Next Article in Journal
Manufacturing of 3D-Printed Hybrid Scaffolds with Polyelectrolyte Multilayer Coating in Static and Dynamic Culture Conditions
Previous Article in Journal
Research on Thermal Stability and Flammability of Wood Scob-Based Loose-Fill Thermal Insulation Impregnated with Multicomponent Suspensions
Previous Article in Special Issue
Analysis of Phase Mismatch for Mercurous Bromide-Based Non-Collinear AOTF Design in Spectral Imaging Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of the Acousto-Optic Figure of Merit and the Maximum Value of the Elasto-Optic Constant of Liquids

Scientific and Technological Centre of Unique Instrumentation RAS, 117342 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2024, 17(12), 2810; https://doi.org/10.3390/ma17122810
Submission received: 5 April 2024 / Revised: 2 May 2024 / Accepted: 6 June 2024 / Published: 8 June 2024
(This article belongs to the Special Issue Acousto-Optical Spectral Technologies (2nd Edition))

Abstract

:
The elasto-optic properties of liquids on the basis of the first principles of acousto-optics were theoretically investigated. A relationship for calculating the elasto-optic constant of liquids using only the refractive index was obtained. The refractive index values corresponding to the maximum elasto-optic constant for polar and nonpolar liquids were determined. Calculations for about 100 liquids were performed and compared with known experimental data. This study significantly extends our understanding of the acousto-optic effect and has practical applications for predicting the elasto-optic constant of a liquid and estimating its wavelength dispersion.

1. Introduction

Acousto-optic devices are used today in many fields, e.g., in laser technology, medicine and optical information-processing systems [1,2,3]. Due to the formation of a phase structure in the medium under the influence of ultrasound, acousto-optic devices allow for the parameters of the light beam to be controlled in real time. It is assumed that the ultrasound frequency is less than the inverse time of the relaxation processes [4]. For the rotation of aspherical molecules due to the ultrasound wave, the relaxation time is about 0.1–10 ns [5]. Therefore, the ultrasound frequency has to be less than 100 MHz, which is reasonably accurate for the majority of practical applications. The greatest success was achieved by using birefringent single crystals (e.g., paratellurite, TeO2) as the medium for the acousto-optic interaction [6]. Such crystals also have a pronounced acoustic anisotropy, which makes it possible to select the optimal propagation directions of light and ultrasound beams for a given task.
The energy efficiency of acousto-optic devices is determined by the coefficient M 2 , which characterizes the acousto-optic figure of merit of the medium in which the diffraction of radiation by ultrasound takes place. This coefficient is a complex function of the components p i j of the elasto-optic tensor [7], the speed of sound V, the density of the medium ρ and the refractive index n. For reasons of simplicity, the effective elasto-optic constant p is used. In this case, the formula for the acousto-optic figure of merit reads as follows [6]:
M 2 = p 2 n 6 ρ V 3 .
Note that all known values of the components p i j of the elasto-optic tensor of crystals and liquids are less than one [8]. Therefore, the value of the effective elasto-optic constant p, which is a combination of the p i j values, is also bounded from above. In our opinion, this fact is of fundamental importance and, therefore, requires a detailed investigation. In this study, the maximum value of the effective elasto-optic constant was determined using liquids as an example. This is not only important from a theoretical point of view but can also increase the energy efficiency of acousto-optic devices by choosing the optimal interaction medium.

2. Theory of the Elasto-Optic Effect in Liquids

2.1. Elasto-Optic Effect

The elasto-optic effect consists of variation in the permittivity ε of the medium caused by its deformation. Let us define an elementary volume in the medium in the form of a cube. Its compression by Δ x along an edge with length L x parallel to the O x axis leads to a displacement of the matter particles u and, as a consequence, to a change in the density Δ ρ and relative deformation S:
u = x Δ x L x = x Δ ρ ρ ,
S = d u d x = Δ ρ ρ .
At the same time, it is known that in the first approximation, the change in the dielectric impermeability Δ B (where B = 1 / ε = 1 / n 2 ) is proportional to the relative strain S, and the proportionality coefficient is the elasto-optic constant p [9]:
Δ B = 1 n 2 1 ( n + Δ n ) 2 2 Δ n n 3 = p S .
It follows from (4) that the ratio of the change in the refractive index Δ n to the relative strain S for a given medium is a constant that is independent of the strain. This constant is sometimes used and is referred to as the elasto-optic coefficient ρ n / ρ (not to be confused with the elasto-optic constant p) [10]:
ρ n ρ ρ Δ n Δ ρ = Δ n S .
By substituting Expression (5) into (4), the relationship between the elasto-optic constant p and the elasto-optic coefficient ρ n / ρ can be obtained [11]:
p = 2 n 3 ρ n ρ .
Note, that sometimes ρ ε / ρ is used instead of ρ n / ρ , and the elasto-optic constant p can be calculated as follows (assuming ε = n 2 ):
ρ ε ρ = 2 n ρ n ρ ,
p = 1 n 4 ρ ε ρ .

