4/TiCl4

TiCl4/SiCl4

SnCl4/SiCl4

T (°K)

293

293

293

293

293

293

*0

(mix) 2.413 2.547 2.676 2.417 2.572 2.748 2.2373 2.2382 2.2377 2.833 2.849 2.864 2.368 2.534 2.678 2.387 2.547 2.709

C-M (cm3/mole)

32.14 35.15 38.14 32.83 36.88 41.47 29.47 30.87 32.10 42.36 43.28 44.21 35.46 37.81 39.69 36.46 39.45 42.27

* i6

0.2869 0.5068 0.7402 0.2657 0.5034 0.7765 0.2426 0.5014 0.7359 0.2571 0.5084 0.7708 0.2381 0.5182 0.7627 0.2478 0.5119 0.7563

α From Sackmann and Kehlen [78].

b Data are given for other concentrations in Sackmann and Kehlen [78].

Deviations from simple additivity were observed in CCl4/TiCl4 , CCl4/SnCl4 , and SiCl4/TiCl4 systems. This may be attributed to limited association between dissimilar molecules.

(18)

4* Conclusion

We have noted throughout the sparsity of adequately studied systems.

Admittedly, there are difficulties involved in the measurement of properties in transition regions. However, the significant lack of dielectric data even in "convenient" regions is disturbing.

On the other hand, as an experimentalist it is exciting to look upon a still-fertile field for investigation. Practically every liquid listed in this chapter as well as many not so named are in need of further study as dielectric liquids. Most pressing is the need for thorough and careful measurements of the dielectric constant on a number of liquids as a function of density in the region of their critical point. Larsen et al. [88]

have suggested that the L-L function will become suspect at the critical point for two further reasons: (1) the magnitude of the imaginary part of the refractive index increases significantly as the critical point is approached; (2) density fluctuations at the critical point remove assump- tion that the medium is homogeneous over distances of the order of optical wavelengths. The deviations predicted by Larsen et al. should be more readily observed with molecules having a large polarizability.

Because calculation of both the C-M and L-L functions requires a knowledge of the density, it is important that careful density deter- minations be made on a number of liquids in conjunction with dielectric measurements. An interesting observation made by Stewart [73] is that once it becomes possible to calculate the molar polarization of a liquid from theory, measurement of its dielectric constant will provide a means of obtaining density data for the liquid with a precision greater than that possible by PVT methods.

As a final liquid, let us suggest the study of the dielectric behavior of liquid xenon hexafluoride (melting point = 322.63°K, boiling point = 348.72°K). Such an investigation could shed light on the sug- gestion [89] that molecular association occurs in the liquid. Perhaps more importantly, it would provide needed dielectric data on a distorted octahedral configuration which could not be obtained conveniently elsewhere.

REFERENCES

The survey of literature was completed by September, 1966.

1. C. A. Coulson, Discussions Faraday Soc. 40, 285 (1965).

la. M. Karplus, J. Chem. Phys. 41, 880 (1964).

2. M. Karplus and H. J. Kolker, J. Chem. Phys. 39, 2997 (1963).

3. R. Yaris, J. Chem. Phys. 40, 667 (1964).

4. J. R. Hoyland, J. Chem. Phys. 41, 3153 (1964).

(19)

5. C. J. F. Böttcher, "Theory of Electric Polarization,'' p. 28. Elsevier, Amsterdam, 1952.

6. M. Davies, Discussion Faraday Soc. 40, 279 (1965).

7. O. F. Mossotti, Mem. Math. Fis. Modena 24 / / , 49 (1850).

8. R. Clausius, "Die mechanische Wärmetheorie," Vol. II, p. 62. Vieweg, Braunschweig, 1864-67.

9. H. A. Lorentz, "Theory of Electrons." Teubner, Leipzig, 1909 (also Dover, New York, 1952).

10. R. H. Cole, in "Progress in Dielectrics" (J. B. Birks and J. Hart, eds.), Vol. 3, p. 47.

Heywood, London, 1961.

11. J. G. Kirkwood, J. Chem. Phys. 4, 592 (1936).

12. J. Yvon, "Recherches sur la Theorie Critique des Liquides." Hermann, Paris, 1937.

13. A. D. Buckingham, Discussion Faraday Soc. 40, 171 (1965).

14. W. F. Brown, Jr., J. Chem. Phys. 18, 1193 (1950); 21, 1121 (1923).

14a. W. F. Brown, Jr., in "Handbuch der Physik" (S. Flügge, ed.), Vol. XVII. Springer, Berlin, 1956.

