DOI:
10.1039/D5RA06386H
(Paper)
RSC Adv., 2025,
15, 38865-38875
1-Ethyl-4-butyl-1,2,4-triazolium acetyl amino acid ionic liquids: preparation and characterization of physicochemical properties
Received
27th August 2025
, Accepted 7th October 2025
First published on 16th October 2025
Abstract
In order to address certain performance limitations of imidazolium amino acid ionic liquids, three novel 1-ethyl-4-butyl-1,2,4-triazolium acetyl amino acid ionic liquids, namely [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] were synthesized and characterized in this work. Firstly, based on experimentally determined refractive indices and previously reported data including density, surface tension, viscosity, and heat capacity, the polarization properties, surface properties, and volumetric properties were calculated. Special emphasis was placed on elucidating the influence of acetyl amino acid anion structure on the physicochemical properties of these ionic liquids. Subsequently, by integrating molar surface Gibbs energy with the Lorentz–Lorenz relation, we established a predictive equation for surface tension, which demonstrated excellent agreement with experimental values. Analysis using the polarity coefficient equation revealed that the polarity of the ionic liquids decreases with increasing anion side-chain length, corresponding to enhanced hydrophobicity. Finally, the Eyring viscosity equation and thermodynamic formulas were employed to derive a computational expression for viscous flow activation energy, yielding results consistent with those obtained from the Arrhenius equation. Current limited research on 1-ethyl-4-butyl-1,2,4-triazolium acetyl amino acid ionic liquids necessitates combined experimental and semi-empirical approaches to guide their functional design and industrial applications.
1. Introduction
Ionic liquids (ILs), recognized as green and designer solvents, have garnered significant attention and application across various industries.1–5 While imidazolium cations are prevalent in current ILs, their inherent drawback lies in the acidic C2 proton of the imidazolium ring. This acidity facilitates decomposition under basic conditions via carbene formation, potentially inducing side reactions.6 Substituting imidazolium with triazolium cations—either 1,2,3-triazolium or 1,2,4-triazolium—mitigates this instability.7–10 Given the hazardous azides typically employed in synthesizing 1,2,3-triazole derivatives, 1,2,4-triazolium is frequently the more suitable cationic platform.11
Amino acids are often employed as biocompatible anions to impart specific functionality to ILs.12–16 A key limitation of amino acid-based ILs, however, is their relatively low thermal decomposition temperature compared to ILs with conventional anions, hindering broader application. The corresponding acetylated amino acids offer enhanced thermal stability and higher molar heat capacity compared to their amino acid precursors.17 Consequently, this study focuses on the development and characterization of novel, non-toxic ionic liquids comprising 1-ethyl-4-butyl-1,2,4-triazolium cations paired with acetyl amino acid anions. Investigation of their thermal properties aims to support practical utilization.18
Although our prior work has evaluated the feasibility of these novel ILs as sustainable heat-transfer fluids,19–24 their characterization primarily focused on key properties for thermal applications, such as heat capacity, thermal conductivity, heat-storage density, thermal stability, and heat-transfer coefficients/areas in heat exchangers. Nevertheless, research on triazolium acetyl amino acid ILs, particularly their fundamental physicochemical properties, remains far from exhaustive, given the vast potential for novel structural designs and applications. Industry advancement necessitates such development, driving the demand for robust methods to characterize physico-chemical properties. Accurate prediction of IL behavior via semi-empirical methods critically depends on this data.
Building upon prior work with ILs,24,25 the compounds [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] were synthesized. Their refractive indices were experimentally determined. Utilizing these refractive indices alongside established literature data (density, surface tension, viscosity, heat capacity),24 a range of thermophysical properties were computed. It is noteworthy that while numerous studies have examined the impact of varying methylene chain lengths on cations (e.g., imidazolium) on IL properties, analogous effects on anions remain comparatively underexplored.26–29 Therefore, this study not only addresses the fundamental research gap left by the previous application-oriented work and advances the quantitative understanding of the structure–property relationships in this class of ILs, but also complements the fundamental database of ILs by offering new analytical tools and design concepts.
2. Experimental
2.1 Reagents
The CAS number, purities and sources of the reagents are listed in Table 1.
Table 1 The CAS number, purity and source of the reagents
Reagent |
CAS |
Purity |
Source |
Bromoethane |
74-96-4 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
Bromobutane |
190-65-9 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
1,2,4-Triazole |
288-88-0 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
N-Acetyl-glycine |
543-24-8 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
N-Acetyl-L-alanine |
97-69-8 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
N-Acetyl-L-cysteine |
616-91-1 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
Methanol |
67-56-1 |
99% |
Tianjin Fuyu chemical Co.,Ltd |
Sodium methoxide |
124-41-4 |
99% |
Shanghai Macklin Biochemical Co.,Ltd |
Acetonitrile |
75-05-8 |
99% |
Tianjin Fuyu chemical Co.,Ltd |
Dichloromethane |
75-09-2 |
99% |
Tianjin Fuyu chemical Co.,Ltd |
Amberjet 4200-CL |
60 177-39-1 |
Particle size > 0.950 |
Sinopharm Chemical Reagent Co., Ltd |
Deionized water |
— |
— |
Manufactured by multiple distillation |
2.2 Preparation
In this study, the synthesis of ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala], andaz(2,4)][Accys] was accomplished through a four-step procedure, as illustrated in Fig. 1. Firstly, a mixture of 1,2,4-triazole, methanol, and sodium methoxide was reacted with bromoethane at 65 °C for 48 h to yield 1-ethyl-1,2,4-triazole. Secondly, the intermediate 1-ethyl-1,2,4-triazole was further reacted with bromopentane at 70 °C for 24 h to produce the bromide-based IL precursor, [Taz(2,4)][Br]. Then, [Taz(2,4)][Br] was dissolved in distilled water and passed through a pre-treated anion exchange column packed with Amberjet 4200-Cl resin. The effluent was collected once it showed strong alkalinity and the absence of bromide ions was confirmed by AgNO3–HNO3 testing, yielding [Taz(2,4)][OH]. Finally, the aqueous solution of [Taz(2,4)][OH] was reacted with an equimolar amount of N-acetyl-L-glycine/alanine/cysteine at room temperature to obtain the target ILs. The products were subsequently purified using organic solvents, including acetonitrile and dichloromethane.
 |
| Fig. 1 Synthesis of the novel ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys]. | |
2.3 Characterization
The [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] were characterized by 1H-NMR spectroscopy, 13C-NMR spectroscopy, FT-IR spectroscopy and thermogravimetry (TG) (see Fig. S1–S12) and the water content was determined by a Karl Fischer moisture titrator (ZSD-2 type).
[Taz(2,4)][Acgly] was obtained as a light yellow transparent liquid. 1H-NMR (DMSO-d6): δH = 10.59 (s, 1H), 9.32 (s, 1H), 7.55 (d, 1H), 4.44 (m, 2H), 4.26 (t, 2H), 3.43 (d, 2H), 1.84 (m, 5H), 1.49 (t, 3H), 1.32 (m, 2H), 0.93 (t, 3H). 13C-NMR (DMSO-d6): 172.03, 168.88, 144.91, 142.97, 47.22, 47.15, 43.34, 31.08, 22.71, 18.97, 13.79, 13.43. IR (500–4500 cm−1, KBr pellet): 3294, 2964, 2935, 2883, 1954, 1457, 1364, 1164, 776. According to the whole TG curve, the sample initial decomposition temperature is about 453.16 K, and the mass fraction of water content is (0.00525 ± 0.0001). Thus, the purity of the [Taz(2,4)][Acgly] could be estimated more than 99%.