2.2. Review of Permittivity Models

The dependence of the permittivity ε on the density ρ is non-linear and also implicitly includes temperature and pressure. In most works, the relationships between ε and ρ , as well as the expressions for the elasto-optic coefficient ρ n / ρ , were determined. Using (8), we derive expressions for ε / ρ and for the elasto-optic constant p. The models can be divided into three groups: (1) the general theories (Lorentz, Onsager, Kirkwood, Proutiere), (2) the simple approximations to the known relationships (Rocard, Looyenga) and (3) the empirical rules for the best fit of some experimental data (Gladstone–Dale, Eykman, Wahid, Meeten).
The simplest model uses the Lorentz–Lorenz Formula [12,13], which is also known as the Clausius–Mossotti relation [14,15]:
ε 1 ε + 2 = 4 π 3 N A α M ρ ρ , ε ρ ( ε + 2 ) 2 3 , p = ( n 2 1 ) ( n 2 + 2 ) 3 n 4 ,
where M is the molar mass, N A is the Avogadro number and α is the molecular polarizability.
Let us emphasize the assumptions in the Lorentz–Lorenz model. First, it is assumed that the molecules have a spherical shape. Second, the dipole polarizability caused by the rotation of dipole molecules is neglected. And finally, the influence of neighboring molecules on each other is not taken into account.
The approximation of the Lorentz–Lorenz model was given by Rocard, who assumed ( ε + 2 ) as a constant in the Lorentz–Lorenz Formula (9) [16]:
ε 1 ρ , ε ρ const , p = n 2 1 n 4 .
One more model that does not involve the shapes of the molecules was developed by Looyenga [17] and suggests the following equation as a simple approximation of the Lorentz–Lorenz relation (9) for ε 1 :
ε 1 / 3 1 ρ , ε ρ ε 2 / 3 , p = 3 ( n 2 / 3 1 ) n 8 / 3 .
The Onsager formula [18,19], which is referred to as Oster’s rule in some works [20,21], is used to calculate the dielectric constant of liquids. In this model, the molecules are assumed to be a polarizable point dipole located at the center of a spherical cavity (the Onsager cavity):
( ε 1 ) ( 2 ε + 1 ) ε ρ , ε ρ ε 2 2 ε 2 + 1 , p = ( n 2 1 ) ( 2 n 2 + 1 ) ( 2 n 4 + 1 ) n 2 .
To provide a more precise description of polar liquids, the Kirkwood model introduces a g-factor to consider the short-range intermolecular interaction [22,23], resulting in the following expressions [24]:
ε 1 ρ ( 1 + a ρ ) , p = ( n 2 1 ) ( 2 n 2 + 1 ) ( n 2 + 2 ) n 4 .
In refining the Kirkwood model, Niedrich assumed that the molecules are not identical and that the local electric field is determined solely by the dielectric constant, independent of temperature and density [24]:
( ε 1 ) ( 2 ε + 1 ) ε ρ exp ( b ρ 2 ) , p = ( n 2 1 ) ( 2 n 2 + 1 ) ( n 2 + 2 ) n 4 3 ( n 4 + 2 ) ( 2 n 2 + 1 / n 2 ) ( n 2 + 2 ) .
Proutiere in [25,26,27] showed that the approximation of the local electric field must be made after averaging the molecular dipole moment and that this moment is independent of density fluctuations [28]. It was established that this model gives more accurate results than others [29,30]:
ε 1 ρ ε α 2 ε ¯ + 1 2 ( ε ¯ 1 ) φ ρ ( α N A / 3 ε 0 M ) ,
p = n 2 1 n 4 1 + 2 ( n 2 1 ) 2 φ 3 ( 2 n 2 + 1 ) 6 ( n 2 1 ) 2 ( φ 1 ) / ( 2 n 2 + 1 ) ,
where ε 0 is the dielectric constant of a vacuum, the bar indicates a mean value in the bulk liquid that surrounds the Onsager cavity and φ = 6 / π 2 is the inverse value of a part of the available space filled by the molecular Onsager cavities.
As can be seen, numerous models describing the elasto-optic effect have been developed to date. It is worth mentioning empirical relations: Gladstone–Dale rule [21,31]:
ε 1 ρ , ε ρ ε , p = 2 ( n 1 ) n 3 ,
Eykman’s rule [26], which is applicable for organic solvents within the refractive index range 1.35 < n < 1.5 [32]:
ε 1 ε + 0.4 ρ , ε ρ ε ( ε + 0.4 ) 2 ε + 0.8 ε + 1 , p = 2 ( n 2 1 ) ( n + 0.4 ) n 3 ( n 2 + 0.8 n + 1 ) ,
Wahid’s rule [16] for 1.3 < n < 1.6:
ε 1 ε 1 / 3 ρ , ε ρ ε 4 / 3 2 ε + 1 , p = 3 ( n 2 1 ) n 2 ( 2 n 2 + 1 ) .
and Meeten’s rule for temperatures from −50 °C to +35 °C, for pressures from 1 to 10 3 bars and over the whole visible spectrum [32]:
ρ ε ρ = ( ε 1 ) ( 7 ε + 23 ) 30 , p = ( n 2 1 ) ( 7 n 2 + 23 ) 30 n 4 .
Usually, the models under consideration are used to calculate the permittivity ε or molecular polarizability. Verification of the models (excluding the Proutiere model) in the experiment on compressing liquids to a pressure of 14 kbar revealed that the Lorentz–Lorenz model has the highest error rate (approximately 10%), while the Niedrich formula has the lowest (about 5%) [24,33]. It was demonstrated [29] that the Proutiere formula provides a more accurate result (with an error of about 2%) for polar liquids, such as water, compared with the Niedrich formula. At the same time, the error in Meeten’s rule is within 1% relative to experimental data for most liquids [32].
The main theoretical problem is to determine the local electric field of the molecules, taking into account the shape of the molecule and the influence of the neighboring molecules. The only way to estimate the applicable range of the model is to compare the consequences of the theory with experimental data: (1) the pressure dependence of the refractive index [24,33], (2) the light scattering in a liquid due to thermal fluctuations in the permittivity [29] and (3) the pressure dependence of the elasto-optic constant [26,32]. But even if a model gives the correct result for one physical parameter (molecular polarizability, permittivity), it does not mean that the model is correct for another physical parameter (elasto-optic constant). In fact, the elasto-optic constant comparison derived from the most known models is performed here for the first time. The results are summarized in the following section.

3. Comparison of Permittivity Models

The models work fine to give an accurate relationship between the permittivity ε and the density ρ of the medium. At the same time, it is interesting to compare all these models in detail with regard to the calculation of the elasto-optic constant p. For this purpose, the theoretical dependencies p ( n ) , as well as experimental data from [15,26], are presented in Figure 1. It should be noted that the data for the same fluid scatter by 5–10%. Furthermore, the dynamic elasto-optic constant could even be only half as large as that determined under static conditions [34,35]. Nevertheless, as one can see, the data agree qualitatively with the models.
According to (9)–(19), the density derivative of the permittivity ε / ρ at ε 1 is estimated as (1) a constant ( ε / ρ const in the Rocard and Onsager models), which is determined by the polarizability of the medium; (2) a linear dependence on the refractive index ( ε / ρ n in the Gladstone–Dale and Eykman models); and even (3) a significantly nonlinear function ( ε / ρ n 4 in the Lorentz–Lorenz model). It should also be noted that only in the Lorentz–Lorenz (9) and Meeten (20) models, there is a plateau of p ( n ) at n 2 . However, there is a common feature in the considered dependencies of the elasto-optic constant p on the refractive index n. In most of the considered models, the elasto-optic constant p increases linearly with the refractive index n according to the law p n 1 , and at n 2 , the elasto-optic constant decreases as p 1 / n 2 . For example, in (13), (17) and (18),
p 2 ( n 1 ) at n 1 1 2 / n 2 at n 2 .
The analysis of the models allowed us to determine the value of the refractive index n opt corresponding to the maximum value of the elasto-optic constant max(p). The results are summarized in Table 1.
One can see from Figure 1 that the Proutiere relation (16) describes the experimental data the most accurately, while the Lorentz–Lorenz (9) and Rocard (10) models are the crudest approximations. At the same time, the Proutiere model predicts a clearly different dependence of the elasto-optic constant p ( n ) on the refractive index at n > 1.6 . According to Formula (16), the elasto-optic constant of liquids with a high optical density is an increasing function of the refractive index, while according to other models, in contrast, it is a decreasing function. As the Lorentz–Lorenz formula was one of the first models, it gives significant errors. At the same time, the Proutiere relation is a result of the much more modern theory with an accuracy of about a few percent. Certainly, there is the significant difference between these and other models at n > 1.6 , but which ones are correct could not be clarified owing to a lack of experimental data.
It is important to note that the Lorentz–Lorenz model predicts the largest value of the elasto-optic constant and all experimental values of p are below this dependence (9). This fact was also established by Niedrich [24]. Therefore, Formula (9) can be used to estimate the maximum value of the elasto-optic constant p.

