Next Article in Journal
Effect of Ionic and Non-Ionic Surfactant on Bovine Serum Albumin Encapsulation and Biological Properties of Emulsion-Electrospun Fibers
Previous Article in Journal
Evaluation of TILI-2 as an Anti-Tyrosinase, Anti-Oxidative Agent and Its Role in Preventing Melanogenesis Using a Proteomics Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Compartmentalized Scale (PN) for Measuring Polarity Applied to Novel Ether-Functionalized Amino Acid Ionic Liquids

College of Chemistry, Liaoning University, Shenyang 110036, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(10), 3231; https://doi.org/10.3390/molecules27103231
Submission received: 23 April 2022 / Revised: 12 May 2022 / Accepted: 17 May 2022 / Published: 18 May 2022
(This article belongs to the Section Physical Chemistry)

Abstract

:
Functionalized and environmentally friendly ionic liquids are required in many fields, but convenient methods for measuring their polarity are lacking. Two novel ether-functionalized amino acid ionic liquids, 1-(2-methoxyethyl)-3-methylimidazolium alanine ([C1OC2mim][Ala]) and 1-(2-ethoxyethyl)-3-methylimidazolium alanine ([C2OC2mim][Ala]), were synthesized by a neutralization method and their structures confirmed by NMR spectroscopy. Density, surface tension, and refractive index were determined using the standard addition method. The strength of intermolecular interactions within these ionic liquids was examined in terms of standard entropy, lattice energy, and association enthalpy. A new polarity scale, PN, is now proposed, which divides polarity into two compartments: the surface and the body of the liquid. Surface tension is predicted via an improved Lorentz-Lorenz equation, and molar surface entropy is used to determine the polarity of the surface. This new PN scale is based on easily measured physicochemical parameters, is validated against alternative polarity scales, and is applicable to both ionic and molecular liquids.

1. Introduction

The green chemistry concept is a widely accepted focus of modern chemical research, including the design of new materials. Ionic liquids (ILs) have emerged as useful green reaction media due to their many unique features, such as low vapor pressure and high thermal stability [1,2]. They play important roles in many fields, including energy storage [3], catalysis [4], pharmaceuticals, and medicine [5,6]. However, the relatively high viscosity of ILs is a barrier to further practical applications. Inserting ether groups into the cations of ILs has been shown to substantially reduce their viscosity without lowering thermal stability, while also reducing their toxicity [7,8,9,10]. Ether-functionalized ILs (EFILs) have demonstrated remarkable performance in many fields. For example, they dissolve lignocellulosic biomasses [7], reduce viscosity and provide coordination sites for lithium ions in Li/Li-ion batteries [8], and enhance CO2 selectivity during CO2 capture [11]. However, traditional ILs containing anions such as Cl, [BF4], and [PF6] are environmentally hazardous. The development of environmentally friendly, task-specific ILs based on renewable bioresources (such as amino acids and fatty acids) is an environmental necessity. Ohno and Fukumoto were the first to prepare amino acid ILs (AAILs), using 20 different amino acids as the anion; these AAILs demonstrated lower toxicity [12,13]. AAILs have now been utilized in enantioselective separation [14], extraction separation [15], and CO2 capture [11]. ILs with imidazolium-based cations exhibit low toxicity. Meanwhile, the shorter the alkyl chain on the imidazole ring, the lower the toxicity [16,17]. Thus, imidazolium ILs have proved more attractive than ILs based on ammonium, phosphonium, and pyridinium cations. The present study describes the preparation of two novel ether-functionalized, imidazolium-based AAILs: 1-(2-methoxyethyl)-3-methylimidazolium alanine ([C1OC2mim][Ala]) and 1-(2-ethoxyethyl)-face Tension, and Refractive Inde ([C2OC2mim][Ala]), together abbreviated as [CnOC2mim][Ala](n = 1, 2).
There is scope for considerable further research into EFILs because their physiochemical properties are highly susceptible to changes in structure. More experimental data are required to elucidate their structure–property relationships. This study measures the density, surface tension, and refractive index of [CnOC2mim][Ala](n = 1, 2). Density correlates with packing efficiency and intermolecular interactions and is a critical design property in chemical engineering [18]. Its study provides insight into the microstructure and macroscopic properties of ILs [19]. Surface tension is a crucial property at liquid–gas interfaces [20], affecting how the phases interact [21]. The surface tension of ILs is between those of alkanes and water [22]. It can be measured directly or predicted using, for example, the parachor formula [23] or group contribution methods [24]. This study predicts surface tension using an improved Lorentz-Lorenz equation. ILs are often used as solvents, so determining their polarity is crucial. Due to their non-structured nature, polarity cannot be determined by traditional methods such as relative permittivity (εr) and dipole moment (δ) [25]. The most widely used experimental method for IL polarity is the ET(30) scale, which measures the solvatochromic UV−Vis absorbance shift of a solute. However, this method is time-consuming and expensive, so attempts have been made to develop predictive models [26,27].
The present study proposes a new polarity scale, PN, which enables polarity to be predicted from the easily measured physicochemical properties of density, surface tension, and refractive index. Following on from our previous studies [28,29,30], (i) [CnOC2mim][Ala](n = 1, 2) are synthesized and their structures confirmed by nuclear magnetic resonance spectroscopy (NMR); (ii) their density, surface tension, and refractive index are measured from 288.15 to 328.15 K at 5 K intervals; (iii) the strength of their molecular interactions are studied based on standard entropy, lattice energy, and association enthalpy; (iv) an improved Lorentz-Lorenz equation is used to predict the surface tension of ILs and molecular liquids; and (v) a new scale, PN, for estimating polarity is proposed, combining molar surface entropy s (which measures the polarity of the surface of a liquid) and the polarity coefficient P2 (which measures the polarity of the body of a liquid).

2. Results and Discussion

2.1. Density, Surface Tension, and Refractive Index of [CnOC2mim][Ala](n = 1, 2)

The density (ρ), surface tension (γ), and refractive index (nD) of [CnOC2mim][Ala](n = 1, 2) with various water contents over 288.15–328.15 K (at 5 K intervals) are shown in Tables S1–S3 Supplementary Materials, with each value being an average of three measurements using the standard addition method. These parameters were plotted against water content (Figure 1), producing a series of straight lines with correlation coefficient squares (r2) consistently greater than 0.99. The y-axis intercepts of these lines give the experimental value of each parameter in anhydrous [CnOC2mim][Ala](n = 1, 2) (Table 1).

2.2. Strength of [CnOC2mim][Ala](n = 1, 2) Intermolecular Interactions

The density of ILs increases gradually as temperature rises. At higher temperatures, the mobility of constituent ions improves and the unit volume increases [31]. The thermal expansion coefficient (α) is defined as
α = ( 1 / V ) ( V / Y ) p = (   l n ρ / T ) p
where V is molar volume. Molar volume is defined as
V = M / ρ
The molecular volume Vm is defined as
V m = M / N ρ = V / N
where N is the Avogadro constant, V is molar volume, and M is molar mass.
At 298.15 K, Vm is 0.3300 nm3 for [C1OC2mim][Ala] and 0.3571 nm3 for [C2OC2mim][Ala] (Table S4). The difference between these indicates that the contribution of methylene (-CH2-) to Vm is 0.0271 nm3. This is close to the average contribution methylene makes to the Vm of several other ILs listed in Table S5 (0.0278 nm3).
Lattice energy (UPOT) and standard entropy (Sθ298) can be calculated according to Glasser’s theory; Equation (4) is suitable for MX(1:1) type ionic salts. The constants in Equation (5) are empirical values [32].
U POT = 1981.2   ( ρ / M ) 1 / 3 + 103.8
S θ 298 = 1246.5 V m + 29.5
UPOT reflects the strength of intermolecular interactions and can be used to measure the stability of ILs [32,33]. UPOT is 443 kJ·mol−1 for [C1OC2mim][Ala] and 435 kJ·mol−1 for [C2OC2mim][Ala], implying that methylene’s average contribution to UPOT is −8 kJ·mol−1. UPOT is inversely related to molar volume. Addition of a methylene group will reduce the ionic or molecular packing efficiency, decreasing the strength of the interactions.
To some extent, standard entropy reflects the degree of disorder of molecular arrangements [33]. Sθ298 is 441 J·K−1·mol−1 for [C1OC2mim][Ala] and 475 J·K−1·mol−1 for [C2OC2mim][Ala], suggesting that methylene’s average contribution to Sθ298 is 34 J·K−1·mol−1 (Table S5). This indicates that ILs with longer aliphatic chains are more disordered. Standard entropy increases as molecular volume increases. Sθ298 of ILs is usually greater than 200 J·K−1·mol−1, compared with the more molecularly ordered conventional inorganic salts such as NaCl (72.1 J·K−1·mol−1) and KCl (82.6 J·K−1·mol−1) [34]. This may explain why ILs are molten below 373 K.
The association enthalpy (ΔAHm0) also reflects the strength of intermolecular interactions: the higher the ΔAHm0, the stronger the gaseous state interactions. ΔAHm0 of ILs in the gaseous phase can be calculated based on the thermodynamic cycle shown in Scheme 1 [29].
Vaporization enthalpy (ΔlgHm0) is a key parameter for calculating ΔAHm0. It is estimated from Equation (6):
g s = a + b   ( Δ l g H m 0     R T )
g s = γ V 2 / 3 N 1 / 3
where gs is the molar surface Gibbs energy; γ is surface tension; V is molar volume; N is the Avogadro constant; ΔlgHm0 is vaporization enthalpy; R is the gas constant; T is temperature; and a and b are the empirical constants −1.519 kJ·mol−1 and 0.09991, respectively [29].
The estimated vaporization enthalpy is 164.1 kJ·mol−1 for [C1OC2mim][Ala] and 165.1 kJ·mol−1 for [C2OC2mim][Ala]. Consequently, ΔAHm0 is −278.9 kJ·mol−1 for [C1OC2mim][Ala] and −269.9 kJ·mol−1 for [C2OC2mim][Ala]. The ΔAHm0 of other ILs are listed in Table S5. Again, addition of methylene reduces packing efficiency and increases the degree of molecular disorder. Thus, for ILs with the same anions, the absolute value of ΔAHm0 decreases as the length of the imidazole ring alkyl side chains increases. For ILs with the same cations, ΔAHm0 decreases with increasing anion volume.