[Taz(2,4)][Acala] was obtained as a light yellow transparent liquid. 1H-NMR (DMSO-d6): δH = 10.61 (s, 1H), 9.32 (s, 1H), 7.59 (d, 1H), 4.44 (m, 2H), 4.29 (t, 2H), 3.88 (m, 1H), 1.82 (m, 5H), 1.49 (t, 3H), 1.29 (m, 2H), 1.16 (d, 3H), 0.93 (t, 3H). 13C-NMR (DMSO-d6): 175.34, 168.05, 144.98, 143.15, 49.71, 47.21, 47.13, 31.12, 22.90, 19.04, 18.98, 13.80, 13.16. IR (500–4500 cm−1, KBr pellet): 3400, 3274, 2966, 2930, 2877, 1651, 1594, 1454, 1398, 1164, 646. According to the whole TG curve, the sample initial decomposition temperature is about 453.67 K, and the mass fraction of water content is (0.00521 ± 0.0001). Thus, the purity of the [Taz(2,4)][Acala] could be estimated more than 99%.
[Taz(2,4)][Accys] was obtained as a colorless transparent liquid. 1H-NMR (DMSO-d6): δH = 10.59 (s, 1H), 9.32 (s, 1H), 7.58 (d, 1H), 4.42 (m, 2H), 4.27 (t, 2H), 4.01 (m, 1H), 2.78 (m, 3H), 1.82 (m, 5H), 1.47 (t, 3H), 1.32 (m, 2H), 0.91 (t, 3H). 13C-NMR (DMSO-d6): 172.00, 168.87, 144.90, 142.98, 55.96, 47.25, 47.18, 31.11, 27.14, 22.95, 19.01, 13.84, 13.47. IR (500–4500 cm−1, KBr pellet): 3401, 3274, 2963, 2930, 2874, 1601, 1464, 1365, 1164, 776. According to the whole TG curve, the sample initial decomposition temperature is about 452.26 K, and the mass fraction of water content is (0.00344 ± 0.0001). Thus, the purity of the [Taz(2,4)][Accys] could be estimated more than 99%.
2.4. Determination of refractive indexes
The refractive index (nD) was measured using a WZS-1 Abbe refractometer with an accuracy of ± 0.0002. The instrument was coupled with a GDH-2008W high-precision low-temperature thermostatic bath (ΔT = ± 0.02 K) to ensure temperature stability during measurements. Prior to measurement, the system was equilibrated at the target temperature for 30 minutes to achieve thermal stability. For hydrophilic ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] that readily form hydrogen bonds with water, the standard addition method (SAM) was employed to eliminate interference from residual water. Refractive index measurements were conducted across a temperature range of 288.15–323.15 K at 5 K intervals for ionic liquid samples with varying water contents. The intrinsic physical properties of anhydrous ILs were determined through linear regression analysis, where the intercept of the refractive index-water content curve provided the water-free reference value.
3. Results and discussion
3.1 Refractive index
The refractive index values of [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] with varying water contents were determined as the average of triplicate measurements, summarized in Table 2 and represented in Fig. S13–S15. At constant temperature, the refractive index exhibited a linear correlation with water content (w2), with all linear regression coefficients exceeding 0.99. Standard deviations of the regressions remained within experimental error ranges, confirming the validity of the SAM for these ILs. The intercepts of these linear fits provided the experimental refractive index values for anhydrous ILs (Fig. 2).
Table 2 Refractive index, nD, values of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] containing various mass fraction of water, w2, at T = (288.15–323.15) Ka
Standard uncertainties u are u(T) = 0.02 K, u(p) = 0.01 MPa, u(w2) = 0.0001, and the expanded uncertainty U(nD) = 0.0032 with 0.95 level of confidence (k ≈ 2); w2 is the various mass fraction of water; r2 is the correlation coefficient square and sd is the standard deviation. |
T/K |
103 w2 = 5.25 |
103 w2 = 8.79 |
103 w2 = 11.76 |
103 w2 = 14.97 |
103 w2 = 17.68 |
103 w2 = 0 |
r2 |
sd ×10−5 |
[Taz(2,4)][Acgly] |
288.15 |
1.4953 |
1.4949 |
1.4945 |
1.4941 |
1.4937 |
1.4960 |
0.998 |
4.00 |
293.15 |
1.4941 |
1.4937 |
1.4933 |
1.4929 |
1.4926 |
1.4948 |
0.999 |
2.69 |
298.15 |
1.4929 |
1.4925 |
1.4921 |
1.4918 |
1.4915 |
1.4935 |
0.996 |
4.29 |
303.15 |
1.4920 |
1.4916 |
1.4912 |
1.4908 |
1.4905 |
1.4927 |
0.999 |
2.69 |
308.15 |
1.4911 |
1.4906 |
1.4902 |
1.4899 |
1.4895 |
1.4917 |
0.994 |
6.11 |
313.15 |
1.4899 |
1.4894 |
1.4891 |
1.4887 |
1.4883 |
1.4906 |
0.996 |
5.03 |
318.15 |
1.4886 |
1.4881 |
1.4878 |
1.4873 |
1.4870 |
1.4893 |
0.996 |
5.28 |
323.15 |
1.4873 |
1.4869 |
1.4865 |
1.4861 |
1.4857 |
1.4880 |
0.998 |
3.99 |
T/K |
103 w2 = 5.21 |
103 w2 = 8.46 |
103 w2 = 11.03 |
103 w2 = 13.28 |
103 w2 = 15.37 |
103 w2 = 0 |
r2 |
sd × 10−5 |
[Taz(2,4)][Acala] |
288.15 |
1.4920 |
1.4917 |
1.4914 |
1.4911 |
1.4909 |
1.4926 |
0.995 |
4.60 |
293.15 |
1.4911 |
1.4907 |
1.4904 |
1.4901 |
1.4899 |
1.4917 |
0.998 |
2.98 |
298.15 |
1.4901 |
1.4896 |
1.4893 |
1.4889 |
1.4887 |
1.4908 |
0.993 |
6.53 |
303.15 |
1.4890 |
1.4886 |
1.4882 |
1.4878 |
1.4876 |
1.4898 |
0.992 |
7.27 |
308.15 |
1.4878 |
1.4873 |
1.4869 |
1.4866 |
1.4863 |
1.4886 |
0.999 |
2.79 |
313.15 |
1.4860 |
1.4856 |
1.4851 |
1.4848 |
1.4845 |
1.4868 |
0.993 |
6.91 |
318.15 |
1.4844 |
1.4839 |
1.4836 |
1.4832 |
1.4829 |
1.4852 |
0.996 |
5.00 |
323.15 |
1.4826 |
1.4822 |
1.4818 |
1.4814 |
1.4811 |
1.4836 |
0.994 |
6.33 |
T/K |
103 w2 = 3.44 |
103 w2 = 7.45 |
103 w2 = 10.94 |
103 w2 = 14.07 |
103 w2 = 17.29 |
103 w2 = 0 |
r2 |
sd × 10−5 |
[Taz(2,4)][Accys] |
288.15 |
1.5183 |
1.5179 |
1.5175 |
1.5172 |
1.5168 |
1.5187 |
0.998 |
2.78 |
293.15 |
1.5172 |
1.5168 |
1.5164 |
1.5161 |
1.5158 |
1.5176 |
0.998 |
2.44 |
298.15 |
1.5162 |
1.5158 |
1.5154 |
1.5151 |
1.5148 |
1.5166 |
0.998 |
2.44 |
303.15 |
1.5149 |
1.5145 |
1.5141 |
1.5137 |
1.5134 |
1.5153 |
0.997 |
3.44 |
308.15 |
1.5134 |
1.5130 |
1.5126 |
1.5123 |
1.5120 |
1.5138 |
0.998 |
2.44 |
313.15 |
1.5113 |
1.5110 |
1.5107 |
1.5105 |
1.5102 |
1.5116 |
0.997 |
2.46 |
318.15 |
1.5101 |
1.5098 |
1.5096 |
1.5093 |
1.5091 |
1.5104 |
0.994 |
3.32 |
323.15 |
1.5089 |
1.5086 |
1.5084 |
1.5081 |
1.5079 |
1.5092 |
0.994 |
3.32 |
 |
| Fig. 2 Plot of refractive index, nD, versus temperature, T. [Taz(2,4)][Acgly]: nD = 1.5599–2.2191 × 10−4 T, r2 = 0.99, sd = 5.21 × 10−6; [Taz(2,4)][Acala]: nD = 1.5677–2.5881 × 10−4 T, r2 = 0.99, sd = 5.61 × 10−6; [Taz(2,4)][Accys]: nD = 1.6008–2.8333 × 10−4 T, r2 = 0.99, sd = 6.77 × 10−6. | |
As shown in Table 2 and Fig. 2, the refractive index of anhydrous ionic liquids exhibits an inverse correlation with temperature, primarily attributed to thermal expansion-induced reduction in molecular packing density. Furthermore, introducing methylene groups or sulfur atoms into the IL anions reduced the refractive index. This observation aligns with prior studies: Almeida et al. reported that IL refractive indices depend on anion volume,30 while Deetlefs et al. demonstrated that tighter packing of anions and cations correlates with higher refractive indices.31 This phenomenon may arise from methylene/sulfur substitutions lowering refractive indices through increased anion volume and reduced polarizability.