4. AO Figure of Merit of Liquids

The result obtained in the previous section is very important: the optimal value of the refractive index n opt of the liquid at which the elasto-optic constant p is a maximum was determined. However, for practical applications, the value of p is not the only one can be used as a criterion for the selection of the optimal acousto-optic interaction medium. As follows from (1), the energy efficiency of the AO devices is determined by the AO figure of merit M 2 , which is a complex parameter. Combining (1) and (6), one can obtain the following relation for M 2 :
M 2 = p 2 n 6 ρ V 3 = 4 ρ V 3 ρ n ρ 2 .
A review of the data [36] for more than 250 liquids allowed us to plot a density–velocity diagram (see Figure 2) in which the points correspond to different liquids. As can be seen, it cannot be said that a denser liquid is characterized by a greater value of the speed of sound. Therefore the further considerations in this section only apply to a single liquid. In this case, one can assume that the speed of sound is proportional to the density of the medium [37]:
V = w ( 1 + v ρ ) ,
where w and v are coefficients that are environment-specific and have the same sign ( w > 0 , v > 0 or w < 0 , v < 0 ).
Under the Lorentz–Lorentz approximation, the relations (1), (9) and (23) allow one to write the expression for the acousto-optic figure of merit in the following way:
M 2 = v w 3 1 v ρ ( v ρ 1 ) 3 ( n 2 1 ) 2 ( n 2 + 2 ) 2 9 n 2
Combining (24) and the relation (9) between density ρ and permittivity ε leads to the following expression for the acousto-optic figure of merit (see Figure 3):
M 2 = v w 3 1 v ρ max [ ( v ρ max 1 ) n 2 ( 2 + v ρ max ) ] 3 ( n 2 1 ) 2 ( n 2 + 2 ) 2 9 n 2 ,
ρ max = 3 M 4 π N A α ,
where ρ max is the maximal value of ρ according to the Lorentz–Lorenz model (9) at infinite permittivity ε [17].
As can be seen from Figure 3, the shape of the graph M 2 ( n ) is determined by the sign of coefficients in the dependence V ( ρ ) (23). For liquids under normal conditions, as follows from [37], the free term of the linear approximation V ( ρ ) is negative. At the same time, the dependence V ( ρ ) for liquefied gases is essentially non-linear. Therefore, the slope angle and, consequently, the free term of the linear approximation depend on the operating point (temperature and pressure). Note, that infinite growth of the acousto-optic figure of merit M 2 with a decrease in the refractive index n is related to the extremely small sound velocity V, as follows from (23) and (25).
There are a few peculiarities that should be mentioned for the dependence of the normalized acousto-optic figure of merit on n and v ρ max . For v ρ max < 0 , the optimal value of the refractive index n opt can be found as follows:
n opt = 10 7 v ρ max + 3 ( v ρ max ) 2 28 v ρ max + 4 2 ( 5 v ρ max 4 ) ,
where the relation for the maximal value of M 2 can be easily obtained, but it is not presented because of its highly complicated nature.
It is interesting that at a high refractive index n 1 , the acousto-optic figure of merit M 2 1 / ρ max w 3 ( v ρ max 1 ) 3 seems to be independent of n and depends only on the density and sound velocity parameters. However, this is not the case, as ρ max depends on the polarizability α , which is an optical parameter that varies with temperature and pressure. This means a limitation for the model used.

5. Discussion

The acousto-optic figure of merit M 2 is a complex parameter of acoustic, optic and elasto-optic properties of the medium, i.e., four items: V, ρ , n and p. At the same time, the models of permittivity give us simple relations p = p ( n ) , reducing the number of parameters to three. As shown in the previous section, the structure of the relationship in (25) allows one to reduce everything to the analysis of the function of only two variables M 2 w 3 / v = f ( v ρ max , n ) . However, it should be noted that there are forbidden zones on this dependence since not any combination of medium parameters can be found in nature. This aspect is very important and requires further consideration, but is beyond the scope of this paper. Nevertheless, from the obtained dependencies (see Figure 3) one can draw a general conclusion that from two liquids with the same optical ( n ) and acoustic properties ( w , v ) , the liquid with the larger value of the molecule polarizability α is more preferable. This study reviewed the literature on optical, acoustic and elasto-optic properties, as well as acousto-optic figure of merit, and the collected data for about 100 liquids are presented in Table A1 in Appendix A.
Since the refractive index n depends on the wavelength λ , the dispersion of the elasto-optic constant p can be estimated using (21):
d p d λ 2 d n d λ at n 1 1 4 n 3 d n d λ at n 2 .
To calculate the dispersion of the elasto-optic constant p, the experimental Eykman’s rule (18) was used, as well as data on the refractive index dispersion n ( λ ) of both polar and nonpolar liquids. Data for alkanes were related to the range 0.32–0.65 μ m at a temperature of 300 K [38], whereas data for alcohols were obtained in the much wider range 0.45–1.55 μ m at the same temperature [39]. In addition, data for liquid xenon for the range 0.18–0.65 μ m at 162.35 K were taken from [40]. The dynamics of p ( λ ) depended on the relationship between the refractive index n and its optimal value n opt for maximal p (see Table 1): (1) for n < n opt , the elasto-optic constant p decreased with the wavelength λ ; (2) for n > n opt , the elasto-optic constant increased with λ ; and (3) for n n opt , the elasto-optic constant did not depend on λ . The results are shown in Figure 4.
For the crystalline media, the components of the elasto-optic tensor were calculated on the basis of density functional theory to determine the structure of the electronic zone [41,42]. The above method was only implemented numerically, and therefore, there was no analytical formula for the p i j components. It is worth mentioning that a method for estimating the p 11 and p 12 components for electro-optic crystals based on electro-optic, optical and acoustic properties was recently proposed [43]. However, this method has not yet found wide application and was only validated for one crystal. At the same time, for liquids, as can be seen from the relationships in Section 2.2, the elasto-optic constant p depended only on the refractive index n. Therefore, now (1) it is possible to calculate the elasto-optic constant using a simple relationship and (2) the analysis of the results is reduced to the study of a function of one variable. In our opinion, it is one of the fundamental laws of nature that the function p ( n ) is bounded. However, this law has so far only been shown for liquids in this study.
Let us try to give a physical interpretation of the limitation of the maximum value of the elasto-optic constant p. For this purpose, let us use the original relation (4), which is the definition of this constant, and rewrite it so that it contains relative physical quantities:
Δ n n = p n 2 2 S .
As can be seen from the literature data, for most liquids, 1.3 < n < 1.6 in the range of transparency (see Figure 1), while the maximum refractive index is n = 2.1 for the liquid Se2Br2 [44] (the experimental value of p, however, is unknown). At the same time, the maximum value of the elasto-optical constant is limited to p < 0.375 (see Table 1). Finally, we arrive at the following relationship: Δ n / n < 0.8 S , i.e., the relative change in the refractive index is less than the relative strain.