2.3. Prediction of Surface Tension Based on Molar Surface Gibbs Energy

The parameter gs used to estimate vaporization enthalpy was developed in our previous work by modifying Li’s model [35]. The definition of gs is consistent with the concept presented by Myers [36]. Thus, gs is a true thermodynamic function that integrates volumetric and surface properties.
Plotting gs against T for [CnOC2mim][Ala](n = 1, 2) yields straight lines (Figure 2) such that their relationship can be expressed as
g s = G 0     G 1 T
G0 and G1 for [CnOC2mim][Ala](n = 1, 2) are obtained from Figure 2 and substituted into Equation (8) to give the estimated molar surface Gibbs energy, gs(est). Values of gs, G0, G1, and gs(est) are listed in Table 2.
The Lorentz-Lorenz equation expresses the relationship between nD and the mean molecular polarizability (αp) [37]:
R m = [ ( n D 2     1 ) / ( n D 2 + 2 ) ] · V = ( 4 π N / 3 ) · α p
where Rm is molar refraction, αp is mean molecular polarizability, and nD is refractive index. This has been combined with gs to give an improved Lorentz-Lorenz equation [38] that can predict surface tension, γ(est):
γ ( est ) 3 / 2 = [ g s ( est ) 3 / 2 / N 1 / 2 R m ] ( n D 2 1 ) / ( n D 2 + 2 )
The Rm, αp, and γ(est) for [CnOC2mim][Ala](n = 1, 2) are listed in Table 3. Plotting estimated surface tension values against their corresponding experimental values produces a straight line (Figure 3). A similar plot for other ionic and molecular liquids also illustrates a linear relationship (Table S6, Figure 4), showing that this method is applicable for the surface tension prediction of both types of liquid.
For ILs sharing the same anions (Table 1 and Table S6), nD decreases as the length of the alkyl chain in the cations increases. Refractive index correlates with dipole moment [39,40], which increases with higher molecular packing density [41]. Thus, a larger dipole moment results in a higher refractive index. The nD is larger in [C1OC2mim][Ala] than [C2OC2mim][Ala], so the packing density of the former is higher, which is consistent with the density trend of the two ILs.
As polarizability increases, Coulomb interactions are reduced and ion mobility rises [42]. The αp of [C1OC2mim][Ala] is lower than [C2OC2mim][Ala], so its Coulomb interactions are stronger. This finding is similar to the pattern observed for UPOT and ΔAHm0.

2.4. Molar Surface Entropy: Polarity Contribution from Surface Liquid

For most liquids, surface tension declines as temperature increases, as shown by the Eötvös equation:
γ V 2 / 3 = k ( T c     T )
where Tc is critical temperature and k is the Eötvös equation parameter, which is associated with polarity. For some organic liquids with weak polarity, k is nearly 2.2 × 10−7 J·mol−2/3·K−1, while for some with strong polarity, such as molten NaCl, it is nearly 0.4 × 10−7 J·mol−2/3·K−1 [35,43]. However, the physical significance of k is not clear. Multiplying both sides of the Eötvös equation by N1/3 [35] gives
g s = C 0     C 1 T
which fits the fundamental thermodynamic concept:
G = H T S
C1 denotes the molar surface entropy and is given by C1= −( g s T )p.
The relationship between molar surface entropy (defined here as s [29]) and k is expressed as
s = N 1 / 3 k
Entropy is directly linked to the number of microstates [44]. Higher entropy means molecules can be arranged in more ways, while the total energy remains constant. Thus, the physical significance of s is clear—it reflects the polarity of a liquid’s surface (higher s, lower surface polarity). The value of s is 17.39 J·mol−1·K−1 for [C1OC2mim][Ala] and 19.94 J·mol−1·K−1 for [C2OC2mim][Ala]. Values for other EFILs are listed in Table 4. The overall trend is that for ILs with the same cations, such as [C1OC2mim]+, [C2OC2mim]+, and [C1OC4mim]+, s increases as the volume of anions increases. Cl, [Ala], [Thr], and [Gly] clearly obey this rule. It can be speculated that larger anions cause greater disordering of surface molecules, leading to lower polarity. However, [NTf2] does not conform to this rule. This may be due to it being more symmetrical than other anions, facilitating a more orderly arrangement of surface molecules and increasing the polarity. For ILs sharing the same anions, the general trend is that s increases as the volume of cations increases. Values of s for different other ILs (Table S7) confirm this effect.
Data for [C1OC2mim][NTf2], [C2OC1mim][NTf2], [C2OC2mim][NTf2], and [C1OC3mim][NTf2] (Table 4) imply that, for ILs with the same number of alkyl side chain carbons, the position of the ether group also affects s, presumably because it affects packing efficiency on the liquid surface.