State and liquid state in the corres
3.2 Derived properties
3.2.1 Polarization properties. According to the Lorentz–Lorentz relation for non-polar materials connecting the static refractive index with molecular polarizability,32 the molar refraction (Rm) is defined as: |
Rm = [(nD2 − 1)/(nD2 + 2)]V = (4πN/3)αP
| (1) |
where nD is the refractive index and αp represents the molar polarizability. The Rm and αp values listed in Table 3 demonstrate that [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] exhibit temperature-independent Rm and αp, indicating that these parameters characterize the induced dipole effects in the ILs.
Table 3 Values of the refractive index (nD), molar refraction (Rm), molar polarizability (αp), and surface tension (γ) of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] in the range of T = (288.15–323.15) Ka
T/K |
nD |
Rm/cm3 mol−1 |
1024 αp/cm3 |
aγ/mJ m−2 |
Sa/mJ K−1 m−2 |
Ea/mJ m−2 |
Derived from previous work.24 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acgly] |
288.15 |
1.4960 |
68.64 |
27.21 |
45.8 |
70.2 |
66.0 |
293.15 |
1.4948 |
68.69 |
27.23 |
45.5 |
70.2 |
66.1 |
298.15 |
1.4935 |
68.74 |
27.25 |
45.2 |
70.2 |
66.1 |
303.15 |
1.4927 |
68.85 |
27.29 |
44.9 |
70.2 |
66.2 |
308.15 |
1.4917 |
68.94 |
27.33 |
44.5 |
70.2 |
66.1 |
313.15 |
1.4906 |
69.00 |
27.35 |
44.1 |
70.2 |
66.1 |
318.15 |
1.4893 |
69.05 |
27.37 |
43.7 |
70.2 |
66.0 |
323.15 |
1.4880 |
69.10 |
27.39 |
43.4 |
70.2 |
66.1 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acala] |
288.15 |
1.4926 |
73.56 |
29.16 |
46.4 |
75.5 |
66.6 |
293.15 |
1.4917 |
73.68 |
29.21 |
46.1 |
75.5 |
66.7 |
298.15 |
1.4908 |
73.78 |
29.25 |
45.7 |
75.5 |
66.6 |
303.15 |
1.4898 |
73.88 |
29.29 |
45.3 |
75.5 |
66.6 |
308.15 |
1.4886 |
73.96 |
29.32 |
44.9 |
75.5 |
66.5 |
313.15 |
1.4868 |
73.96 |
29.32 |
44.5 |
75.5 |
66.5 |
318.15 |
1.4852 |
73.98 |
29.33 |
44.2 |
75.5 |
66.5 |
323.15 |
1.4836 |
74.01 |
29.34 |
43.8 |
75.5 |
66.5 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Accys] |
288.15 |
1.5187 |
81.03 |
32.12 |
46.7 |
72.6 |
66.9 |
293.15 |
1.5176 |
81.12 |
32.16 |
46.4 |
72.6 |
67.0 |
298.15 |
1.5166 |
81.22 |
32.20 |
46.1 |
72.6 |
67.0 |
303.15 |
1.5153 |
81.29 |
32.23 |
45.7 |
72.6 |
67.0 |
308.15 |
1.5138 |
81.34 |
32.24 |
45.3 |
72.6 |
66.9 |
313.15 |
1.5116 |
81.28 |
32.22 |
44.9 |
72.6 |
66.9 |
318.15 |
1.5104 |
81.37 |
32.26 |
44.6 |
72.6 |
66.9 |
323.15 |
1.5092 |
81.45 |
32.29 |
44.2 |
72.6 |
66.9 |
3.2.2 Surface properties. The surface entropy (Sa) of a pure liquid, defined as the magnitude of the temperature-dependent slope of γ, quantifies the degree of surface disorder per unit volume. By applying a least-squares linear regression to the surface tension data of the ionic liquids (at zero water content) as a function of temperature (see Fig. 3), the empirical equation is derived:where A0 is an empirical parameter and Sa = – (∂γ/∂T)p represents the surface entropy. The Sa values for [Taz(2,4)][Acgly], [Taz(2,4)][Acala], and [Taz(2,4)][Accys] are summarized in Table 3.
 |
| Fig. 3 Plot of surface tension, γ, versus temperature, T. [Taz(2,4)][Acgly]: γ = 66.1–7.024 × 10−2 T, r2 = 0.996, sd = 1.72 × 10−3; [Taz(2,4)][Acala]: γ = 68.2–7.548 × 10−2 T, r2 = 0.999, sd = 1.07 × 10−3; [Taz(2,4)][Accys]: γ = 67.7–7.262 × 10−2 T, r2 = 0.998, sd = 1.35 × 10−3. | |
The surface energy (Ea) of ionic liquids can be calculated from their γ and Sa at zero water content using the equation:
The calculated Ea values for the ILs at various temperatures are listed in Table 3. The results indicate that the surface energy of ionic liquids decreases with the elongation of alkyl chains on the anion or the introduction of sulfur atoms, which aligns with the observations reported by Yan and Wei.23,33 This reduction is likely attributed to enhanced steric hindrance and diminished electronic effects within the ionic liquid structure. Furthermore, the Ea remains nearly constant across the temperature range of 288.15–323.15 K, indicating negligible temperature dependence of Ea. The temperature insensitivity of Ea can be attributed to the compensatory effect of Sa on energy variations and the dynamic equilibrium between intermolecular interactions and interfacial structures.34,35
Notably, the Ea of ILs is significantly lower than that of molten salts but closer to that of organic liquids. This suggests weaker interionic interactions in ILs compared to inorganic molten salts, as surface energy is determined by the strength of these interactions. The weaker interactions also explain the lower standard entropy of ILs relative to molten salts.23
3.2.3 Volumetric properties. The molecular volume (Vm) of an IL, defined as the combined volume of one cation and one anion, is calculated from density values using the equation:where M denotes molar mass, ρ represents density, N is Avogadro's constant, and V corresponds to molar volume. As shown in Table 3 for synthesized ILs at 298.15 K, the introduction of a methylene group (–CH2–) or sulfur atom (–S–) into the anion significantly increases Vm, which is closely related to changes in steric hindrance and electron cloud distribution.36Specifically, the volumetric contributions from a methylene group and sulfur atom in acetylated amino acid anions were quantified as Vm(–CH2–) = 0.0307 nm3 and Vm(–S–) = 0.0229 nm3, respectively. The larger contribution from the methylene group likely stems from alkyl chain elongation, leading to linear molecular skeleton extension and cumulative van der Waals volume effects. In contrast, the relatively smaller contribution of sulfur atoms may be attributed to their enhanced electron polarizability. The higher electronegativity of sulfur promotes structural compactness in anions, partially offsetting steric effects.37
According to Glasser's theory,19 the standard molar entropy (S0) and the crystal energy (UPOT) of the ILs can be estimated:
|
S0 = 1246.5 Vm + 29.5
| (5) |
|
UPOT = 1981.2(ρ/M)1/3 + 103.8
| (6) |
The S0 and UPOT of ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] at 298.15 K are summarized in Table 4. The data reveal that the S0 increases with the elongation of the anion side chain, indicating that longer side chains reduce structural order. Specifically, each additional –CH2– group on the acetylated amino acid anion contributes 38 J K−1 mol−1 to the S0, which aligns closely with the reported average contribution of 39 J K−1 mol−1 per –CH2– group in [C4mim][Gly] and [C4mim][Ala].19–21,38,39 Similarly, the incorporation of a sulfur atom into the anion side chain contributes 29 J K−1 mol−1 to the S0.