6. Conclusions

This article deals with the basics of the theory of the elasto-optic effect in liquids. Starting from first principles, an expression for the acousto-optic figure of merit is derived, which allows one to estimate the optimal ratios of the material parameters of the medium. The physical properties, as well as the calculated acousto-optic figure of merit, of about 100 liquids are summarized. Formulas for estimating the elasto-optic constant from the known refractive index are derived. It was found that the elasto-optic constant of any liquid cannot be greater than 0.375. This fact is of fundamental importance, as it provides a qualitative explanation as to why the elasto-optic constants of known liquid and crystalline media are less than one. In the future, it is planned to develop the proposed method for isotropic dielectrics.

Author Contributions

Conceptualization, P.A.N. and V.E.P.; methodology, P.A.N.; formal analysis, V.E.P.; investigation, P.A.N.; writing—original draft preparation, P.A.N.; writing—review and editing, V.E.P.; visualization, P.A.N.; supervision, V.E.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education of the Russian Federation under State contract no. FFNS-2022-0009.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within this article.

Acknowledgments

This work was done with the use of equipment of the Shared Research Facilities of the Scientific and Technological Centre of Unique Instrumentation of the Russian Academy of Sciences (STC UI RAS).

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

The data on the refractive index n, the density ρ and the experimental values of the elasto-optic coefficient ( ρ n / ρ ) of the liquids were taken from a single source for each liquid and systematized in Table A1, namely, [15] at 25 °C and [26] at 20 °C. It should be noted that the experimental data on the elasto-optic constant were obtained only from [15], directly from the acousto-optic interaction, whereas in [26], it was from the temperature dependencies of the density and refractive index:
ρ n ρ = ρ n / T ρ / T ,
where T is the temperature.
The experimental value of the elasto-optic constant p was calculated from the values of the elasto-optic coefficient ( ρ n / ρ ) using the known relationship (6). Table A1 also summarizes the results of our calculations of the elasto-optic constant p using the Lorentz–Lorenz formula (9) and Eykman’s rule (18). We also estimated the acousto-optic figure of merit M 2 for each liquid using Formula (1).
Table A1. Optical, acoustical and acousto-optical properties of liquids.
Table A1. Optical, acoustical and acousto-optical properties of liquids.
Mediumn ρ
(g/cm)3
V
(m/s)
p
Exp
p
Eykman
p
Lorentz
M 2 · 10 15
(s3/kg)
Exp
M 2 · 10 15
(s3/kg)
Eykman
M 2 · 10 15
(s3/kg)
Lorentz
1,1,1-Trichloroethane1.438 [26]1.338 [26]943 [45]0.282 [26]0.3130.339625771903
1,1,2,2-Tetrachloroethane1.494 [26]1.595 [26]1170 [46]0.307 [26]0.3160.349411435530
1,2-Dichloroethane1.445 [26]1.253 [26]1216 [46]0.293 [26]0.3140.340347397467
1-Butanol1.399 [26]0.810 [26]1257 [46]0.302 [26]0.3090.330426445507
1-Chlorobutane1.402 [26]0.886 [26]1117 [46]0.296 [26]0.3090.330538587671
1-Chloropropane1.388 [26]0.891 [26]1091 [46]0.293 [26]0.3070.327530582659
1-Hexane1.388 [26]0.673 [26]1098 [46]0.305 [26]0.3070.327745756856
1-Nitropropane1.402 [26]1.001 [26]1252 [47]0.275 [26]0.3090.330292368421
2-Aminoethanol1.454 [26]1.016 [26]1741 [46]0.302 [26]0.3140.342161174206
2-Ethoxyethyl Acetate1.404 [26]0.973 [26]1129 [46]0.374 [26]0.3090.331768525601
2-Methyl-1-Propanol1.396 [26]0.802 [26]1188 [48]0.302 [26]0.3080.329502523595
2-Nitropropane1.394 [26]0.988 [26]1201 [47]0.272 [26]0.3080.328317407462
3-Methyl-1-Butanol1.407 [26]0.810 [26]1220 [46]0.289 [26]0.3100.332442506580
a-Bromonaphthalene1.649 [15]1.479 [15]1376 [15]0.326 [15]0.3120.366553507698
Acetaldehyde1.331 [26]0.778 [26]1137 [46]0.280 [26]0.2950.309381424465
Acetic Acid1.372 [26]1.049 [26]1134 [46]0.264 [26]0.3040.322303403452
Acetic Anhydride1.390 [26]1.081 [26]1249 [46]0.272 [26]0.3070.327254324368
Acetone1.356 [15]0.785 [15]1170 [15]0.298 [15]0.3010.317438448498
Acetonitrile1.344 [26]0.782 [26]1300 [46]0.268 [26]0.2980.314247306337
a-Chloronaphthalene1.626 [15]1.188 [15]1454 [15]0.327 [15]0.3130.364541497671
Acrolein1.402 [26]0.839 [26]1207 [49]0.300 [26]0.3090.330462491561
Acrylonitrile1.392 [26]0.806 [26]1172 [50]0.292 [26]0.3070.328478529601
Aniline1.579 [15]1.017 [15]1639 [15]0.332 [15]0.3160.360382345448
Anisole1.517 [26]0.994 [26]1425 [46]0.306 [26]0.3170.352397425526
Benzene1.495 [15]0.874 [15]1290 [15]0.329 [15]0.3160.349643595725