2.5. A New Model for Predicting Polarity

Experimental methods for measuring polarity, such as ET(30), are time consuming and laborious. Here, we present a predictive model that establishes a relationship between polarity and the easily determined physicochemical properties of density, surface tension, and refractive index.
Our previous work [28], based on Hildebrand and Scott’s theory [47], proposed δμ as a polarity scale. δμ is the solubility parameter derived from the contribution of the average permanent dipole moment:
δ μ 2 = Δ g l H 0 m μ / V     ( 1     x ) RT / V
where V is the molar volume and ΔglH0 is the contribution of the average permanent dipole moment to ΔglH0m, such that
Δ g l H 0 m μ = Δ g l H 0 m     Δ g l H 0 mn
where ΔglH0mn is the contribution of the induced dipole moment to ΔglH0m and can be calculated from the Lawson–Ingham equation [48]:
Δ g l H 0 mn = C   [ ( n D 2     1 )   /   ( n D 2 + 2 ) ] V = C   R m
where C is the empirical constant 1.297 kJ·cm−3. In Equation (15), x represents ΔglH0mnglH0m (at 298.15 K). The δμ polarity of [C1OC2mim][Ala] (21.03 J1/2·cm−3/2) is larger than [C2OC2mim][Ala] (19.65 J1/2·cm−3/2).
However, there is an obvious drawback to δμ: it has a dimension (J1/2·cm−3/2), while some polarity scales, such as the dielectric constant, are non-dimensional. Furthermore, the contribution of the induced dipole moment was neglected. Therefore, δμ was improved as follows and designated as P [45]:
P = δ μ / δ n = [ ( Δ g l H 0 m μ / V ( 1 x )   R T / V ) / ( Δ g l H 0 mn / V x R T / V ) ] 1 / 2
where δn is the solubility parameter from the contribution of the induced dipole moment. Comparison of ΔglH0 and ΔglH0mn shows that (1−x)RT/V and xRT/V can be omitted and Equation (18) can be expressed as
P = ( Δ g l H 0 m μ / Δ g l H 0 mn ) 1 / 2
The effects of average permanent dipole moment and induced dipole moment are both considered within P, which is dimensionless, with a large value indicating high polarity. [C4mim][BF4] is hydrophilic and [C4mim][NTf2] is hydrophobic. According to Seddon et al. [49], P is 1.226 for [C4mim][BF4] and 0.401 for [C4mim][NTf2]. This higher polarity of [C4mim][BF4] fits practical experience. Thus, the parameter P is capable of measuring the polarity of ILs. P is 1.191 for [C1OC2mim][Ala] and 1.043 for [C2OC2mim][Ala].
This study divided polarity into two compartments: the contribution from the body of a liquid, and the contribution from the surface. Cohesive energy density can demonstrate the strength of intermolecular interactions within the body [48]. δμ2 is the cohesive energy density from the average permanent dipole moment and δn2 is from the induced dipole moment. Consequently, P2 can describe the polarity of the body of a liquid:
P 2 = δ μ 2 / δ n 2
Molar surface entropy (s) was proven above to reflect the polarity of a liquid surface. Combining s and P2, a new polarity scale, PN, is now proposed:
P N = s / P 2 = s / ( δ μ 2 / δ n 2 )
This compartmentalized scale is a novel method to evaluate polarity, with a large PN indicating weak polarity. Based on literature data [50,51,52,53,54,55], PN is 22.03 J·mol−1·K−1 for [C4mim][NTf2] and 14.50 J·mol−1·K−1 for [C4mim][BF4]. These results fit with practical experience and demonstrate the rationality of the PN scale. The PN of [C1OC2mim][Ala] is 12.26 J·mol−1·K−1 and is 18.33 J·mol−1·K−1 for [C2OC2mim][Ala], which is the same polarity trend observed using δμ and P. As discussed above, ILs with longer alkyl chains exhibit higher standard entropy and lower molecular packing efficiency. The observed higher polarity of [C1OC2mim][Ala] may be due to its stronger intermolecular interactions and more ordered molecular arrangement. This fact can be explained as follows: a polar molecule has a permanent electric dipole moment, and a molecule may be polar if it has low symmetry [44]. ILs have asymmetric structures, and an orderly arrangement of ILs will maintain this structure. In this situation, their permanent electric dipole moments will not counteract each other, and polarity will be enhanced. The PN of various ether-functionalized ILs are listed in Table 5. For ILs with the same anion, PN declines as the length of the imidazole ring alkyl side chain increases. This trend supports the contention that the strength of intermolecular interactions and the degree of disorder of molecular arrangements influence polarity.
The PN scale can be validated by comparison with other polarity scales (Table 6). FTIR spectroscopy probes [56] and ETN [57] show that the polarity of [C2mim]BF4 is larger than [C4mim]BF4. The PN scale gives the same qualitative result. Wu et al. determined the order of polarity of several ILs to be [C4mim]BF4 > [C4mim]NTf2 > [C4mim]OAc using ETN [58]. Again, PN gives the same result.
Furthermore, when PN is applied to molecular liquids, the estimated polarity is broadly, and inversely, consistent with the dielectric constant ε [59] (Table 7). The correlation between PN and the inverse ε−1 (Figure 5) is r2 = 0.94, demonstrating that PN is also suitable for molecular liquids.
ET(30) is one of the most popular polarity scales for evaluating ILs, but it is laborious and costly [60]. Moreover, the results vary depending on the molecular probe used [61]. However, using PN, polarity can be determined simply from density, surface tension, and refractive index. The dielectric constant is the traditional polarity scale for organic solvents. The comparison of dielectric constant and PN proves the universal applicability of PN. Thus, this new PN scale is demonstrated to be a viable predictive method for evaluating the polarity of both ionic and molecular liquids based on easily measured physicochemical properties. It will find applications in many fields, particular those employing novel ionic liquids for which the evaluation of polarity is expensive and time consuming.

3. Materials and Methods

3.1. Materials

The sources and purity of reagents are listed in Table 8. N-Methylimidazole (AR grade) was purified by distillation, while 2-chloroethyl methyl ether and 2-chloroethyl ethyl ether (both AR grade) were used as purchased. DL-Alanine was recrystallized from water and dried in a vacuum oven [62].

3.2. Preparation of ILs [CnOC2mim][Ala](n = 1, 2)

[CnOC2mim][Ala](n = 1, 2) were prepared by Fukumoto’s neutralization method [12], and [CnOC2mim]Cl(n = 1, 2) by Sheldon’s method [63]. An equal molar amount of 2-chloroethyl methyl ether or 2-chloroethyl ethyl ether was added dropwise to N-methylimidazole under nitrogen in a three-necked round-bottom flask while stirring at 298.15 K. The reaction temperature reached 353.15 K with 2-chloroethyl methyl ether and 373.15 K with 2-chloroethyl ethyl ether. The reactions lasted for 48 h and produced light yellow liquids, which were then washed three times with ethyl acetate, producing [CnOC2mim]Cl(n = 1, 2). These were transformed into [CnOC2mim]OH(n = 1, 2) using basic anion exchange resin conditioned with sodium hydroxide [29]. The aqueous [CnOC2mim]OH(n = 1, 2) were then added dropwise (at slight excess) to aqueous DL-alanine and reacted for 72 h, yielding [CnOC2mim][Ala](n = 1, 2). Water was removed by rotary evaporation and excess DL-alanine by ethanol:acetonitrile (9:1). The solvents were evaporated under reduced pressure and [CnOC2mim][Ala](n = 1, 2) were dried in a vacuum oven for 60 h at 353.15 K. The chemical structures of [CnOC2mim][Ala](n = 1, 2) are shown in Figure 6.

3.3. Analytical Methods

The structures of [CnOC2mim][Ala](n = 1, 2) were characterized by NMR (Varian XL-300), as shown in the Supplementary Materials. The final water contents (w2) of [CnOC2mim][Ala](n = 1, 2), as measured using a ZSD-2 Karl Fischer moisture titrator, were 0.00472 and 0.00640 ± 0.0001 (mass fraction), respectively.
Since [CnOC2mim][Ala](n = 1, 2) form strong hydrogen bonds with water, it is difficult to remove all traces of water from these ILs, which affects their density, surface tension, and refractive index. Therefore, the standard addition method was used to determine these properties. Each parameter was measured in [CnOC2mim][Ala](n = 1, 2) at different water contents following heating at graduated temperatures. Values were then plotted against water content, and the intercept of the regression lines yielded the parameter values in the anhydrous ILs at a given temperature.

4. Conclusions

[CnOC2mim][Ala](n = 1, 2) were prepared and their structures confirmed by NMR. Density, surface tension, and refractive index were determined by the standard addition method. Adding methylene to the aliphatic chain of an IL increased its standard entropy. Lattice energy and association enthalpy measurements showed that molecules of [C1OC2mim][Ala] were more compacted, and their intermolecular interactions stronger, than [C2OC2mim][Ala]. An improved Lorentz-Lorenz equation predicted the surface tension of both ionic and molecular liquids. A new compartmentalized polarity scale (PN) based on molar surface Gibbs energy and dipole moments is presented. It encompasses the polarity of both the surface and body of a liquid. [C1OC2mim][Ala] is shown to have higher polarity than [C2OC2mim][Ala] based on PN. The validity of PN is demonstrated by comparison with alternative polarity scales and published polarities of both ionic and molecular liquids.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27103231/s1, Figure S1. 1H NMR spectrum of IL [COC2mim][Ala]; Figure S2. 1H NMR spectrum of IL [C2OC2mim][Ala]; Figure S3. 13C NMR spectrum of IL [COC2mim][Ala]; Figure S4. 13C NMR spectrum of IL [C2OC2mim][Ala]; Table S1. Densities of Ionic Liquids Containing Various Amounts of Water at pressure p = 0.1 MPa; Table S2. Surface Tensions of Ionic Liquids Containing Various Amounts of Water at pressure p = 0.1 MPa; Table S3. Refractive Indexes of Ionic Liquids Containing Various Amount of Water at pressure p = 0.1 MPa; Table S4. The values of molar volume, V/cm3·mol−1 and molecular volume, Vm/nm3 for the ILs [CnOC2mim][Ala](n = 1, 2); Table S5. Molecular volume, Vm, standard molar entropy, S0, lattice energy, UPOT, vaporization enthalpy, ΔlgHm0, association enthalpy, ΔAHm0 for some ionic liquids; Table S6. The estimated surface tension, γ(est), experimental surface tension, γ(exp), refractive index, nD, experimental density ρ(exp), molar surface Gibbs energy, gs, molar refraction, Rm, molar volume, V for different ILs and molecular liquids; Table S7. The molar surface entropy, s, for some ILs; Density, ρ, surface tension, and refractive index, nD measuring methods; citation of ref. [30,39,43,52,54,55,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106].

Author Contributions

Writing—original draft preparation, X.Z.; data curation, C.G.; visualization, W.W.; supervision, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 21773100; the Education Bureau of Liaoning Province, grant number LJC202005; and the Liaoning BaiQianWan Talents Program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of compounds are available from the authors.