Table 4 Volumetric contributions of a methylene group or a sulfur atom in anions at 298.15 K
ILs |
M/g mol−1 |
aρ/g cm−3 |
104 V/m3 mol−1 |
Vm/nm3 |
ΔVm(–CH2–)/nm3 |
ΔVm(–S–)/nm3 |
Derived from previous work.24 Derived from ref. 38. |
[Taz(2,4)][Acgly] |
270.37 |
1.14395 |
2.363 |
0.3925 |
0.0307 |
|
[Taz(2,4)][Acala] |
284.41 |
1.11596 |
2.549 |
0.4232 |
0.0229 |
[Taz(2,4)][Accys] |
316.48 |
1.17804 |
2.687 |
0.4461 |
|
ILs |
S0/J K−1 mol−1 |
ΔS0(–CH2–)/J K−1 mol−1 |
ΔS0(–S–)/J K−1 mol−1 |
UPOT/kJ mol−1 |
ΔUPOT(–CH2–)/kJ mol−1 |
ΔUPOT(–S–)/kJ mol−1 |
[Taz(2,4)][Acgly] |
519 |
38 |
|
424 |
8 |
|
[Taz(2,4)][Acala] |
557 |
29 |
416 |
5 |
[Taz(2,4)][Accys] |
586 |
|
411 |
|
b[C4mim][Gly] |
426 |
39 |
|
447 |
10 |
|
b[C4mim][Ala] |
465 |
|
437 |
|
Conversely, the UPOT decreases with the extension of the anion side chain. Each –CH2– group added to the acetylated amino acid anion reduces the UPOT by 8 kJ mol−1, consistent with the average contribution of −8 kJ mol−1 per –CH2– group reported for 79 imidazolium-based ILs.40 Additionally, the introduction of a sulfur atom into the anion decreases the UPOT by 5 kJ mol−1. Compared to conventional inorganic salts, the significantly lower UPOT of ILs is the fundamental reason for their liquid state at room temperature.
3.3. The molar surface gibbs free energy and its applications
The relationship between the molar surface Gibbs energy, gs, and temperature, T, is obtained by substituting eqn (7) into the traditional Eötvös equation,31 which can be expressed aswhere V is molar volume, N is Avogadro constant and γ is surface tension. The values of gs for [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] are calculated, listed in Table 5 and plotted in Fig. 4. The predicted value of the molar surface Gibbs energy (gs(pre)) can be calculated by the intercept C0 and the slope C1 into the eqn (8) and are shown in Table 5. According to thermodynamic relations, C1 = –(∂gs/∂T)p is the molar surface entropy (s) of ILs, and the molar surface enthalpy h = gs + sT, and their values are listed in Table 5. The gs(pre) versus gs of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] were plotted in Fig. 4A. As can be seen, the h is independent with T, molar surface heat capacity is close to zero which indicates that the process from the inside to the surface of the liquid is an isocoulombic process.41
Table 5 Values of the molar surface Gibbs energy (gs), molar surface entropy (s), molar surface enthalpy (h), predicted values of molar surface Gibbs energy (gs(pre)), molar refraction (Rm), molar polarizability (αp) and predicted values of surface tension (γpre) of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] in the range of T = (288.15–323.15) K
T/K |
aρ/kg m−3 |
a104V/m3 |
gs/kJ mol−1 |
s/J mol−1 K−1 |
h/kJ mol−1 |
gs(pre)/kJ mol−1 |
γpre/mJ m−2 |
Derived from previous work.24 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acgly] |
288.15 |
1150.61 |
2.350 |
14.93 |
17.12 |
19.66 |
14.93 |
45.9 |
293.15 |
1147.28 |
2.357 |
14.83 |
17.12 |
19.68 |
14.84 |
45.5 |
298.15 |
1143.95 |
2.364 |
14.75 |
17.12 |
19.70 |
14.75 |
45.2 |
303.15 |
1140.62 |
2.370 |
14.68 |
17.12 |
19.71 |
14.66 |
44.8 |
308.15 |
1137.29 |
2.377 |
14.58 |
17.12 |
19.70 |
14.56 |
44.5 |
313.15 |
1133.97 |
2.384 |
14.47 |
17.12 |
19.68 |
14.47 |
44.1 |
318.15 |
1130.66 |
2.391 |
14.36 |
17.12 |
19.66 |
14.38 |
43.8 |
323.15 |
1127.33 |
2.398 |
14.29 |
17.12 |
19.68 |
14.29 |
43.4 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acala] |
288.15 |
1122.90 |
2.533 |
15.69 |
19.30 |
21.25 |
15.70 |
46.4 |
293.15 |
1119.43 |
2.541 |
15.62 |
19.30 |
21.27 |
15.60 |
46.1 |
298.15 |
1115.96 |
2.549 |
15.51 |
19.30 |
21.27 |
15.51 |
45.7 |
303.15 |
1112.50 |
2.557 |
15.41 |
19.30 |
21.26 |
15.41 |
45.3 |
308.15 |
1109.03 |
2.564 |
15.30 |
19.30 |
21.25 |
15.31 |
44.9 |
313.15 |
1105.56 |
2.573 |
15.20 |
19.30 |
21.24 |
15.22 |
44.6 |
318.15 |
1102.08 |
2.581 |
15.13 |
19.30 |
21.27 |
15.12 |
44.2 |
323.15 |
1098.62 |
2.589 |
15.02 |
19.30 |
21.26 |
15.02 |
43.8 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Accys] |
288.15 |
1184.92 |
2.671 |
16.36 |
19.26 |
21.91 |
16.38 |
46.8 |
293.15 |
1181.47 |
2.679 |
16.28 |
19.26 |
21.93 |
16.28 |
46.4 |
298.15 |
1178.04 |
2.687 |
16.21 |
19.26 |
21.95 |
16.19 |
46.0 |
303.15 |
1174.55 |
2.694 |
16.10 |
19.26 |
21.94 |
16.09 |
45.7 |
308.15 |
1171.04 |
2.703 |
15.99 |
19.26 |
21.93 |
15.99 |
45.3 |
313.15 |
1167.55 |
2.711 |
15.88 |
19.26 |
21.91 |
15.90 |
44.9 |
318.15 |
1164.07 |
2.719 |
15.81 |
19.26 |
21.93 |
15.80 |
44.6 |
323.15 |
1160.58 |
2.727 |
15.70 |
19.26 |
21.92 |
15.70 |
44.2 |
 |
| Fig. 4 Plot of (A) gs(pre) vs. gs and (B) γpre vs. γ for [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys]. gs(pre) = 1.00021 gs-0.00329, r2 = 0.999, sd = 0.0043; γ(pre) = 0.99767 γ + 0.11354, r2 = 0.998, sd = 0.0116. | |
The equation for predicting the surface tension of ILs can be obtained using eqn (7) and (9):
|
γpre3/2 = [gs(pre)3/2/N1/2Rm]nD2 − 1)/(nD2 + 2)
| (9) |
The predicted values of the surface tensions, γpre, for [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] were also listed in Table 5. Fig. 4B shows the predicted surface tension γpre versus experimental surface tension γ of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys]. As can be seen, the value of r2 is equal to 0.99 suggesting that the proposed equation can well predict the ILs surface tension with low errors.