Benzonitrile1.528 [26]1.005 [26]1603 [51]0.308 [26]0.3170.354291308385
Bicyclohexyl1.480 [26]0.886 [26]1422 [46]0.336 [26]0.3160.347466411495
Bromobenzene1.552 [15]1.488 [15]1160 [15]0.320 [15]0.3160.357616602766
Bromoform1.591 [15]2.877 [15]921 [15]0.325 [15]0.3150.361764717940
Butyl Acetate1.394 [26]0.881 [26]1226 [46]0.299 [26]0.3080.328403428487
Butyric Acid1.398 [26]0.958 [26]1203 [46]0.271 [26]0.3080.329328426485
Carbon Disulphide1.617 [15]1.256 [15]1144 [15]0.338 [15]0.3140.36310879371255
Carbon Tetrachloride1.460 [26]1.594 [26]936 [46]0.280 [26]0.3150.343583734873
Carbon Tetrachloride1.456 [15]1.584 [15]922 [15]0.331 [15]0.3140.342841758899
Chlorobenzene1.525 [26]1.106 [26]1289 [46]0.311 [26]0.3170.353513532662
Chlorobenzene1.519 [15]1.101 [15]1270 [15]0.330 [15]0.3170.353593546677
Chloroform1.446 [26]1.489 [26]1001 [46]0.314 [26]0.3140.340601601708
Chloroform1.444 [15]1.480 [15]978 [15]0.305 [15]0.3130.340609643756
Cinnamaldehyde1.611 [15]1.049 [15]1555 [15]0.341 [15]0.3140.363514438583
Crotonaldehyde1.437 [26]0.852 [26]1344 [49]0.305 [26]0.3130.338397417489
Cyclohexane1.426 [26]0.779 [26]1279 [46]0.307 [26]0.3120.336486503584
Cyclohexylamine1.459 [26]0.867 [26]1430 [52]0.321 [26]0.3150.343393377448
Decane1.412 [26]0.730 [26]1253 [46]0.309 [26]0.3100.333525530610
Dichloromethane1.424 [26]1.326 [26]1092 [46]0.280 [26]0.3120.336380469545
Dimethoxymethane1.353 [26]0.860 [26]1146 [46]0.269 [26]0.3010.317344429476
Dodecane1.422 [26]0.749 [26]1298 [46]0.308 [26]0.3110.335478489566
Ethanol1.361 [26]0.789 [26]1160 [46]0.287 [26]0.3020.319427472527
Ethyl Acetate1.372 [26]0.901 [26]1162 [46]0.284 [26]0.3040.322382438491
Ethyl Acetate1.370 [15]0.895 [15]1148 [15]0.281 [15]0.3040.322385451505
Ethyl Alcohol1.359 [15]0.785 [15]1146 [15]0.280 [15]0.3020.318417485540
Ethyl Butyrate1.393 [26]0.879 [26]1197 [46]0.292 [26]0.3080.328413458521
Ethyl Formate1.360 [26]0.923 [26]1136 [53]0.248 [26]0.3020.319288426475
Ethyl Malonate1.414 [26]1.055 [26]1267 [54]0.277 [26]0.3100.333286358413
Ethyl Oxalate1.410 [26]1.079 [26]1276 [55]0.278 [26]0.3100.332272337388
Ethyl Propionate1.384 [26]0.890 [26]1183 [46]0.277 [26]0.3060.326365447505
Ethylenediamine1.457 [26]0.897 [26]1672 [56]0.414 [26]0.3140.342391226268
Formic Acid1.371 [26]1.220 [26]1287 [46]0.290 [26]0.3040.322215237265
iso-Butil Alcohol1.390 [15]0.798 [15]1194 [15]0.287 [15]0.3070.327436501569
Methyl Benzoate1.517 [26]1.089 [26]1380 [57]0.300 [26]0.3170.352384426528
Methyl Oleate1.452 [26]0.874 [26]1425 [58]0.299 [26]0.3140.341331366432
Methylene Iodide1.732 [15]3.308 [15]962 [15]0.332 [15]0.3050.37010128511257
n-Butil Alcohol1.395 [15]0.806 [15]1245 [15]0.309 [15]0.3080.329451449512
n-Heptane1.388 [26]0.684 [26]1152 [46]0.310 [26]0.3070.327656643729
n-Hexane1.375 [26]0.659 [26]1098 [46]0.308 [26]0.3050.323735719808
n-Hexane1.372 [15]0.655 [15]1090 [15]0.322 [15]0.3040.322816728817
Nitrobenzene1.552 [26]1.203 [26]1475 [46]0.298 [26]0.3160.357323363461
Nitrobenzene1.546 [15]1.198 [15]1448 [15]0.331 [15]0.3160.356412376476
Nitroethane1.392 [26]1.051 [26]1272 [59]0.288 [26]0.3080.328280318362
Nitromethane1.381 [26]1.138 [26]1338 [46]0.281 [26]0.3060.325202238269
Nonane1.405 [26]0.718 [26]1297 [46]0.309 [26]0.3090.331470471540
n-Pentane1.358 [26]0.626 [26]1030 [46]0.273 [26]0.3010.318681830924
o-Dichlorobenzene1.552 [26]1.306 [26]1296 [46]0.309 [26]0.3160.357468491624
Oleic Acid1.460 [26]0.891 [26]1400 [60]0.298 [26]0.3150.343352392466
o-Toluidine1.565 [15]0.994 [15]1598 [15]0.326 [15]0.3160.358385362465
Phenil hydrazine1.599 [15]1.094 [15]1716 [15]0.334 [15]0.3150.362337300396
Phosphorus tribromide1.690 [15]2.861 [15]930 [15]0.331 [15]0.3090.36811079641373
Piperidine1.453 [26]0.860 [26]1400 [46]0.234 [26]0.3140.342218393464
Propionaldehyde1.362 [26]0.797 [26]1379 [61]0.280 [26]0.3020.319240279311
Propionic Acid1.387 [26]0.993 [26]1166 [62]0.266 [26]0.3070.326320425481
Propionitrile1.366 [26]0.782 [26]1271 [46]0.267 [26]0.3030.320289371415
Propyl Acetate1.384 [26]0.888 [26]1198 [46]0.276 [26]0.3060.326350433489
Quinoline1.615 [15]1.092 [15]1567 [15]0.331 [15]0.3140.363464416557
Styrene1.547 [26]0.906 [26]1354 [46]0.291 [26]0.3160.356516610773
Tetrachloroethylene1.506 [26]1.623 [26]1053 [46]0.306 [26]0.3160.351577616756
Trans-Decahydronaphthalene1.469 [26]0.870 [26]1398 [63]0.317 [26]0.3150.345426421503
Trichloroethylene1.478 [26]1.468 [26]1049 [46]0.314 [26]0.3160.346606611736
Trichloroethylene1.472 [15]1.460 [15]1021 [15]0.305 [15]0.3150.345608651780
Trycresyl Phosphate1.549 [15]1.172 [15]1502 [15]0.321 [15]0.3160.356358348442
Water1.330 [15]0.997 [15]1491 [15]0.273 [15]0.2950.309125146160
Water1.333 [26]0.998 [26]1482 [46]0.360 [26]0.2960.310224151166