References

  1. Blanchard, L.A.; Hancu, D.; Beckman, E.J.; Brennecke, J.F. Green processing using ionic liquids and CO2. Nature 1999, 399, 28–29. [Google Scholar] [CrossRef]
  2. Wei, G.-T.; Yang, Z.; Lee, C.-Y.; Yang, H.-Y.; Wang, C.C. Aqueous− organic phase transfer of gold nanoparticles and gold nanorods using an ionic liquid. J. Am. Chem. Soc. 2004, 126, 5036–5037. [Google Scholar] [CrossRef] [PubMed]
  3. Kim, G.-T.; Jeong, S.; Xue, M.-Z.; Balducci, A.; Winter, M.; Passerini, S.; Alessandrini, F.; Appetecchi, G. Development of ionic liquid-based lithium battery prototypes. J. Power Sources 2012, 199, 239–246. [Google Scholar] [CrossRef]
  4. Liu, Y.; Yao, X.; Yao, H.; Zhou, Q.; Xin, J.; Lu, X.; Zhang, S. Degradation of poly(ethylene terephthalate) catalyzed by metal-free choline-based ionic liquids. Green Chem. 2020, 22, 3122–3131. [Google Scholar] [CrossRef]
  5. Williams, H.D.; Sahbaz, Y.; Ford, L.; Nguyen, T.-H.; Scammells, P.J.; Porter, C.J. Ionic liquids provide unique opportunities for oral drug delivery: Structure optimization and in vivo evidence of utility. Chem. Commun. 2014, 50, 1688–1690. [Google Scholar] [CrossRef]
  6. Shamshina, J.L.; Kelley, S.P.; Gurau, G.; Rogers, R.D. Chemistry: Develop ionic liquid drugs. Nat. News 2015, 528, 188. [Google Scholar] [CrossRef] [Green Version]
  7. Rahim, A.H.A.; Yunus, N.M.; Hamzah, W.S.W.; Sarwono, A.; Muhammad, N. Low-Viscosity Ether-Functionalized Ionic Liquids as Solvents for the Enhancement of Lignocellulosic Biomass Dissolution. Processes 2021, 9, 261. [Google Scholar] [CrossRef]
  8. Zhang, S.; Li, J.; Jiang, N.; Li, X.; Pasupath, S.; Fang, Y.; Liu, Q.; Dang, D. Rational Design of an Ionic Liquid-Based Electrolyte with High Ionic Conductivity Towards Safe Lithium/Lithium-Ion Batteries. Chem.—Asian J. 2019, 14, 2810–2814. [Google Scholar] [CrossRef]
  9. Zeng, S.; Wang, J.; Bai, L.; Wang, B.; Gao, H.; Shang, D.; Zhang, X.; Zhang, S. Highly selective capture of CO2 by ether-functionalized pyridinium ionic liquids with low viscosity. Energy Fuels 2015, 29, 6039–6048. [Google Scholar] [CrossRef]
  10. Trush, M.; Metelytsia, L.; Semenyuta, I.; Kalashnikova, L.; Papeykin, O.; Venger, I.; Tarasyuk, O.; Bodachivska, L.; Blagodatnyi, V.; Rogalsky, S. Reduced ecotoxicity and improved biodegradability of cationic biocides based on ester-functionalized pyridinium ionic liquids. Environ. Sci. Pollut. Res. 2019, 26, 4878–4889. [Google Scholar] [CrossRef]
  11. Qu, Y.; Lan, J.; Chen, Y.; Sun, J. Amino acid ionic liquids as efficient catalysts for CO2 capture and chemical conversion with epoxides under metal/halogen/cocatalyst/solvent-free conditions. Sustain. Energy Fuels 2021, 5, 2494–2503. [Google Scholar] [CrossRef]
  12. Fukumoto, K.; Yoshizawa, M.; Ohno, H. Room temperature ionic liquids from 20 natural amino acids. J. Am. Chem. Soc. 2005, 127, 2398–2399. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Z.; Sun, C.; Zhang, C.; Zhao, S.; Cai, M.; Liu, Z.; Yu, Q. Amino acid ionic liquids as anticorrosive and lubricating additives for water and their environmental impact. Tribol. Int. 2021, 153, 106663. [Google Scholar] [CrossRef]
  14. Jiang, J.; Mu, X.; Qiao, J.; Su, Y.; Qi, L. New chiral ligand exchange capillary electrophoresis system with chiral amino amide ionic liquids as ligands. Talanta 2017, 175, 451–456. [Google Scholar] [CrossRef] [PubMed]
  15. Tang, F.; Zhang, Q.; Ren, D.; Nie, Z.; Liu, Q.; Yao, S. Functional amino acid ionic liquids as solvent and selector in chiral extraction. J. Chromatogr. A 2010, 1217, 4669–4674. [Google Scholar] [CrossRef] [PubMed]
  16. Ghavre, M.; Byrne, O.; Altes, L.; Surolia, P.K.; Spulak, M.; Quilty, B.; Thampi, K.R.; Gathergood, N. Low toxicity functionalised imidazolium salts for task specific ionic liquid electrolytes in dye-sensitised solar cells: A step towards less hazardous energy production. Green Chem. 2014, 16, 2252–2265. [Google Scholar] [CrossRef]
  17. Kebaili, H.; Pérez de los Ríos, A.; Salar-García, M.J.; Ortiz-Martínez, V.M.; Kameche, M.; Hernández-Fernández, J.; Hernández-Fernández, F.J. Evaluating the toxicity of ionic liquids on Shewanella sp. for designing sustainable bioprocesses. Front. Mater. 2020, 7, 387. [Google Scholar] [CrossRef]
  18. Jouyban, A.; Mirheydari, S.N.; Barzegar-Jalali, M.; Shekaari, H.; Acree, W.E. Comprehensive models for density prediction of ionic liquid+ molecular solvent mixtures at different temperatures. Phys. Chem. Liq. 2020, 58, 309–324. [Google Scholar] [CrossRef]
  19. Chen, Z.; Huo, Y.; Cao, J.; Xu, L.; Zhang, S. Physicochemical Properties of Ether-Functionalized Ionic Liquids: Understanding Their Irregular Variations with the Ether Chain Length. Ind. Eng. Chem. Res. 2016, 55, 11589–11596. [Google Scholar] [CrossRef]
  20. Pham, T.P.T.; Cho, C.-W.; Yun, Y.-S. Environmental fate and toxicity of ionic liquids: A review. Water Res. 2010, 44, 352–372. [Google Scholar] [CrossRef]
  21. Atashrouz, S.; Amini, E.; Pazuki, G. Modeling of surface tension for ionic liquids using group method of data handling. Ionics 2015, 21, 1595–1603. [Google Scholar] [CrossRef]
  22. Holbrey, J.D.; Rogers, R.D.; Mantz, R.A.; Trulove, P.C.; Cocalia, V.A.; Visser, A.E.; Anderson, J.L.; Anthony, J.L.; Brennecke, J.F.; Maginn, E.J.; et al. Physicochemical Properties. In Ionic Liquids in Synthesis; 2007; pp. 57–174. Available online: https://onlinelibrary.wiley.com/doi/book/10.1002/9783527621194 (accessed on 22 April 2022).
  23. Nemec, T. Prediction of surface tension of binary mixtures with the parachor method. In EPJ Web of Conferences; EDP Sciences: Les Ulis, France, 2015; Volume 92, p. 02054. [Google Scholar]
  24. Paduszynski, K. Extensive Databases and Group Contribution QSPRs of Ionic Liquid Properties. 3: Surface Tension. Ind. Eng. Chem. Res. 2021, 60, 5705–5720. [Google Scholar] [CrossRef]
  25. Shukla, S.K.; Kumar, A. Polarity issues in room temperature ionic liquids. Clean Technol. Environ. Policy 2015, 17, 1111–1116. [Google Scholar] [CrossRef]
  26. Cláudio, A.F.M.; Swift, L.; Hallett, J.P.; Welton, T.; Coutinho, J.A.P.; Freire, M.G. Extended scale for the hydrogen-bond basicity of ionic liquids. Phys. Chem. Chem. Phys. 2014, 16, 6593–6601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Kurnia, K.A.; Lima, F.; Cláudio, A.F.M.; Coutinho, J.A.P.; Freire, M.G. Hydrogen-bond acidity of ionic liquids: An extended scale. Phys. Chem. Chem. Phys. 2015, 17, 18980–18990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Hong, M.; Liu, R.J.; Yang, H.X.; Guan, W.; Tong, J.; Yang, J.Z. Determination of the vaporisation enthalpies and estimation of the polarity for 1-alkyl-3-methylimidazolium propionate {C(n)mim Pro (n = 2, 3)} ionic liquids. J. Chem. Thermodyn. 2014, 70, 214–218. [Google Scholar] [CrossRef]