The association enthalpy of ion pairs in the gas phase (ΔAH0m) serves as a critical parameter for characterizing the strength of interionic interactions within liquid ionic liquids. Thermodynamically, when ΔAH0m < 0, a larger absolute value indicates stronger ion association. The UPOT, vaporization enthalpy (ΔglH0m), and ΔAH0m of ILs are interrelated through the following equation:42
|
ΔAH0m = ΔglH0m − UPOT
| (10) |
|
ΔglH0m = 11.21 (γV2/3N1/3) + 2.4
| (11) |
Here, Δ
glH0m is calculated
via the Kabo empirical equation,
42 which depends on surface tension (
γ), molar volume (
V), and Avogadro's constant (
N). The absolute values of Δ
AH0m for various ionic liquids are listed in
Table 6.
Table 6 Values of the surface tension (γ), molar volume (V), molar surface Gibbs free energy (gs), molar vaporization enthalpy (ΔglH0m), lattice energy (UPOT) and association enthalpy of ion pairs (ΔAH0m) of [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] at 298.15 K
ILs |
γ/mJ m−2 |
104 V/m3 mol−1 |
gs/kJ mol−1 |
ΔglH0m/kJ mol−1 |
UPOT/kJ mol−1 |
ΔAH0m/kJ mol−1 |
[Taz(2,4)][Acgly] |
45.2 |
2.363 |
14.75 |
167.7 |
424 |
−256.3 |
[Taz(2,4)][Acala] |
45.7 |
2.549 |
15.51 |
176.3 |
416 |
−239.7 |
[Taz(2,4)][Accys] |
46.1 |
2.687 |
16.21 |
184.1 |
411 |
−226.9 |
Analysis of Table 6 reveals that introducing a methylene group or sulfur atom into the anion significantly reduces the absolute value of ΔAH0m when the cation structure remains constant. This trend demonstrates that increasing the relative molecular mass of the anion weakens interionic associations within the ILs. Structural modification of anions, introducing methylene group or sulfur atom, significantly modulates the internal association strength of ILs through steric hindrance and charge delocalization effects. This principle provides a theoretical foundation for designing functionalized ILs with low melting points and high fluidity, particularly in applications requiring a balance between ionic conductivity and thermal stability.
3.4. Polarity
3.4.1. New scale of the polarity, δμ. In order to more accurately describe the polarity of ILs, Tong et al. proposed a new scale of the polarity, δμ, on the basis of three assumptions.43 Based on Hildebrand's theory,44 the solubility parameter of the polar part δμ is calculated: |
δμ2 = ΔglH0m,μ/V − (1 − x)RT/V
| (12) |
where x = ΔglH0m,n/ ΔglH0m, ΔglH0m = ΔglH0m,n + ΔglH0m,μ and the δμ is a new scale of the polarity. The vaporization enthalpy ΔglH0m consists of non-polar partΔglH0m,n (the contribution part from the induced dipole moment toΔglH0m) and polar part ΔglH0m,μ (the contribution part from the average permanent dipole moment to ΔglH0m). The values of ΔglH0m and ΔglH0m,n can be estimated by using Kabo empirical equation45 and the Lawson–Ingham equation,46 respectively. |
ΔglH0m,n = C[(nD2 − 1)/(nD2 + 2)]V
| (13) |
where C = 1.297 kJ cm−3 is an empirical constant for organic liquids, nDis the refractive index and V is molar volume.According to the eqn (10)–(13), the values of δμ, ΔglH0m, ΔglH0m,n and ΔglH0m,μ for [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys] were calculated and listed in Table 7. It reveals the polarity order [Taz(2,4)][Acgly] > [Taz(2,4)][Acala] > [Taz(2,4)][Accys], demonstrating that the incorporation of methylene groups or sulfur atoms into the anion structure relatively enhances the hydrophobicity of ILs.
Table 7 The values of vaporization enthalpy (ΔglH0m), the contribution part from the induced dipole moment to vaporization enthalpy (ΔglH0m,n), the contribution part from the average permanent dipole moment to vaporization enthalpy (ΔglH0m,μ), new scale of polarity (δμ) and polarity coefficient (P) for [Taz(2,4)][Acgly], [Taz(2,4][Acala] and [Taz(2,4][Accys] at T = 298.15 K
ILs |
ΔglH0m/kJ mol−1 |
ΔglH0m,n/kJ mol−1 |
ΔglH0m,μ/kJ mol−1 |
δμ/J1/2·cm−3/2 |
P |
[Taz(2,4)][Acgly] |
167.7 |
89.15 |
78.60 |
18.10 |
0.939 |
[Taz(2,4)][Acala] |
176.3 |
95.72 |
80.55 |
17.65 |
0.917 |
[Taz(2,4)][Accys] |
184.1 |
105.37 |
78.75 |
17.00 |
0.865 |
3.4.2. Improved new scale of the polarity, P. Unfortunately, the polarity of the solvent is the sum of all possible interactions between solvent and solute except chemical action, while the δμ only considers the dipole contribution part and does not consider the non-polar contribution part. Thus, Zhang et al.27–29 took both dipole and non-polar contributions into account, modified δμ and deriving the equation for the polarity coefficient, P: |
P = δμ/δn = [(ΔglH0m,μ/V −( 1 − x)RT/V)/(ΔglH0m,n/V − xRT/V)]1/2
| (14) |
In eqn (14), (ΔglH0m,μ/V) ≫ [(1 – x) RT/V] and (ΔglH0m,n/V) ≫ (xRT/V). Thus, the following equation for calculating the polarity coefficient, P, |
P = (ΔglH0m,μ/ΔglH0m,n)1/2
| (15) |
P is called the polarity coefficient of ionic liquid, and the larger the value of polarity coefficient, the stronger the polarity of ILs (see Table 7).
From Table 7, the trends in P and δμ are identical, indicating that introducing methylene groups or sulfur atoms into the anion structure reduces the hydrophilicity of ILs, with sulfur atoms exhibiting a more pronounced effect. This is attributed to sulfur atoms significantly diminishing hydrophilicity through mechanisms such as charge delocalization, structural asymmetry, and dynamic hydrogen bond modulation-effects that surpass the steric hindrance-dominated contribution of methylene groups. Furthermore, the comparison between [C4mim][BF4] (P = 1.210) and [C4mim][NTf2] (P = 0.647) demonstrates that the polarity of [C4mim][BF4] is substantially higher than that of [C4mim][NTf2], consistent with empirical observations,27 thereby confirming the broader applicability of P as a predictive metric for ILs polarity.