15. J. deBoer, F. van der Maesen, and C. A. Ten Seldam, Physica 19, 265 (1953).

16. M. Mandel a n d P . Mazur, Physica 24, 116 (1958).

17. L. Jansen and P. Mazur, Physica 21, 193 (1955).

18. L. Jansen, Phys. Rev. 112, 434 (1958).

19. L. Jansen and A. D. Solem, Phys. Rev. 104, 1291 (1956).

20. R. Yaris and B. Kirtman, J. Chem. Phys. 37, 1775 (1962).

21. C. A. Ten Seldam and S. R. deGroot, Physica 18, 905 (1952).

22. B. G. Dick, Jr. and A. W. Overhauser, Phys. Rev. 112, 90 (1958).

23. R. L. Amey, J. Phys. Chem. 69, 702 (1965).

24. A. D. Buckingham and J. A. Pople, Trans. Faraday Soc. 51, 1029 (1955); J. Chem.

Phys. 27, 820 (1957).

25. R. W. Zwanzig, J. Chem. Phys. 25, 211 (1956); 27, 821 (1957).

26. A. D. Buckingham, Quart. Rev. (London) 13, 183 (1959).

27. D. R. Johnston, G. J. Oudemans, and R. H. Cole, J. Chem. Phys. 33, 1310 (1960).

28. A. A. Maryott and G. Birnbaum, J. Chem. Phys. 36, 2032 (1962).

29. D. R. Johnston and R. H. Cole, J. Chem. Phys. 36, 318 (1962).

30. A. D. Buckingham, J. Chem. Phys. 23, 412 (1955).

31. W. V. Smith, J. Chem. Phys. 25, 510 (1956).

32. H. Feeny, W. M. Madigosky, and B. Winters, J. Chem. Phys. 27, 898 (1957).

33. A. A. Maryott and S. J. Kryder, J. Chem. Phys. 41, 1580 (1964).

34. R. H. Orcutt, J. Chem. Phys. 39, 605 (1963).

35. J. P. Kopla and J. A. A. Ketelaar, Mol. Phys. 1, 343 (1958).

36. N. J. Harrick and N. F. Ramsey, Phys. Rev. 88, 228 (1952).

37. S. Kielich, Physica 28, 511 (1962).

38. J. Van Kranendonk, Physica 24, 347 (1958).

39. W. S. Benedict and L. D. Kaplan, J. Chem. Phys. 30, 388 (1959).

40. A. G. Turner, B. H. Honig, R. G. Parr, and J. R. Hoyland, / . Chem. Phys. 40, 1919 (1964).

41. S. Kielich, Acta Phys. Polon. 24, 389 (1963).

42. N. Bauer, K. Fajans, and S. Z. Lewin, in "Technique of Organic Chemistry"

(A. Weissberger, ed.), Vol. I, p. 1139. Wiley (Interscience), New York, 1960.

43. H. A. Lorentz, Wied. Ann. Phys. 9, 641 (1880).

44. L. V. Lorenz, Wied. Ann. Phys. 11, 70 (1880).

45. R. E. Kagarise, J. Opt. Soc. Am. 51, 830 (1961).

46. M. Cameo-Bosco, Compt. Rend. 252, 1434 (1961).

(20)

47. A. A. Maryott and E. R. Smith, Table of dielectric constants of pure liquids. Natl.

Bur. Std. (U.S.) Circ. 514 (1951).

48. C. P. Smyth, "Dielectric Behavior and Structure., , McGraw-Hill, New York, 1955.

49. H. Fröhlich, "Theory of Dielectrics," 2nd ed. Oxford Univ. Press, London and New York, (1958).

50. R. J. W. LeFevre, Adv. Phys. Org. Chem. 3, 1 (1965).

51. A. Dalgarno, Advan. Phys. 11, 281 (1962).

52. A. R. von Hippel, "Dielectric Materials and Applications." Wiley, New York, 1954.

53. Committee on Digest of Literature of the Conference on Electrical Insulation. Dig.

Lit. Dielectrics 1 (1936 to date). (Nat. Res. Council, Washington, D.C.) 54. R. W. Crowe and C. P. Smyth, } . Am. Chem. Soc. 73, 5406 (1951).