  29. Zheng, X.; Gong, Y.; Jiang, W.; Yu, K.; Tong, J.; Yang, J. The application of molar surface Gibbs energy to predicting ILs’ properties. J. Mol. Liq. 2019, 288, 111004. [Google Scholar] [CrossRef]
  30. Tong, J.; Hong, M.; Chen, Y.; Wang, H.; Guan, W.; Yang, J.-Z. The surface tension, density and refractive index of amino acid ionic liquids: C(3)mim Gly and C(4)mim Gly. J. Chem. Thermodyn. 2012, 54, 352–357. [Google Scholar] [CrossRef]
  31. Singh, D.; Gardas, R.L. Influence of Cation Size on the Ionicity, Fluidity, and Physiochemical Properties of 1,2,4-Triazolium Based Ionic Liquids. J. Phys. Chem. B 2016, 120, 4834–4842. [Google Scholar] [CrossRef]
  32. Glasser, L. Lattice and phase transition thermodynamics of ionic liquids. Thermochim. Acta 2004, 421, 87–93. [Google Scholar] [CrossRef]
  33. Gusain, R.; Panda, S.; Bakshi, P.S.; Gardas, R.L.; Khatri, O.P. Thermophysical properties of trioctylalkylammonium bis(salicylato)borate ionic liquids: Effect of alkyl chain length. J. Mol. Liq. 2018, 269, 540–546. [Google Scholar] [CrossRef]
  34. Lambert, F.L.; Leff, H.S. The Correlation of Standard Entropy with Enthalpy Supplied from 0 to 298.15 K. J. Chem. Educ. 2009, 86, 94–98. [Google Scholar] [CrossRef]
  35. Tong, J.; Qu, Y.; Li, K.; Chen, T.-F.; Tong, J.; Yang, J.-Z. The molar surface Gibbs energy of the aqueous solution of the ionic liquid [C6mim][OAc]. J. Chem. Thermodyn. 2016, 97, 362–366. [Google Scholar] [CrossRef]
  36. Myers, R.T. True molar surface energy and alignment of surface molecules. J. Colloid Interface Sci. 2004, 274, 229–236. [Google Scholar] [CrossRef] [PubMed]
  37. Ersfeld, B.; Felderhof, B.U. Retardation correction to the Lorentz-Lorenz formula for the refractive index of a disordered system of polarizable point dipoles. Phys. Rev. E 1998, 57, 1118–1126. [Google Scholar] [CrossRef]
  38. Zhang, D.; Jiang, W.; Liu, L.; Yu, K.; Hong, M.; Tong, J. The molar surface Gibbs energy and polarity of ether-functionalized ionic liquids. J. Chem. Thermodyn. 2019, 138, 313–320. [Google Scholar] [CrossRef]
  39. Hong, M.; Sun, A.; Liu, C.; Guan, W.; Tong, J.; Yang, J.-Z. Physico-Chemical Properties of 1-Alkyl-3-methylimidazolium Propionate Ionic Liquids {[Cnmim][Pro](n = 3, 4, 5, 6)} from 288.15 K to 328.15 K. Ind. Eng. Chem. Res. 2013, 52, 15679–15683. [Google Scholar] [CrossRef]
  40. Seoane, R.G.; Corderí, S.; Gómez, E.; Calvar, N.; González, E.J.; Macedo, E.A.; Domínguez, Á. Temperature Dependence and Structural Influence on the Thermophysical Properties of Eleven Commercial Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 2492–2504. [Google Scholar] [CrossRef]
  41. Sun, Q.; Yang, W.; Cheng, Y.; Dong, J.; Zhu, M.; Zhang, B.; Dubois, A.; Zhu, M.; Jie, W.; Xu, Y. Anisotropic dielectric behavior of layered perovskite-like Cs3Bi2I9 crystals in the terahertz region. Phys. Chem. Chem. Phys. 2020, 22, 24555–24560. [Google Scholar] [CrossRef]
  42. Weiß, N.; Schmidt, C.H.; Thielemann, G.; Heid, E.; Schröder, C.; Spange, S. The physical significance of the Kamlet–Taft π* parameter of ionic liquids. Phys. Chem. Chem. Phys. 2021, 23, 1616–1626. [Google Scholar] [CrossRef]
  43. Wei, J.; Fan, B.-H.; Pan, Y.; Xing, N.-N.; Men, S.-Q.; Tong, J.; Guan, W. Vaporization enthalpy and the molar surface Gibbs free energy for ionic liquids C(n)Dmim NTF2 (n = 2, 4). J. Chem. Thermodyn. 2016, 101, 278–284. [Google Scholar] [CrossRef]
  44. Peter Atkins, P.; De Paula, J. Atkins’ Physical Chemistry; OUP Oxford: Oxford, UK, 2014. [Google Scholar]
  45. Zhang, D.; Li, B.; Hong, M.; Kong, Y.-X.; Tong, J.; Xu, W.-G. Synthesis and characterization of physicochemical properties of new ether-functionalized amino acid ionic liquids. J. Mol. Liq. 2020, 304, 112718. [Google Scholar] [CrossRef]
  46. Zhao, Y.; Lv, J.; Liu, H.; Wu, J.; Tong, J. Study on the polarity and molar surface Gibbs energy of ether-based amino acid ionic liquids [COC4mim][Gly], [COC4mim][Ala] and [COC4mim][Thr]. J. Mol. Liq. 2021, 336, 116094. [Google Scholar] [CrossRef]
  47. Hildebrand, J.H.; Scott, R.L. The Solubility of Nonelectrolytes; Reinhold Publishing Corporation: New York, NY, USA; London Chapman and Hall: London, UK, 1950. [Google Scholar]
  48. Lawson, D.D.; Ingham, J. Estimation of solubility parameters from refractive index data. Nature 1969, 223, 614. [Google Scholar] [CrossRef]
  49. Deetlefs, M.; Seddon, K.R.; Shara, M. Predicting physical properties of ionic liquids. Phys. Chem. Chem. Phys. 2006, 8, 642–649. [Google Scholar] [CrossRef]
  50. Masaki, T.; Nishikawa, K.; Shirota, H. Microscopic Study of Ionic Liquid− H2O Systems: Alkyl-Group Dependence of 1-Alkyl-3-Methylimidazolium Cation. J. Phys. Chem. B 2010, 114, 6323–6331. [Google Scholar] [CrossRef]
  51. Chen, Y.; Sun, Y.; Li, Z.; Wang, R.; Hou, A.; Yang, F. Volumetric properties of binary mixtures of ionic liquid with tributyl phosphate and dimethyl carbonate. J. Chem. Thermodyn. 2018, 123, 165–173. [Google Scholar] [CrossRef]
  52. Verevkin, S.P.; Zaitsau, D.H.; Emel’yanenko, V.N.; Yermalayeu, A.V.; Schick, C.; Liu, H.; Maginn, E.J.; Bulut, S.; Krossing, I.; Kalb, R. Making sense of enthalpy of vaporization trends for ionic liquids: New experimental and simulation data show a simple linear relationship and help reconcile previous data. J. Phys. Chem. B 2013, 117, 6473–6486. [Google Scholar] [CrossRef]
  53. Li, Z.; Sun, Y.; Zhao, D.; Zhuang, Y.; Wang, R.; Yang, F.; Liu, X.; Chen, Y. Surface tension of binary mixtures of (ionic liquid+ tributyl phosphate). J. Chem. Thermodyn. 2019, 132, 214–221. [Google Scholar] [CrossRef]
  54. Luo, H.; Baker, G.A.; Dai, S. Isothermogravimetric determination of the enthalpies of vaporization of 1-alkyl-3-methylimidazolium ionic liquids. J. Phys. Chem. B 2008, 112, 10077–10081. [Google Scholar] [CrossRef]
  55. Xu, W.-G.; Li, L.; Ma, X.-X.; Wei, J.; Duan, W.-B.; Guan, W.; Yang, J.-Z. Density, surface tension, and refractive index of ionic liquids homologue of 1-alkyl-3-methylimidazolium tetrafluoroborate [C n mim][BF4](n = 2, 3, 4, 5, 6). J. Chem. Eng. Data 2012, 57, 2177–2184. [Google Scholar] [CrossRef]
  56. Tao, G.-H.; Zou, M.; Wang, X.-H.; Chen, Z.-Y.; Evans, D.; Kou, Y. Comparison of Polarities of Room-Temperature Ionic Liquids Using FT-IR Spectroscopic Probes. Aust. J. Chem.—AUST J CHEM 2005, 58, 327–331. [Google Scholar] [CrossRef]
  57. Park, S.; Kazlauskas, R.J. Improved Preparation and Use of Room-Temperature Ionic Liquids in Lipase-Catalyzed Enantio- and Regioselective Acylations. J. Org. Chem. 2001, 66, 8395–8401. [Google Scholar] [CrossRef] [PubMed]