3.5. Viscosity
The temperature dependence of ILs viscosity can be correlated with various theoretical models. In addition to the VFT equation, the Arrhenius equation47 is commonly employed to describe the temperature-dependent viscosity of ILs and calculate the viscous flow activation energy (Eη), which represents the minimum energy required for ion movement within the ILs. Another widely used equation relating viscosity to temperature is: |
ln η = ln η∞ − Eη/RT
| (16) |
where η∞ is an empirical constant, Eη denotes the viscous flow activation energy, R represents the gas constant, and T is the absolute temperature. Using eqn (18), a linear fit of ln
η versus T−1 was performed, with the fitting parameters and calculated results summarized in Table 8 and illustrated in Fig. 5. The Eη follow this order: [Taz(2,4)][Accys] (48.58 kJ mol−1) < [Taz(2,4)][Acala] (51.18 kJ mol−1) < [Taz(2,4)][Acgly] (64.68 kJ mol−1).
Table 8 The values of activation Gibbs energy and activation energy for viscous flow of studied ILs
T/K |
aη/mPa s |
Eη/kJ mol−1 |
104V/m−3 mol−1 |
ΔH≠/kJ mol−1 |
ΔG≠/kJ mol−1 |
TΔS≠/kJ mol−1 |
Derived from previous work.24 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acgly] |
323.15 |
175.0 |
48.58 |
2.398 |
48.00 |
31.06 |
16.93 |
333.15 |
97.8 |
48.58 |
2.410 |
47.89 |
30.43 |
17.46 |
343.15 |
59.2 |
48.58 |
2.424 |
47.91 |
29.93 |
17.98 |
353.15 |
37.6 |
48.58 |
2.437 |
47.99 |
29.48 |
18.51 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Acala] |
323.15 |
245.4 |
51.18 |
2.589 |
50.63 |
32.18 |
18.45 |
333.15 |
137.4 |
51.18 |
2.606 |
50.61 |
31.59 |
19.02 |
343.15 |
82.4 |
51.18 |
2.622 |
50.69 |
31.10 |
19.59 |
353.15 |
48.2 |
51.18 |
2.638 |
50.61 |
30.44 |
20.16 |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
[Taz(2,4)][Accys] |
323.15 |
509.1 |
64.68 |
2.727 |
64.14 |
34.28 |
29.86 |
333.15 |
242.1 |
64.68 |
2.742 |
64.08 |
33.30 |
30.78 |
343.15 |
120.1 |
64.68 |
2.758 |
64.02 |
32.31 |
31.71 |
353.15 |
66.4 |
64.68 |
2.774 |
64.16 |
31.53 |
32.63 |
 |
| Fig. 5 Plot of ln η vs. 1/T for ILs [Taz(2,4)][Accys], [Taz(2,4)][Acala] and [Taz(2,4)][Acgly]. [Taz(2,4)][Acgly]: ln η = −12.94 + 5843.25/T, r2 = 0.999, sd = 0.40962; [Taz(2,4)][Acala]: ln η = −13.54 + 6155.32/T, r2 = 0.999, sd = 0.24198; [Taz(2,4)][Accys]: ln η = −17.86 + 7779.78/T, r2 = 0.999, sd = 0.40962. | |
The Eη reflects the resistance to ion mobility and the energy barrier that ion pairs must overcome for mutual displacement. Higher activation energies indicate greater difficulty in ion pair separation, which may be attributed to factors such as larger ionic sizes, stronger intermolecular interactions, or more complex entanglement between ions in the ILs structure.48 This observation aligns with previous reports correlating activation energy with the strength of cation–anion interactions and ionic liquid microstructural complexity.28
In studies of liquid transport properties, Eyring extended transition state theory to liquid flow phenomena, establishing a viscosity theory. According to the Eyring viscosity equation,49 the relationship between viscosity and the activation Gibbs energy of viscous flow (ΔG≠) is expressed as:
|
η = (hN/V) exp(ΔG≠/RT)
| (17) |
where
η is viscosity,
h is Planck's constant,
N is Avogadro's number,
V is molar volume,
R is the gas constant, and
T is temperature. Substituting experimental data into
eqn (17), the calculated Δ
G≠ values for ILs at different temperatures are summarized in
Table 8 and shown in
Fig. 6. The descending order of Δ
G≠ magnitudes is: [Taz(2,4)][Accys] > [Taz(2,4)][Acala] > [Taz(2,4)][Acgly].
 |
| Fig. 6 Plot of ΔG≠ vs. T for ILs [Taz(2,4)][Accys], [Taz(2,4)][Acala] and [Taz(2,4)][Acgly]. [Taz(2,4)][Acgly]: ΔG≠ = 47.94–0.0524 T, r2 = 0.991, sd = 0.00842; [Taz(2,4)][Acala]: ΔG≠ = 50.64–0.0571 T, r2 = 0.996, sd = 0.00487; [Taz(2,4)][Accys]: ΔG≠ = 64.10–0.0924 T, r2 = 0.996, sd = 0.01242. | |
Using the thermodynamic relationship:
where Δ
H≠ and Δ
S≠ denote the activation enthalpy and entropy of viscous flow, respectively. Substituting
eqn (18) into the Eyring viscosity equation and taking the natural logarithm yields:
|
ln η = (R/ΔS≠) ln (hN/V) − ΔH≠/RT
| (19) |
For moderate temperature ranges, ΔS≠ can be considered constant. Notably, the Eη in the Arrhenius equation corresponds to ΔH≠ in the Eyring framework. This equivalence arises because Eη represents the energy barrier for ion mobility, analogous to the enthalpy required to overcome intermolecular interactions during viscous flow.
As illustrated in Fig. 6, the ΔS≠ for ILs is numerically equivalent to the slope of the linear fitting line. The calculated TΔS≠ values and ΔH≠ derived from eqn (19) at various temperatures are compiled in Table 8. A comparison between ΔH≠ values of the ILs [Taz(2,4)][Accys], [Taz(2,4)][Acala], and [Taz(2,4)][Acgly] and the Arrhenius Eη demonstrates close agreement between these two parameters. Crucially, eqn (19) enables a thermodynamic distinction between enthalpic (ΔH≠) and entropic (TΔS≠) contributions to viscous flow resistance. Table 8 reveals that ΔH≠ > TΔS≠ for all ILs, indicating that the resistance to viscous flow is predominantly governed by enthalpic effects. This implies that ion mobility is primarily determined by the energy barrier ΔH≠ required to overcome intermolecular interactions. Higher ΔH≠ values correlate with stronger constraints on ion pair separation, reflecting increased difficulty for ions to move freely past one another. The dominance of enthalpic contributions aligns with the structural characteristics of ILs, where coulombic forces and hydrogen-bonding networks impose significant energy barriers to ion mobility.
4. Conclusions
This study developed and characterized three novel ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys], investigating their physicochemical properties and structure–property relationships. Quantitative analysis of volumetric properties revealed that specific structural features of the anion systematically influence key molecular parameters. The introduction of each additional –CH2– group resulted in a substantial increase in molecular volume (Vm), a significant elevation in standard molar entropy (S0), and a noticeable reduction in lattice energy (UPOT). Furthermore, the incorporation of a sulfur atom also contributed to measurable increases in both Vm and S0, while concurrently decreasing UPOT, albeit to a less pronounced extent compared to the –CH2– group. Integration of molar surface Gibbs energ with the Lorentz–Lorenz equation yielded a predictive model for surface tension, demonstrating strong agreemen (r2 = 0.99) between calculated and experimental values. Polarity characterization employed a new scale δμ, subsequently refined to the polarity coefficient P. Both metrics confirmed that anion modification via –CH2– or –S– introduction reduces hydrophilicity, with sulfur imparting a more significant hydrophobic enhancement. Viscosity analysis leveraged Eyring's equation and thermodynamic relations to derive activation enthalpy for viscous flow. These results corresponded closely with values from the Arrhenius equation, confirming enthalpic dominance in ion mobility resistance. Systematic modulation of IL properties through anion engineering enables predictive design for applications requiring tailored thermophysical behavior.