References

  1. Schrodel, Y.; Hartmann, C.; Zheng, J.; Lang, T.; Steudel, M.; Rutsch, M.; Salman, S.H.; Kellert, M.; Pergament, M.; Hahn-Jose, T.; et al. Acousto-optic modulation of gigawatt-scale laser pulses in ambient air. Nat. Photonics 2023, 18, 54–59. [Google Scholar] [CrossRef]
  2. Omidali, M.; Mardanshahi, A.; Särestoniemi, M.; Zhao, Z.; Myllyla, T. Acousto-Optics: Recent studies and medical applications. Biosensors 2023, 13, 186. [Google Scholar] [CrossRef]
  3. Pozhar, V.E.; Bulatov, M.F.; Machikhin, A.S.; Shakhnov, V.A. Technical implementation of acousto-optical instruments: Basic types. J. Phys. Conf. Ser. 2019, 1421, 012058. [Google Scholar] [CrossRef]
  4. Klein, W.; Cook, B. Unified Approach to Ultrasonic Light Diffraction. IEEE Trans. Sonics Ultrason. 1967, 14, 123–134. [Google Scholar] [CrossRef]
  5. Klein, W.; Fitts, W. Optical Diffraction by an Acoustically Oriented Liquid. IEEE Trans. Sonics Ultrason. 1974, 21, 204–208. [Google Scholar] [CrossRef]
  6. Zhang, H.; Zhao, H. Accurate design of a TeO2 noncollinear acousto-optic tunable filter with refractive index correction. Opt. Lett. 2023, 48, 3395. [Google Scholar] [CrossRef]
  7. Sando, D.; Yang, Y.; Bousquet, E.; Carretero, C.; Garcia, V.; Fusil, S.; Dolfi, D.; Barthelemy, A.; Ghosez, P.; Bellaiche, L.; et al. Large elasto-optic effect and reversible electrochromism in multiferroic BiFeO3. Nat. Commun. 2016, 7, 10718. [Google Scholar] [CrossRef]
  8. Martienssen, W.; Warlimont, H. (Eds.) Springer Handbook of Condensed Matter and Materials Data; Springer: Berlin/Heidelberg, Germany, 2005; p. 1121. [Google Scholar] [CrossRef]
  9. Rinaldi, D.; Natali, P.P.; Montalto, L.; Davì, F. The refraction indices and Brewster law in stressed isotropic materials and cubic crystals. Crystals 2021, 11, 1104. [Google Scholar] [CrossRef]
  10. Li, J.; Zhang, H.; Chen, X.; Le, T.; Wei, H.; Li, Y. High-speed non-contact measurement of elasto-optic coefficient via laser-induced phonons. Appl. Phys. Lett. 2022, 121, 251102. [Google Scholar] [CrossRef]
  11. Uchida, N. Comment on “On the efficiency of liquids in acoustooptic devices”. Jpn. J. Appl. Phys. 1972, 11, 415–416. [Google Scholar] [CrossRef]
  12. Kragh, H. The Lorenz-Lorentz formula: Origin and early history. Substantia 2018, 2, 7–18. [Google Scholar] [CrossRef]
  13. Neumaier, L.; Schilling, J.; Bardow, A.; Gross, J. Dielectric constant of mixed solvents based on perturbation theory. Fluid Phase Equilibria 2022, 555, 113346. [Google Scholar] [CrossRef]
  14. Talebian, E.; Talebian, M. A general review on the derivation of Clausius–Mossotti relation. Optik 2013, 124, 2324–2326. [Google Scholar] [CrossRef]
  15. Uchida, N. Elastooptic coefficient of liquids determined by ultrasonic light diffraction method. Jpn. J. Appl. Phys. 1968, 7, 1259–1266. [Google Scholar] [CrossRef]
  16. Wahid, H. Molecular scattering of light in pure liquids. J. Opt. 1995, 26, 109–121. [Google Scholar] [CrossRef]
  17. Weiss, L.; Tazibt, A.; Tidu, A.; Aillerie, M. Water density and polarizability deduced from the refractive index determined by interferometric measurements up to 250 MPa. J. Chem. Phys. 2012, 136, 124201. [Google Scholar] [CrossRef]
  18. Onsager, L. Electric moments of molecules in liquids. J. Am. Chem. Soc. 1936, 58, 1486–1493. [Google Scholar] [CrossRef]
  19. Maribo-Mogensen, B.; Kontogeorgis, G.M.; Thomsen, K. Modeling of dielectric properties of complex fluids with an equation of state. J. Phys. Chem. B 2013, 117, 3389–3397. [Google Scholar] [CrossRef]
  20. Oster, G. The scattering of light and its applications to chemistry. Chem. Rev. 1948, 43, 319–365. [Google Scholar] [CrossRef]
  21. Abdel-Azim, A.A.A.; Munk, P. Light scattering of liquids and liquid mixtures. 1. Compressibility of pure liquids. J. Phys. Chem. 1987, 91, 3910–3914. [Google Scholar] [CrossRef]
  22. Mandel, M. A statistical approach to the Onsager and Kirkwood dielectric theory for liquids of rigid dipoles. Physica 1972, 57, 141–151. [Google Scholar] [CrossRef]
  23. Zhang, C.; Hutter, J.; Sprik, M. Computing the Kirkwood g-factor by combining constant Maxwell electric field and electric displacement simulations: Application to the dielectric constant of liquid water. J. Phys. Chem. Lett. 2016, 7, 2696–2701. [Google Scholar] [CrossRef]
  24. Niedrich, Z. Alternative optical equation for dielectric liquids. Phys. Rev. E 1999, 60, 4099–4104. [Google Scholar] [CrossRef]
  25. Proutiere, A. Agreement between molecular polarizability anisotropies (γ2) deduced from Rayleigh light scattering and static Kerr birefringence in liquids—Experimental and theoretical aspects. Mol. Phys. 1988, 65, 499–512. [Google Scholar] [CrossRef]
  26. Proutiere, A.; Megnassan, E.; Hucteau, H. Refractive index and density variations in pure liquids: A new theoretical relation. J. Phys. Chem. 1992, 96, 3485–3489. [Google Scholar] [CrossRef]
  27. Hucteau, H.; Proutiere, A. Refractive index and density variations in liquids. Progress in theoretical formulation. J. Mol. Liq. 1994, 62, 93–112. [Google Scholar] [CrossRef]
  28. Bot, A. Comment on “Refractive index variations in pure liquids: A new theoretical relation”. J. Phys. Chem. 1993, 97, 2804. [Google Scholar] [CrossRef]