55. R. Mecke and R. Joeckle, Z. Electrochem. 66, 255 (1962).

56. L. Hartshorn, J. V. L. Parry, and L. Essen, Proc. Phys. Soc. (London) £ 6 8 , 422 (1955).

57. J. Timmermans, A. M. Piette, and R. Philippe, Bull. Soc. Chim. Beiges 64, 5 (1955).

58. R. Philippe and A. M. Piette, Bull. Soc. Chim. Beiges 64, 600 (1955).

59. E. G. Claeys, G. P. Van der Kelen, and Z. Eeckhaut, Bull. Soc. Chim. Beiges 70, 462 (1961).

60. S. Katagiri, Tohoku Imperial Univ., Sei. Repts. 1, Ser. 44, 165 (1960).

61. J. Hurivic and J. Michalczyk, Roczniki Chem. 34, 1423 (1960).

62. E. Trieber, J. Schurz, and H. Koren, Monatsh. Chem. 82, 32 (1951).

63. C. G. Miller and O. Maass, Can. J. Chem. 38, 1606 (1960).

64. J. J. Lindberg, Acta Chem. Scand. 14, 379 (1960).

65. A. S. Brown, P. M. Levin, and E. W. Abrahamson, J. Chem. Phys. 19, 1226 (1951).

66. T . N. Pliev, Zh. Fiz. Khim. 35, 2144 (1961).

67. L. deBrouckere and A. Lecoq-Robert, Bull. Soc. Chim. Beiges 70, 549 (1961).

68. D. M. Lee, H. A. Fairbank, and E. J. Walker, Phys.Rev. 121, 1258 (1961).

69. C. E. Chase and E. Maxwell, Physica 27, 1129 (1961).

70. R. L. Amey and R. H. Cole, J. Chem. Phys. 40, 1^6 (1962).

71. G. O. Jones and B. L. Smith, Phil. Mag. 5, 355 (1960).

72. C. P. Abbiss, C. M. Knobler, R. K. Teague, and C. J. Pings, / . Chem. Phys. 42, 4145 (1965).

73. J. W. Stewart, / . Chem. Phys. 40, 3297 (1964).

74. E. U. Franck, Naturwissenschaften 41, 37 (1954).

75. G. Fröhlich and W. Jost, Chem. Ber. 86, 1184 (1953).

76. A. V. Belonogov and V. M. Gorbunkov, Opt. Spectry. (USSR) (English Transi.) 14, 234 (1963).

77. D. W. Barnes and P. N . Schatz, J. Chem. Phys. 38, 2662 (1963).

78. H. Sackmann and H. Kehlen, Z. Electrochem. 66, 446 (1962).

79. H. Tsubomura, Bull. Chem. Soc. Japan 27, 1 (1954).

80. R. M. Waxier and C. E. Weir, J. Res. Natl. Bur. Std. .468, 489 (1964).

81. H. E. Wirth and E. D. Palmer, J. Phys. Chem. 60, 911 (1956).

82. D. Berg, J. Chem. Phys. 31, 572 (1959).

83. C. C. Addison, H. C. Bolton, and J. Lewis, J. Chem. Soc. p. 1294 (1957).

84. P. Goth, Nature 198, 1081 (1963).

85. L. Hartshorn and I. Saxton, in "Handbuch der Physik" (S. Flügge, ed.), Vol. XVI.

Springer, Berlin, 1958.

86. D. Decroocq and J. C. Jungers, Compt. Rend. 252, 1454 (1961).

87. S. Kielich, Physica 28, 1116 (1962).

88. S. Y. Larsen, R. D. Mountain, and R. Zwanzig, J. Chem. Phys. 42, 2187 (1965).

89. F. Schreiner et al.t 152nd Natl. Meeting Am. Chem. Soc, New York, September 1966.

Ábra

FIG. 1. Effective polarizability (a), harmonic oscillator potential (HO), and Lennard- Lennard-Jones potential (L-J) as a function of intermolecular distance
TABLE I  QUADRUPOLE MOMENTS  Ô x 10 2e  esu cm 2  ±1.7  1.8  4.4  4.59  4.6  5.2  5.3  2.6  3.92  4  &lt;0.3  0.52  0.6  0.63  0.95  &lt;1.0  ±0.80  1.0  1.64  1.8  1.80  1.90  &lt;0.3  Method Pressure broadening Pressure broadening  Microwave absorption
TABLE II  REFERENCE LIQUIDS  T ( ° K )  282  293  293  293  293  298  303  283  293  293  293  293  298  298  298  298  303  303  303  293  293  293  293  293  *o 2.05  2.0230 ± 0.0002 2.0250 2.021 2.021 2.0173 2.010 2.300 2.2832 ± 0.0002 2.2836 2.285 2.28
TABLE IV
+2

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

A heat flow network model will be applied as thermal part model, and a model based on the displacement method as mechanical part model2. Coupling model conditions will

The present paper reports on the results obtained in the determination of the total biogen amine, histamine and tiramine content of Hungarian wines.. The alkalized wine sample

István Pálffy, who at that time held the position of captain-general of Érsekújvár 73 (pre- sent day Nové Zámky, in Slovakia) and the mining region, sent his doctor to Ger- hard

Major research areas of the Faculty include museums as new places for adult learning, development of the profession of adult educators, second chance schooling, guidance

The decision on which direction to take lies entirely on the researcher, though it may be strongly influenced by the other components of the research project, such as the

In this article, I discuss the need for curriculum changes in Finnish art education and how the new national cur- riculum for visual art education has tried to respond to

By examining the factors, features, and elements associated with effective teacher professional develop- ment, this paper seeks to enhance understanding the concepts of

The method discussed is for a standard diver, gas volume 0-5 μ,Ι, liquid charge 0· 6 μ,Ι. I t is easy to charge divers with less than 0· 6 μΐ of liquid, and indeed in most of