Author contributions
Conceptualization, D. F. and J. Q.; methodology, D. F. and L. L.; formal analysis, K. L. and J. M.; investigation, K. L. and H. Y.; resources, K. L. and J. Q.; data curation, K. L.; writing – original draft, K. L. and J. M.; writing – review and editing, K. L. and H. Y.; supervision, J. Q, and L. L.; funding acquisition, L. L., H. Y., and K. L. All authors have read and agreed to the published version of the manuscript.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the supplementary information (SI), Supplementary information: 12 figures showing the 1H NMR, 13C NMR, IR, and TGA spectroscopy of the ILs [Taz(2,4)][Acgly], [Taz(2,4)][Acala] and [Taz(2,4)][Accys]. See DOI: https://doi.org/10.1039/d5ra06386h.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (NSFC) (62201496), Special Funding of the Taishan Scholar Project (tsqn201909150), funding from the Qingchuang Science and Technology Plan of Shandong Universities (2023KJ283) and College Student Innovation and Entrepreneurship Training Program (S202310904130). Youth Entrepreneurship Talents Introduction Project in Shandong Province Higher Education Institutions.
Notes and references
- C. Baudequin, D. Bregeon, J. G. F. Levillain, J. C. Plaquevent and A. C. Gaumont, Chiral ionic liquids, a renewal for the chemistry of chiral solvents? Design, synthesis and applications for chiral recognition and asymmetric synthesis, Tetrahedron Asymmetry, 2005, 16, 3921–3945 CrossRef CAS.
- H. Clavier, L. Boulanger, N. Audi, L. Toupet, M. Mauduit and J. C. Guillemin, Design and synthesis of imidazolinium salts derived from (L)-valine. Investigation of their potential in chiral molecular recognition, Chem. Commun., 2004, 10, 1224–1225 RSC.
- Y. Shimizu, Y. Ohte, Y. Yamamura, K. Saito and T. Atake, Low-temperature heat capacity of room-temperature ionic liquid, 1-hexyl-3-methylimidazolium bis(triflfluoromethylsulfonyl)imide, J. Phys. Chem. B, 2006, 110, 13970–13975 CrossRef CAS PubMed.
- U. P. Yauheni, G. K. Andrey and V. B. Andrey, Calorimetric determination of the enthalpy of 1-butyl-3-methylimidazolium bromide synthesis: a key quantity in thermodynamics of ionic liquids, J. Phys. Chem. B, 2009, 113, 14742–14746 CrossRef PubMed.
- T. Brünig, K. Krekić, C. Bruhn and R. Pietschnig, Calorimetric Studies and Structural Aspects of Ionic Liquids in Designing Sorption Materials for Thermal Energy Storage, Chem.–Eur. J., 2016, 22, 16200–16212 CrossRef PubMed.
- S. Chowdhury, R. S. Mohan and J. L. Scott, Reactivity of ionic liquids, Tetrahedron, 2007, 63, 2363–2389 CrossRef CAS.
- K. Stappert, D. Unal, E. T. Spielberg and A. V. Mudring, Influence of the Counteranion on the Ability of 1-Dodecyl-3-methyltriazolium Ionic Liquids to Form Mesophases, Cryst. Growth Des., 2015, 15, 752–758 CrossRef CAS.
- K. Stappert and A. V. Mudring, Triazolium based ionic liquid crystals: Effect of asymmetric substitution, RSC Adv., 2015, 5, 16886–16896 RSC.
- K. Stappert, D. Unal, B. Mallick and A. V. Mudring, New triazolium based ionic liquid crystals, J. Mater. Chem. C, 2014, 2, 7976–7986 RSC.
- Y. Jeong and J. S. Ryu, Synthesis of 1,3-dialkyl-1,2,3-triazolium ionic liquids and their applications to the Baylis- Hillman reaction, J. Org. Chem., 2010, 75, 4183–4191 CrossRef CAS PubMed.
- D. Chand, M. Wilk-Kozubek, V. Smetana and A. V. Mudring, Alternative to the Popular Imidazolium Ionic Liquids: 1,2,4-Triazolium Ionic Liquids with Enhanced Thermal and Chemical Stability, ACS Sustainable Chem. Eng., 2019, 7, 15995–16006 CrossRef CAS.
- S. Kirchhecker and D. Esposito, Amino acid based ionic liquids: A green and sustainable perspective, Curr. Opin. Green Sustain. Chem., 2016, 2, 28–33 CrossRef.
- H. Ohno and K. Fukumoto, Amino Acid Ionic Liquids, Acc. Chem. Res., 2007, 40, 1122–1129 CrossRef CAS PubMed.
- K. Fukumoto, M. Yoshizawa and H. Ohno, Room Temperature Ionic Liquids from 20 Natural Amino Acids, J. Am. Chem. Soc., 2005, 127, 2398–2399 CrossRef CAS PubMed.
- O. J. Curnow and R. Yunis, Synthesis, characterization and properties of amino acid ionic liquids derived from the triaminocyclopropenium cation, RSC Adv., 2016, 6, 70152–70164 RSC.
- M. P. Pereira, R. D. Souza Martins, M. A. L. De Oliveira and F. I. Bombonato, Amino acid ionic liquids as catalysts in a solvent-free Morita-Baylis-Hillman reaction Bombonato, RSC Adv., 2018, 8, 23903–23913 RSC.
- S. Bhattacharyya and F. U. Shah, Thermal stability of choline based amino acid ionic liquids, J. Mol. Liq., 2018, 266, 597–602 CrossRef CAS.
- A. R. Jesusa, M. R. C. Soromenhoa, L. R. Raposob, J. M. S. S. Esperançaa, P. V. Baptistab, A. R. Fernandesb and P. M. Reisa, Enhancement of water solubility of poorly water-soluble drugs by new biocompatible N-acetyl amino acid N-alkyl cholinium-based ionic liquids, Eur. J. Pharm. Biopharm., 2019, 137, 227–232 CrossRef PubMed.
- D. W. Fang, W. Guan, J. Tong, Z. W. Wang and J. Z. Yang, Study on physicochemical properties of ionic liquids based on alanine, J. Phys. Chem. B, 2008, 112, 7499–7505 CrossRef CAS PubMed.
- D. W. Fang, J. Tong, W. Guan, H. Wang and J. Z. Yang, Predicting Properties of Amino Acid Ionic Liquid Homologue of 1-Alkyl-3-methylimidazolium Glycine, J. Phys. Chem. B, 2010, 114, 13808–13814 CrossRef CAS PubMed.
- J. Z. Yang, Q. G. Zhang, B. Wang and J. Tong, Study on the Properties of Amino Acid Ionic Liquid EMIGly, J. Phys. Chem. B, 2006, 110, 22521–22524 CrossRef CAS PubMed.
- Q. Yan, M. Liu, C. Y. Xiao, D. L. Fu, J. Wei, D. W. Fang and J. Z. Yang, Predicting properties of ionic liquid homologue of N-alkylpyridinium acetate, J. Mol. Liq., 2021, 324, 114720 CrossRef CAS.