  29. Zhang, X.; Hu, L. Estimating scattering of pure water from density fluctuation of the refractive index. Opt. Express 2009, 17, 1671. [Google Scholar] [CrossRef]
  30. Evain, K.; Illien, B.; Chabanel, M.; Beignon, M. Comparison of Proutière and Carr-Zimm theories for static light-scattering molecular weight determination: Application to surfactants in solution. J. Phys. Chem. B 2007, 111, 1597–1603. [Google Scholar] [CrossRef]
  31. Dewaele, A.; Eggert, J.H.; Loubeyre, P.; Le Toullec, R. Measurement of refractive index and equation of state in dense He, H2, H2O, and Ne under high pressure in a diamond anvil cell. Phys. Rev. B 2003, 67, 094112. [Google Scholar] [CrossRef]
  32. Meeten, G.H. Empirical relation for the isothermal density derivative of the optical dielectric constant. Nature 1968, 218, 761. [Google Scholar] [CrossRef]
  33. Vedam, K.; Limsuwan, P. Piezo- and elasto-optic properties of liquids under high pressure. II. Refractive index vs density. J. Chem. Phys. 1978, 69, 4772–4778. [Google Scholar] [CrossRef]
  34. Ramesh, K. Photoelasticity. In Springer Handbooks; Springer: New York, NY, USA, 2008; pp. 701–742. [Google Scholar] [CrossRef]
  35. Sapozhnikov, O.A.; Maxwell, A.D.; Bailey, M.R. Modeling of photoelastic imaging of mechanical stresses in transparent solids mimicking kidney stones. J. Acoust. Soc. Am. 2020, 147, 3819–3829. [Google Scholar] [CrossRef]
  36. Technical Devices Inc. Liquid Sound Speed Chart. 2024. Available online: https://tdi-pm.com/images/tdi/PDF/Resources-Brochures/Liquid-Sound-Speed-Charts.xls (accessed on 29 January 2024).
  37. Aziz, R.A.; Bowman, D.H.; Lim, C.C. An Examination of the Relationship between Sound Velocity and Density in Liquids. Can. J. Phys. 1972, 50, 646–654. [Google Scholar] [CrossRef]
  38. Kerl, K.; Varchmin, H. Refractive index dispersion (RID) of some liquids in the UV/VIS between 20 °C and 60 °C. J. Mol. Struct. 1995, 349, 257–260. [Google Scholar] [CrossRef]
  39. Moutzouris, K.; Papamichael, M.; Betsis, S.C.; Stavrakas, I.; Hloupis, G.; Triantis, D. Refractive, dispersive and thermo-optic properties of twelve organic solvents in the visible and near-infrared. Appl. Phys. B 2013, 116, 617–622. [Google Scholar] [CrossRef]
  40. Grace, E.; Butcher, A.; Monroe, J.; Nikkel, J.A. Index of refraction, Rayleigh scattering length, and Sellmeier coefficients in solid and liquid argon and xenon. Nucl. Instruments Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2017, 867, 204–208. [Google Scholar] [CrossRef]
  41. Donadio, D.; Bernasconi, M.; Tassone, F. Photoelasticity of sodium silicate glass from first principles. Phys. Rev. B 2004, 70, 214205. [Google Scholar] [CrossRef]
  42. Liang, X.; Ismail-Beigi, S. Degree of locality of elasto-optic response in solids. Phys. Rev. B 2019, 100, 245204. [Google Scholar] [CrossRef]
  43. Imai, T.; Kawamura, S.; Oka, S. Space charge assisted evaluation method of the elasto-optic coefficients of electrooptic crystals. Opt. Mater. Express 2020, 10, 2181. [Google Scholar] [CrossRef]
  44. Laskar, J.M.; Shravan Kumar, P.; Herminghaus, S.; Daniels, K.E.; Schröter, M. High refractive index immersion liquid for superresolution 3D imaging using sapphire-based aplanatic numerical aperture increasing lens optics. Appl. Opt. 2016, 55, 3165. [Google Scholar] [CrossRef]
  45. Renuka Kumari, S.; Venkateswarlu, P.; Prabhakar, G. Excess volumes and speeds of sound of mixtures of 1,2-dibromoethane with chlorinated ethanes and ethenes at 303.15 K. J. Chem. Thermodyn. 2005, 37, 737–742. [Google Scholar] [CrossRef]
  46. Springer Nature. SpringerMaterials. Available online: https://materials.springer.com/ (accessed on 7 February 2024).
  47. Mehta, S.; Chauhan, R. Volume and compressibility of mixtures of γ-butyrolactam (n = 5) with nitro-compounds. Fluid Phase Equilibria 2001, 187–188, 209–220. [Google Scholar] [CrossRef]
  48. Langa, E.; Mainar, A.M.; Pardo, J.I.; Urieta, J.S. Excess enthalpy, density, and speed of sound for the mixtures β-pinene + 2-methyl-1-propanol or 2-methyl-2-propanol at several semperatures. J. Chem. Eng. Data 2007, 52, 2182–2187. [Google Scholar] [CrossRef]
  49. SensoTech GmbH. Density Measurement in Liquids. Available online: https://www.sensotech.com/en/density-measurement-in-liquids (accessed on 7 February 2024).
  50. Oswal, S.L.; Patel, N.B. Speeds of sound, isentropic compressibilities, and excess volumes of binary mixtures of acrylonitrile with organic solvents. J. Chem. Eng. Data 2000, 45, 225–230. [Google Scholar] [CrossRef]
  51. Shukla, R.; Kumar, A.; Awasthi, N.; Srivastava, U.; Srivastava, K. Speed of sound and isentropic compressibility of benzonitrile, chlorobenzene, benzyl chloride and benzyl alcohol with benzene from various models at temperature range 298.15–313.15 K. Arab. J. Chem. 2017, 10, 895–905. [Google Scholar] [CrossRef]
  52. Burakowski, A.; Gliński, J. Dimerization constants from acoustic measurements: Solutions of benzene, cyclohexylamine and aniline in cyclohexane. J. Solut. Chem. 2017, 46, 1501–1513. [Google Scholar] [CrossRef]
  53. Tabuchi, D. Dispersion and absorption of sound in ethyl formate and study of the rotational isomers. J. Chem. Phys. 1958, 28, 1014–1021. [Google Scholar] [CrossRef]