  58. Wu, Y.; Sasaki, T.; Kazushi, K.; Seo, T.; Sakurai, K. Interactions between Spiropyrans and Room-Temperature Ionic Liquids: Photochromism and Solvatochromism. J. Phys. Chem. B 2008, 112, 7530–7536. [Google Scholar] [CrossRef] [PubMed]
  59. Lide, D.R. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2004; Volume 85. [Google Scholar]
  60. Potangale, M.; Tiwari, S. Correlation of the empirical polarity parameters of solvate ionic liquids (SILs) with molecular structure. J. Mol. Liq. 2020, 297, 111882. [Google Scholar] [CrossRef]
  61. Spange, S.; Lienert, C.; Friebe, N.; Schreiter, K. Complementary interpretation of ET (30) polarity parameters of ionic liquids. Phys. Chem. Chem. Phys. 2020, 22, 9954–9966. [Google Scholar] [CrossRef]
  62. Tong, J.; Zheng, X.; Li, H.; Tong, J.; Liu, Q. Densities and Viscosities of Aqueous Amino Acid Ionic Liquids [Cnmim][Ala](n = 3, 4, 5). J. Chem. Eng. Data 2017, 62, 2501–2508. [Google Scholar] [CrossRef]
  63. Liu, Q.; Janssen, M.H.; van Rantwijk, F.; Sheldon, R.A. Room-temperature ionic liquids that dissolve carbohydrates in high concentrations. Green Chem. 2005, 7, 39–42. [Google Scholar] [CrossRef]
  64. Ma, X.-X.; Wei, J.; Zhang, Q.-B.; Tian, F.; Feng, Y.-Y.; Guan, W. Prediction of Thermophysical Properties of Acetate-Based Ionic Liquids Using Semiempirical Methods. Ind. Eng. Chem. Res. 2013, 52, 9490–9496. [Google Scholar] [CrossRef]
  65. Wei, J.; Bu, X.; Guan, W.; Xing, N.; Fang, D.; Wu, Y. Measurement of vaporization enthalpy by isothermogravimetrical method and prediction of the polarity for 1-alkyl-3-methylimidazolium acetate {[C n mim][OAc](n= 4, 6)} ionic liquids. RSC Adv. 2015, 5, 70333–70338. [Google Scholar] [CrossRef]
  66. Tong, J.; Yang, H.X.; Liu, R.J.; Li, C.; Xia, L.X.; Yang, J.Z. Determination of the Enthalpy of Vaporization and Prediction of Surface Tension for Ionic Liquid 1-Alkyl-3-methylimidazolium Propionate C(n)mim Pro (n=4, 5, 6). J. Phys. Chem. B 2014, 118, 12972–12978. [Google Scholar] [CrossRef] [PubMed]
  67. Tong, J.; Hong, M.; Liu, C.; Sun, A.; Guan, W.; Yang, J.-Z. Estimation of Properties of Ionic Liquids 1-Alkyl-3-methylimidazolium Lactate Using a Semiempirical Method. Ind. Eng. Chem. Res. 2013, 52, 4967–4972. [Google Scholar] [CrossRef]
  68. Wei, J.; Li, Z.; Gu, C.; Pan, Y.; Xing, N.-N.; Tong, J.; Guan, W. Determination of vaporization enthalpy for ionic liquids C(n)mim Lact (n=2, 3, 5) and applications of the molar surface Gibbs free energy. J. Therm. Anal. Calorim. 2016, 125, 547–556. [Google Scholar] [CrossRef]
  69. Wei, J.; Ma, T.; Ma, X.; Guan, W.; Liu, Q.; Yang, J. Study on thermodynamic properties and estimation of polarity of ionic liquids { C(n)mmim NTf2 (n=2, 4)}. Rsc Adv. 2014, 4, 30725–30732. [Google Scholar] [CrossRef]
  70. Verevkin, S.P. Predicting enthalpy of vaporization of ionic liquids: A simple rule for a complex property. Angew. Chem.-Int. Ed. 2008, 47, 5071–5074. [Google Scholar] [CrossRef]
  71. Hong, M.; Sun, A.; Yang, Q.; Guan, W.; Tong, J.; Yang, J.-Z. Studies on properties of ionic liquids 1-alkyl-3-methylimidazolium lactate at temperatures from (288.15 to 333.15)K. J. Chem. Thermodyn. 2013, 67, 91–98. [Google Scholar] [CrossRef]
  72. Fang, D.-W.; Tong, J.; Guan, W.; Wang, H.; Yang, J.-Z. Predicting Properties of Amino Acid Ionic Liquid Homologue of 1-Alkyl-3-methylimidazolium Glycine. J. Phys. Chem. B 2010, 114, 13808–13814. [Google Scholar] [CrossRef]
  73. Wei, J.; Chang, C.; Zhang, Y.; Hou, S.; Fang, D.; Guan, W. Prediction of thermophysical properties of novel ionic liquids based on serine C(n)mim Ser (n=3,4) using semiempirical methods. J. Chem. Thermodyn. 2015, 90, 310–316. [Google Scholar] [CrossRef]
  74. Clara, R.A.; Marigliano, A.C.G.; Solimo, H.N. Density, viscosity, isothermal (vapour plus liquid) equilibrium, excess molar volume, viscosity deviation, and their correlations for chloroform plus methyl isobutyl ketone binary system. J. Chem. Thermodyn. 2007, 39, 261–267. [Google Scholar] [CrossRef]
  75. Dragoescu, D. Refractive indices and their related properties for several binary mixtures containing cyclic ketones and chloroalkanes. J. Mol. Liq. 2015, 209, 713–722. [Google Scholar] [CrossRef]
  76. Fenclová, D.; Vrbka, P.; Dohnal, V.; Řehák, K.; García-Miaja, G. (Vapour + liquid) equilibria and excess molar enthalpies for mixtures with strong complex formation. Trichloromethane or 1-bromo-1-chloro-2,2,2-trifluoroethane (halothane) with tetrahydropyran or piperidine. J. Chem. Thermodyn. 2002, 34, 361–376. [Google Scholar] [CrossRef]
  77. Helm, R.V.; Lanum, W.J.; Cook, G.L.; Ball, J.S. Purification and Properties of Pyrrole, Pyrrolidine, Pyridine and 2-Methylpyridine. J. Phys. Chem. 1958, 62, 858–862. [Google Scholar] [CrossRef]
  78. Lagemann, R.T.; McMillan, D.R., Jr.; Woolf, W.E. Temperature Variation of Ultrasonic Velocity in Liquids. J. Chem. Phys. 1949, 17, 369–373. [Google Scholar] [CrossRef]
  79. Colnay, M.E.; Vasseur, A.; Guerin, M. The influence of non-polar solvents on molecular dielectric polarization. 1: An attempt to select by experiment appropriate expressions. J. Chem. Res 1983, 9, 220–221. [Google Scholar]
  80. Gill, D.S.; Singh, P.; Singh, J.; Singh, P.; Senanayake, G.; Hefter, G.T. Ultrasonic velocity, conductivity, viscosity and calorimetric studies of copper(i) and sodium perchlorates in cyanobenzene, pyridine and cyanomethane. J. Chem. Soc.-Faraday Trans. 1995, 91, 2789–2795. [Google Scholar] [CrossRef]
  81. Sharma, B.R.; Singh, P.P. Excess Gibbs energies of mixing for some binary mixtures. J. Chem. Eng. Data 1975, 20, 360–363. [Google Scholar] [CrossRef]
  82. Kyte, C.; Jeffery, G.; Vogel, A. 864. Physical properties and chemical constitution. Part XXVIII. Pyridine derivatives. J. Chem. Soc. (Resumed) 1960, 4454–4472. [Google Scholar] [CrossRef]
  83. Ortega, J. Densities and refractive-indexes of pure alcohols as a function of temperature. J. Chem. Eng. Data 1982, 27, 312–317. [Google Scholar] [CrossRef]
  84. Vijande, J.; Piñeiro, M.M.; García, J.; Valencia, A.J.L.; Legido, J.L. Density and Surface Tension Variation with Temperature for Heptane + 1-Alkanol. J. Chem. Eng. Data 2006, 51, 1778–1782. [Google Scholar] [CrossRef]
  85. Das, K.N.; Habibullah, M.; Rahman, I.M.M.; Hasegawa, H.; Uddin, M.A.; Saifuddin, K. Thermodynamic Properties of the Binary Mixture of Hexan-1-ol with m-Xylene at T = (303.15, 313.15, and 323.15) K. J. Chem. Eng. Data 2009, 54, 3300–3302. [Google Scholar] [CrossRef]
  86. Neyband, R.S.; Yousefi, A.; Zarei, H. Experimental and Computational Thermodynamic Properties of (Benzyl Alcohol plus Alkanols) Mixtures. J. Chem. Eng. Data 2015, 60, 2291–2300. [Google Scholar] [CrossRef]
  87. Venkatramana, L.; Sivakumar, K.; Gardas, R.; Reddy, K.D. Effect of chain length of alcohol on thermodynamic properties of their binary mixtures with benzylalcohol. Thermochim. Acta 2014, 581, 123–132. [Google Scholar] [CrossRef]
  88. Weissler, A. Ultrasonic Investigation of Molecular Properties of Liquids. II. 1 The Alcohols1a. J. Am. Chem. Soc. 1948, 70, 1634–1640. [Google Scholar] [CrossRef] [PubMed]
  89. Huang, T.-T.; Yeh, C.-T.; Tu, C.-H. Densities, Viscosities, Refractive Indices, and Surface Tensions for the Ternary Mixtures of 2-Propanol + Benzyl Alcohol + 2-Phenylethanol at T = 308.15 K. J. Chem. Eng. Data 2008, 53, 1203–1207. [Google Scholar] [CrossRef]
  90. Hiers, G.S.; Adams, R. Omega-cyclohexyl derivatives of various normal aliphatic acids. IV. J. Am. Chem. Soc. 1926, 48, 2385–2393. [Google Scholar] [CrossRef]
  91. Hovorka, S.; Roux, A.H.; Roux-Desgranges, G.; Dohnal, V. Limiting partial molar excess heat capacities and volumes of selected organic compounds in water at 25 degrees C. J. Solut. Chem. 1999, 28, 1289–1305. [Google Scholar] [CrossRef]
  92. Shinomiya, T. Dielectric Relaxation and Intermolecular Association of Alicyclic Alcohols in Liquid and Solid States. Bull. Chem. Soc. Jpn. 1990, 63, 1087–1092. [Google Scholar] [CrossRef]
  93. Lasich, M.; Moodley, T.; Bhownath, R.; Naidoo, P.; Ramjugernath, D. Liquid–Liquid Equilibria of Methanol, Ethanol, and Propan-2-ol with Water and Dodecane. J. Chem. Eng. Data 2011, 56, 4139–4146. [Google Scholar] [CrossRef]
  94. Zarei, H.A.; Shahvarpour, S. Volumetric Properties of Binary and Ternary Liquid Mixtures of 1-Propanol (1) + 2-Propanol (2) + Water (3) at Different Temperatures and Ambient Pressure (81.5 kPa). J. Chem. Eng. Data 2008, 53, 1660–1668. [Google Scholar] [CrossRef]
  95. Ashcroft, S.J.; Clayton, A.D.; Shearn, R.B. Isothermal vapor-liquid equilibriums for the systems toluene-n-heptane, toluene-propan-2-ol, toluene-sulfolane, and propan-2-ol-sulfolane. J. Chem. Eng. Data 1979, 24, 195–199. [Google Scholar] [CrossRef]
  96. Ritzoulis, G. Excess properties of the binary liquid systems dimethylsulfoxide + isopropanol and propylene carbonate + isopropanol. Can. J. Chem. 1989, 67, 1105–1108. [Google Scholar] [CrossRef]
  97. Almasi, M. Densities and viscosities of binary mixtures of ethylmethylketone and 2-alkanols; application of the ERAS model and cubic EOS. Thermochim. Acta 2013, 554, 25–31. [Google Scholar] [CrossRef]
  98. Ruostesuo, P.; Pirilahonkanen, P. Thermodynamic and spectroscopic properties of 2-pyrrolidinones. 2. Dielectric-Properties of 2-pyrrolidinone in binary-mixtures. J. Solut. Chem. 1990, 19, 473–482. [Google Scholar] [CrossRef]
  99. Olivieri, G.V.; da Cunha, C.S.; dos Santos Martins, L.; Paegle, P.A.M.; Nuncio, S.D.; de Araújo Morandim-Giannetti, A.; Torres, R.B. Thermodynamic and spectroscopic study of binary mixtures of n-butylammonium oleate ionic liquid plus alcohol at T=288.15-308.15 K. J. Therm. Anal. Calorim. 2018, 131, 2925–2942. [Google Scholar] [CrossRef]
  100. Rafiee, H.R.; Frouzesh, F.; Miri, S. Volumetric properties for binary mixtures of ethyl acetate, vinyl acetate and tert-butyl acetate with 1-propanol and iso-butanol at T = (293.15–313.15) K and P=0.087 MPa. J. Mol. Liq. 2016, 213, 255–267. [Google Scholar] [CrossRef]
  101. Zaoui-Djelloul-Daouadji, M.; Mokbel, I.; Bahadur, I.; Negadi, A.; Jose, J.; Ramjugernath, D.; Ebenso, E.E.; Negadi, L. Vapor-liquid equilibria, density and sound velocity measurements of (water or methanol or ethanol+1,3-propanediol) binary systems at different temperatures. Thermochim. Acta 2016, 642, 111–123. [Google Scholar] [CrossRef]
  102. Vasanthakumar, A.; Bahadur, I.; Redhi, G.G.; Gengan, R.M.; Anand, K. Synthesis, characterization and thermophysical properties of ionic liquid N-methyl-N-(2′,3′-epoxypropyl)-2-oxopyrrolidinium chloride and its binary mixtures with water or ethanol at different temperatures. J. Mol. Liq. 2016, 219, 685–693. [Google Scholar] [CrossRef]
  103. Yao, H.; Zhang, S.; Wang, J.; Zhou, Q.; Dong, H.; Zhang, X. Densities and Viscosities of the Binary Mixtures of 1-Ethyl-3-methylimidazolium Bis(trifluoromethylsulfonyl)imide with N-Methyl-2-pyrrolidone or Ethanol at T = (293.15 to 323.15) K. J. Chem. Eng. Data 2012, 57, 875–881. [Google Scholar] [CrossRef]
  104. Yan, J.-H.; Dai, L.-Y.; Wang, X.-Z.; Chen, Y.-Q. Densities and Viscosities of Binary Mixtures of Cyclopropanecarboxylic Acid with Methanol, Ethanol, Propan-1-ol, and Butan-1-ol at Different Temperatures. J. Chem. Eng. Data 2009, 54, 1147–1152. [Google Scholar] [CrossRef]
  105. Zarei, H.A.; Mirhidari, N.; Zangeneh, Z. Densities, Excess Molar Volumes, Viscosity, and Refractive Indices of Binary and Ternary Liquid Mixtures of Methanol (1) + Ethanol (2) + 1,2-Propanediol (3) at P = 81.5 kPa. J. Chem. Eng. Data 2009, 54, 847–854. [Google Scholar] [CrossRef]
  106. Fan, W.; Zhou, Q.; Zhang, S.; Yan, R. Excess molar volume and viscosity deviation for the methanol plus methyl methacrylate binary system at T = (283.15 to 333.15) K. J. Chem. Eng. Data 2008, 53, 1836–1840. [Google Scholar] [CrossRef]
Figure 1. Density (ρ), surface tension (γ), and refractive index (nD) plotted against water content (w2) at various temperatures for [C1OC2mim][Ala] (a,c,e, respectively) and for [C2OC2mim][Ala] (b,d,f, respectively). ■ 288.15 K; ● 293.15 K; ▲ 298.15 K; ▼ 303.15 K; ◆ 308.15 K; ⏴ 313.15 K; ⏵ 318.15 K; ⯃ 323.15 K; ★ 328.15 K.
Figure 1. Density (ρ), surface tension (γ), and refractive index (nD) plotted against water content (w2) at various temperatures for [C1OC2mim][Ala] (a,c,e, respectively) and for [C2OC2mim][Ala] (b,d,f, respectively). ■ 288.15 K; ● 293.15 K; ▲ 298.15 K; ▼ 303.15 K; ◆ 308.15 K; ⏴ 313.15 K; ⏵ 318.15 K; ⯃ 323.15 K; ★ 328.15 K.
Molecules 27 03231 g001
Scheme 1. Thermodynamic cycle for calculating ΔAHm0.
Scheme 1. Thermodynamic cycle for calculating ΔAHm0.
Molecules 27 03231 sch001
Figure 2. Molar surface Gibbs energy (gs) plotted against temperature (T). ■ [C1OC2mim][Ala]: gs = 19,803 − 17.39T, r2 = 0.995, sd = 16.4; ● [C2OC2mim][Ala]: gs = 20,658 − 19.94T, r2 = 0.996, sd = 17.5.
Figure 2. Molar surface Gibbs energy (gs) plotted against temperature (T). ■ [C1OC2mim][Ala]: gs = 19,803 − 17.39T, r2 = 0.995, sd = 16.4; ● [C2OC2mim][Ala]: gs = 20,658 − 19.94T, r2 = 0.996, sd = 17.5.
Molecules 27 03231 g002
Figure 3. Plot of surface tension γ(est) against γ(exp) for [CnOC2mim][Ala](n = 1, 2). γ(est) =1.0581γ(exp) − 2.9851; r2 = 0.995; sd = 0.11.
Figure 3. Plot of surface tension γ(est) against γ(exp) for [CnOC2mim][Ala](n = 1, 2). γ(est) =1.0581γ(exp) − 2.9851; r2 = 0.995; sd = 0.11.
Molecules 27 03231 g003
Figure 4. Plot of surface tension γ(est) against γ(exp) for various ionic and molecular liquids (see Table S6). γ(est) = 1.0009γ(exp) − 0.03643; r2 = 0.999; sd = 0.10.
Figure 4. Plot of surface tension γ(est) against γ(exp) for various ionic and molecular liquids (see Table S6). γ(est) = 1.0009γ(exp) − 0.03643; r2 = 0.999; sd = 0.10.
Molecules 27 03231 g004
Figure 5. Inverse of the dielectric constant (ε−1) of various molecular liquids plotted against the estimated polarity PN. Linear correlation coefficient (r2) 0.94.
Figure 5. Inverse of the dielectric constant (ε−1) of various molecular liquids plotted against the estimated polarity PN. Linear correlation coefficient (r2) 0.94.
Molecules 27 03231 g005
Figure 6. Chemical structures of ionic liquids [CnOC2mim][Ala](n = 1, 2).
Figure 6. Chemical structures of ionic liquids [CnOC2mim][Ala](n = 1, 2).
Molecules 27 03231 g006
Table 1. Density (ρ), surface tension (γ), refractive index (nD), and thermal expansion coefficient (α) at various temperatures (T) for [CnOC2mim][Ala](n = 1, 2).
Table 1. Density (ρ), surface tension (γ), refractive index (nD), and thermal expansion coefficient (α) at various temperatures (T) for [CnOC2mim][Ala](n = 1, 2).
T (K)ρ (g·cm−3)γ (mJ·m−2)nDα (K−1 × 104)
[C1OC2mim][Ala]
288.151.1607351.61.51125.8555
293.151.1574451.21.50975.8722
298.151.1542350.91.50805.8885
303.151.1508250.51.50665.9060
308.151.1473950.01.50485.9236
313.151.1439549.71.50385.9414
318.151.1404949.31.50195.9595
323.151.1370348.81.50045.9776
328.151.1336548.41.49915.9954
[C2OC2mim][Ala]
288.151.1384949.61.49425.8490
293.151.1347549.31.49275.8683
298.151.1319048.91.49145.8830
303.151.1285448.51.48995.9005
308.151.1251448.01.48855.9184
313.151.1217547.61.48695.9363
318.151.1183847.21.48545.9541
323.151.1152146.71.48395.9711
328.151.1116646.31.48245.9901
Standard uncertainties (u) are u(T) = 0.02 K and u(p) = 10 kPa; expanded uncertainties (U) are U(ρ) = 0.002 g·cm−3, U(γ) = 0.3 mJ·m−2, and U(nD) = 0.003, with 95% confidence (k = 2).
Table 2. Molar surface Gibbs energy (gs), G0, G1, and estimated molar surface Gibbs energy (gs(est)) for [CnOC2mim][Ala](n = 1, 2).
Table 2. Molar surface Gibbs energy (gs), G0, G1, and estimated molar surface Gibbs energy (gs(est)) for [CnOC2mim][Ala](n = 1, 2).
T (K)gs (kJ·mol−1)G0G1gs(est) (kJ·mol−1)
[C1OC2mim][Ala]
288.1514.7819,80317.414.79
293.1514.6919,80317.414.70
298.1514.6319,80317.414.62
303.1514.5519,80317.414.53
308.1514.4319,80317.414.44
313.1514.3719,80317.414.35
318.1514.2919,80317.414.27
323.1514.1719,80317.414.18
328.1514.0819,80317.414.09
[C2OC2mim][Ala]
288.1514.8820,65819.914.92
293.1514.8220,65819.914.82
298.1514.7320,65819.914.72
303.1514.6320,65819.914.63
308.1514.5120,65819.914.53
313.1514.4220,65819.914.43
318.1514.3320,65819.914.33
323.1514.2020,65819.914.23
328.1514.1120,65819.914.13
Table 3. Molar refraction (Rm), mean molecular polarizability (αp), and estimated surface tension (γ(est)) of [CnOC2mim][Ala](n = 1, 2).
Table 3. Molar refraction (Rm), mean molecular polarizability (αp), and estimated surface tension (γ(est)) of [CnOC2mim][Ala](n = 1, 2).
T (K)Rmαp × 1024γ(est)
[C1OC2mim][Ala]
288.1557.7422.9151.6
293.1557.7722.9251.2
298.1557.7622.9250.8
303.1557.8022.9350.4
308.1557.8222.9450.0
313.1557.8422.9549.6
318.1557.8922.9749.2
323.1557.9022.9748.8
328.1557.8922.9748.4
[C2OC2mim][Ala]
288.1562.2424.6949.4
293.1562.2824.7149.0
298.1562.3024.7248.6
303.1562.3224.7348.2
308.1562.3624.7447.7
313.1562.3724.7547.3
318.1562.3924.7646.9
323.1562.4124.7646.5
328.1562.4324.7746.1
Table 4. Molar surface entropy (s) of various EFILs.
Table 4. Molar surface entropy (s) of various EFILs.
ILs (J·mol−1·K−1)V × 104 (m3·mol−1)
[C1OC2mim]Cl [38]15.991.52
[C1OC2mim][Ala] 17.391.99
[C1OC2mim][Thr] [45]26.602.18
[C1OC2mim][NTf2] [19]17.802.80
[C2OC2mim]Cl [38]17.811.68
[C2OC2mim][Ala]19.942.15
[C2OC2mim][Thr] [45]28.122.37
[C2OC2mim][NTf2] [19]19.322.99
[C1OC4mim][Gly] [46]19.562.19
[C1OC4mim][Ala] [46]20.542.29
[C1OC4mim][Thr] [46]22.0822.51
[C2OC1mim][NTf2] [19]18.202.81
[C1OC3mim][NTf2] [19]18.492.97
[C3OC2mim][NTf2] [19]20.573.16
Table 5. Polarity of various ether-functionalized ionic liquids using the new PN scale.
Table 5. Polarity of various ether-functionalized ionic liquids using the new PN scale.
ILPN (J·mol−1·K−1)Reference
[COC2mim]Cl12.86[37]
[C2OC2mim]Cl16.76[37]
[C1OC2mim][Ala]12.26This work
[C2OC2mim][Ala]18.33This work
[COC4mim][Ala]21.83[39]
[COC2mim][Thr]22.18[38]
[C2OC2mim][Thr]25.46[38]
[COC4mim][Thr]24.52[39]
[COC2mim][NTf2]31.31[13]
[C2OCmim][NTf2]35.40[13]
[C1OC3mim][NTf2]35.57[13]
[C2OC2mim][NTf2]42.66[13]
[C3OC2mim][NTf2]55.1[13]
[COC4mim][Gly]19.64[39]
Table 6. Polarity of various ionic liquids estimated using the PN scale.
Table 6. Polarity of various ionic liquids estimated using the PN scale.
ILPN (J·mol−1·K−1)
[C6mim]OAC35.73
[C4mim]OAC23.24
[C4mim]NTf2 [51,53]22.03
[C4mmim]NTf219.96
[C4mim]BF414.50
[C2mmim]NTf214.01
[C5mim]Lact12.87
[C2mim]BF48.24
[C2mim]Lact8.16
Table 7. PN and dielectric constant (ε) of various molecular liquids.
Table 7. PN and dielectric constant (ε) of various molecular liquids.
Molecular LiquidPNε
Ethyl acetate169.146.1
Chloroform167.194.8
Tetrahydrofuran124.297.5
Pyridine60.6412.3
Benzyl alcohol48.1313.0
1-Hexanol27.0613.0
Cyclohexanol25.0915.0
1-Propanol8.4820.3
Ethanol5.5225.3
Methanol2.4033.0
Table 8. Source and purity of reagents.
Table 8. Source and purity of reagents.
Reagent NameCAS No.SourcePurification Mass Fraction PurityAnalysis
Anion exchange resin 717122560-63-8SRCNoneGranularity > 0.950GC
N-Methylimidazole616-47-7ACROSDistillation>0.990FM
2-Chloroethyl methyl ether627-42-9SRCNone>0.995FM
2-Chloroethyl ethyl ether628-34-2SRCNone>0.995FM
DL-Alanine302-72-7SRCRecrystallization>0.990FM
Acetonitrile75-05-8SRCNone>0.995FM
Ethyl acetate141-78-6SRCNone>0.995FM
Anhydrous ethanol64-17-5SRCNone>0.995FM
Sodium hydroxide1310-73-2SRCNone>0.960FM
[C1OC2mim][Ala] -SynthesisSolvent extraction, vacuum drying >0.9901H, 13C NMR
[C2OC2mim][Ala] -SynthesisSolvent extraction, vacuum drying >0.9901H, 13C NMR
FM—Fractional melting; SRC—Shanghai Reagent Co., Ltd.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, X.; Guo, C.; Wu, W.; Tong, J. A New Compartmentalized Scale (PN) for Measuring Polarity Applied to Novel Ether-Functionalized Amino Acid Ionic Liquids. Molecules 2022, 27, 3231. https://doi.org/10.3390/molecules27103231

AMA Style

Zheng X, Guo C, Wu W, Tong J. A New Compartmentalized Scale (PN) for Measuring Polarity Applied to Novel Ether-Functionalized Amino Acid Ionic Liquids. Molecules. 2022; 27(10):3231. https://doi.org/10.3390/molecules27103231

Chicago/Turabian Style

Zheng, Xu, Chun Guo, Wenqing Wu, and Jing Tong. 2022. "A New Compartmentalized Scale (PN) for Measuring Polarity Applied to Novel Ether-Functionalized Amino Acid Ionic Liquids" Molecules 27, no. 10: 3231. https://doi.org/10.3390/molecules27103231

Article Metrics

Back to TopTop