- Q. Yan, J. Wei, J. Liu, Z. H. Zhang and D. W. Fang, Physicochemical properties of hydrophobic hexafluoroantimonate ionic liquids and applications of the mole surface Gibbs free energy, Ionics, 2020, 26, 5585–5595 CAS.
- K. H. Liang, Z. W. Lu, C. X. Ren, J. Wei and D. W. Fang, Feasibility of 1-Ethyl-4-butyl-1,2,4-triazolium Acetyl Amino Acid Ionic Liquids as Sustainable Heat-Transfer Fluids, ACS Sustainable Chem. Eng., 2022, 10, 3417–3429 CAS.
- K. H. Liang, H. Y. Yao, J. Qiao, S. Gao, M. J. Zong, F. S. Liu, Q. L. Yang, L. J. Liang and D. w. Fang, Thermodynamic Evaluation of Novel 1,2,4-Triazolium Alanine Ionic Liquids as Sustainable Heat-Transfer Media, Molecules, 2024, 29, 5227 CAS.
- D. L. Fu, Z. R. Song, J. K. Liu, J. Yang, S. L. Suo, K. H. Liang, X. X. Ma and D. W. Fang, Low-temperature heat capacity and the thermodynamic functions of a novel ether-based ionic liquid 1-(2-ethoxyethyl)-3-ethylimidazolium thiocyanate, RSC Adv., 2024, 14, 34971 RSC.
- D. Zhang, W. Jiang, L. Liu, K. Yu, M. Hong and J. Tong, The molar surface Gibbs energy and polarity of ether-functionalized ionic liquids, J. Chem. Thermodyn., 2019, 138, 313–320 CrossRef CAS.
- D. Zhang, B. Li, M. Hong, Y. X. Kong, J. Tong and W. G. Xu, Synthesis and characterization of physicochemical properties of new ether-functionalized amino acid ionic liquids, J. Mol. Liq., 2020, 304, 112718 CrossRef CAS.
- D. Zhang, Y. Qu, Y. Y. Gong, J. Tong, D. W. Fang and J. Tong, Physicochemical properties of [cnmim][thr] (n= 3,5,6) amino acid ionic liquidss, J. Chem. Thermodyn., 2017, 115, 47–51 CrossRef CAS.
- H. F. D. Almeida, J. A. Lopes-da-Silva, M. G. Freire and J. A. P. Coutinho, Surface tension and refractive index of pure and water-saturated tetradecyltrihexylphosphonium-based ionic liquids, J. Chem. Thermodyn., 2013, 57, 372–379 CrossRef CAS.
- M. Deetlefs, K. R. Seddon and M. Shara, Predicting physical properties of ionic liquids, Phys. Chem. Chem. Phys., 2006, 8, 642–649 RSC.
- B. W. Mountain and T. M. Seward, Hydrosulfide/sulfide complexes of copper(I): Experimental confirmation of the stoichiometry and stability of Cu(HS)2-to elevated temperatures, Geochim. Cosmochim. Acta, 2003, 67, 3005–3014 CrossRef CAS.
- C. L. Wei, K. Jiang, T. M. Fang and X. M. Liu, Effects of Anions and Alkyl Chain Length of Imidazolium-Based Ionic Liquids at the Au(111) Surface on Interfacial Structure: A First Principles Study, Green Chem. Eng., 2021, 2, 402–411 CrossRef.
- A. Bhattacharjee, A. Luís, J. H. Santos, J. A. Lopes-da-Silva, M. G. Freire, P. J. Carvalho and J. A. P. Coutinho, Thermophysical properties of sulfonium- and ammonium-based ionic liquids, Fluid Phase Equilib., 2014, 381, 36–45 CrossRef CAS PubMed.
- Q. Wang, Z. G. Qu and D. Tian, Electric Double Layer Theory of Interfacial Ionic Liquids for Capturing Ion Hierarchical Aggregation and Anisotropic Dynamics, Adv. Energy Mater., 2025, 15, 2402974 CrossRef CAS.
- W. D. Tian, C. K. Zhang, S. Paul, W. D. Si, Z. Wang, P. P. Sun, A. Anoop, C. H. Tung and D. Sun, Lattice Modulation on Singlet-Triplet Splitting of Silver Cluster Boosting Near-Unity Photoluminescence Quantum Yield, Angew. Chem., Int. Ed., 2025, 64, e202421656 CrossRef CAS PubMed.
- S. Kang, Y. G. Chung, J. H. Kang and H. Song, CO2 absorption characteristics of amino group functionalized imidazolium-based amino acid ionic liquids, J. Mol. Liq., 2020, 297, 111825 CrossRef CAS.
- J. Tong, M. Hong, Y. Chen, H. Wang, W. Guan and J. Z. Yang, The surface tension, density and refractive index of amino acid ionic liquids: [C3mim][Gly] and [C4mim][Gly], J. Chem. Thermodyn., 2012, 54, 352–357 CrossRef CAS.
- C. B. Zhou, J. Li, Z. Yi and H. J. Ai, Refractive Properties of Imidazolium Ionic Liquids with Alanine Anion [Cnmim][Ala] (n=2,3,4,5,6), Russ. J. Phys. Chem. A, 2017, 91, 2044–2051 CrossRef CAS.
- D. Zhang, W. Jiang, L. Liu, K. Yu, M. Hong and J. Tong, The molar surface Gibbs energy and polarity of ether-functionalized ionic liquids, J. Chem. Thermodyn., 2019, 138, 313–320 CrossRef CAS.
- A. W. Adamson, Physical Chemistry of Surfaces, Wiley, New York, 1976, 3rd edn, 698 Search PubMed.
- Y. Meng, J. Liu, Z. Li and H. M. Wei, Synthesis and physicochemical properties of two SO3H functionalized ionic liquids with hydrogen sulfate anion, J. Chem. Eng. Data, 2014, 59, 2186–2195 CrossRef CAS.
- J. Tong, H. X. Yang, R. J. Liu, C. Li, L. X. Xia and J. Z. Yang, Determination of the Enthalpy of Vaporization and Prediction of Surface Tension for Ionic Liquid 1-Alkyl-3-methylimidazolium Propionate [Cnmim][Pro](n=4,5,6), Phys. Chem. B, 2014, 118, 12972–12978 CrossRef CAS PubMed.
- J. H. Hildebrand and R. L. Scott, The Solubility of Nonelectrolytes, Reinhold, New York, 1950, 3rd edn Search PubMed.
- A. A. Strechan, G. J. Kabo and Y. U. Paulechka, The correlations of the enthalpy of vaporization and the surface tension of molecular liquids, Fluid Phase Equilib., 2006, 250, 125–130 CrossRef CAS.
- D. D. Lawson and J. D. Ingham, Estimation of Solubility Parameters from Refractive Index Data, Nature, 1969, 223, 614 CrossRef CAS.
- J. Vila, P. Ginés, J. M. Pico, C. Franjo, E. Jiménez, L. M. Varela and O. Cabeza, Temperature dependence of the electrical conductivity in EMIM-based ionic liquids-Evidence of Vogel-Tamman-Fulcher behavior, Fluid Phase Equilib., 2006, 242, 141–146 CrossRef CAS.
- M. Ebrahimi and F. Moosavi, The effects of temperature, alkyl chain length, and anion type on thermophysical properties of the imidazolium based amino acid ILs, J. Mol. Liq., 2018, 250, 121–130 CrossRef CAS.
- H. Eyring, Viscosity, plasticity, and diffusion as examples of absolute reaction rates, J. Chem. Phys., 1936, 4, 283–291 CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.