  54. Baluja, S.; Pandaya, N.; Kachhadia, N.; Solanki, A.; Inamdar, P. Thermodynamic and acoustical studies of binary mixtures of diethyl malonate at 308.15 K. Phys. Chem. Liq. 2005, 43, 309–316. [Google Scholar] [CrossRef]
  55. Nayak, J.N.; Aralaguppi, M.I.; Aminabhavi, T.M. Density, viscosity, refractive index, and speed of sound in the binary mixtures of 1,4-dioxane + ethyl acetoacetate, + diethyl oxalate, + diethyl phthalate, or + dioctyl phthalate at 298.15, 303.15, and 308.15 K. J. Chem. Eng. Data 2003, 48, 1489–1494. [Google Scholar] [CrossRef]
  56. Dubey, G.P.; Kumar, K. Densities, viscosities, and speeds of sound of binary liquid mixtures of ethylenediamine with alcohols at T = (293.15 to 313.15) K. J. Chem. Eng. Data 2011, 56, 2995–3003. [Google Scholar] [CrossRef]
  57. Rathnam, M.V.; Mankumare, S.; Kumar, M.S.S. Density, viscosity, and speed of sound of (methyl benzoate + cyclohexane), (methyl benzoate + n-hexane), (methyl benzoate + heptane), and (methyl benzoate + octane) at temperatures of (303.15, 308.15, and 313.15) K. J. Chem. Eng. Data 2009, 55, 1354–1358. [Google Scholar] [CrossRef]
  58. Ott, L.S.; Huber, M.L.; Bruno, T.J. Density and speed of sound measurements on five fatty acid methyl esters at 83 kPa and temperatures from (278.15 to 338.15) K. J. Chem. Eng. Data 2008, 53, 2412–2416. [Google Scholar] [CrossRef]
  59. Roy, R.; Mondal, S.; Ghosh, S.; Jengathe, S. Acoustic data on molecular interactions in mixtures of nitromethane and nitroethane in acetone at 303–318 K. Russ. J. Phys. Chem. A 2018, 92, 2606–2611. [Google Scholar] [CrossRef]
  60. Rostocki, A.J.; Siegoczyński, R.M.; Kiełczyński, P.; Szalewski, M.; Balcerzak, A.; Zduniak, M. Employment of a novel ultrasonic method to investigate high pressure phase transitions in oleic acid. High Press. Res. 2011, 31, 334–338. [Google Scholar] [CrossRef]
  61. Groot, M.S.d. Ultrasonic Relaxation Due to Rotational Isomerism. Ph.D. Thesis, Universiteit Leiden, Leiden, The Netherlands, 1958. Available online: https://www.lorentz.leidenuniv.nl/history/proefschriften/sources/DeGroot_1958.pdf (accessed on 5 June 2024).
  62. Bahadur, I.; Deenadayalu, N.; Naidoo, P.; Ramjugernath, D. Density, speed of sound, and refractive index measurements for the binary systems (butanoic acid+propanoic acid, or 2-methyl-propanoic acid) at T = (293.15 to 313.15) K. J. Chem. Thermodyn. 2013, 57, 203–211. [Google Scholar] [CrossRef]
  63. Luning Prak, D.J.; Lee, B.G. Density, viscosity, speed of sound, bulk modulus, surface tension, and flash point of binary mixtures of 1,2,3,4-tetrahydronaphthalene and trans-decahydronaphthalene. J. Chem. Eng. Data 2016, 61, 2371–2379. [Google Scholar] [CrossRef]
Figure 1. Dependence of the elasto-optic constant of liquids on the refractive index: experimental data and modeling results.
Figure 1. Dependence of the elasto-optic constant of liquids on the refractive index: experimental data and modeling results.
Materials 17 02810 g001
Figure 2. Data on the density and speed of sound for a number of organic liquids: 1—saturated hydrocarbons; 2—replacement of one hydrogen with F or Cl; 3—from two to six hydrogens are replaced with F or Cl; 4—from one to four hydrogens are replaced with Br; 5—aromatic hydrocarbons.
Figure 2. Data on the density and speed of sound for a number of organic liquids: 1—saturated hydrocarbons; 2—replacement of one hydrogen with F or Cl; 3—from two to six hydrogens are replaced with F or Cl; 4—from one to four hydrogens are replaced with Br; 5—aromatic hydrocarbons.
Materials 17 02810 g002
Figure 3. Dependence of the normalized acousto-optic figure of merit of liquid on the refractive index according to the Lorentz–Lorenz model: (a) w > 0 , v > 0 ; (b) w < 0 , v < 0 .
Figure 3. Dependence of the normalized acousto-optic figure of merit of liquid on the refractive index according to the Lorentz–Lorenz model: (a) w > 0 , v > 0 ; (b) w < 0 , v < 0 .
Materials 17 02810 g003
Figure 4. Dispersion curves for some liquids: (a) experimental data for refractive index; (b) calculated elasto-optic constant.
Figure 4. Dispersion curves for some liquids: (a) experimental data for refractive index; (b) calculated elasto-optic constant.
Materials 17 02810 g004
Table 1. Optimal refractive index and maximum elasto-optic constant for different models.
Table 1. Optimal refractive index and maximum elasto-optic constant for different models.
LorentzRocardLooyengaOnsagerKirkwoodNiedrichGladstoneEykmanWahidMeeten
n opt 2.0001.4141.5401.4211.5471.6251.5001.5811.4921.696
max(p)0.3750.2500.3160.2780.3200.3260.2960.3160.3030.326
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nikitin, P.A.; Pozhar, V.E. Evaluation of the Acousto-Optic Figure of Merit and the Maximum Value of the Elasto-Optic Constant of Liquids. Materials 2024, 17, 2810. https://doi.org/10.3390/ma17122810

AMA Style

Nikitin PA, Pozhar VE. Evaluation of the Acousto-Optic Figure of Merit and the Maximum Value of the Elasto-Optic Constant of Liquids. Materials. 2024; 17(12):2810. https://doi.org/10.3390/ma17122810

Chicago/Turabian Style

Nikitin, Pavel A., and Vitold E. Pozhar. 2024. "Evaluation of the Acousto-Optic Figure of Merit and the Maximum Value of the Elasto-Optic Constant of Liquids" Materials 17, no. 12: 2810. https://doi.org/10.3390/ma17122810

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop