Next Article in Journal
Functionalized Nanomembranes and Plasma Technologies for Produced Water Treatment: A Review
Next Article in Special Issue
Chiral Porous Carbon Surfaces for Enantiospecific Synthesis
Previous Article in Journal
Calcined Co(II)-Chelated Polyazomethine as Cathode Catalyst of Anion Exchange Membrane Fuel Cells
Previous Article in Special Issue
Pure Chitosan-Based Fibers Manufactured by a Wet Spinning Lab-Scale Process Using Ionic Liquids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physical Properties of Betaine-1,2-Propanediol-Based Deep Eutectic Solvents

School of Energy and Power Engineering, Northeast Electric Power University, Jilin City 132012, China
*
Author to whom correspondence should be addressed.
Submission received: 16 March 2022 / Revised: 12 April 2022 / Accepted: 24 April 2022 / Published: 27 April 2022
(This article belongs to the Special Issue Applications of Ionic Liquids in Colloid and Polymer Chemistry)

Abstract

:
Due to their splendid advantages, deep eutectic solvents have attracted high attention and are considered as analogues of ionic liquids. Deep eutectic solvents (DESs) are homogeneous mixtures formed by two or three green and cheap components through hydrogen bond, which is divided into hydrogen bond acceptors (HBA) and hydrogen bond donors (HBD). Recently, Betaine has been widely used as a hydrogen bond acceptor. In this work, four DESs were synthesized by blending betaine as HBA and 1,2-propanediol as HBD in four molar ratios (1:3.5, 1:4, 1:5, 1:6). Then, the physical properties of these DESs were measured. The density values were measured within the temperature range (293.15 K to 363.15 K) at atmospheric pressure, whereas the surface tension and viscosity data were determined in four and seven temperatures between 293.15 K and 353.15 K. The relationship between the density and surface tension with temperature have been analyzed and have been fitted as a linear function. The commonly used Arrhenius model was used to describe the dependence between viscosity and temperature. The results of this study are important not only for the DESs’ industrial applications but also for the research on their synthesis mechanism and microstructure.

1. Introduction

With the introduction of the concept of green chemistry, many scholars have explored ionic liquids more and more deeply. Due to their remarkable properties (low vapor pressure, high boiling points, low flammability, excellent dissolution ability, thermal stability and chemical stability, wide liquid range, and so on), ionic liquids are treated as a green solvent and have been used in the fields of separation technology, biocatalysis, organic synthesis, and electrochemistry. Currently, there are about 250 ionic liquids that have been commercialized. However, it has been reported that some defects, such as complex synthesis, purification difficulties, high price, and poor biodegradability, limit the large-scale industrial application of ionic liquids. Therefore, it is necessary to develop green solvents that can maintain the excellent physical properties of ionic liquids while overcoming their disadvantages [1,2,3,4].
With the improvement of solvent performance requirements, deep eutectic solvents (DESs) are recognized as ionic liquid analogues, have attracted high attention, and have become another novel class of green solvent [5,6,7,8,9]. Since Abbott’s team discovered and named the deep eutectic solvent in 2003 [10], DESs have been extensively studied. It has been found that DESs can not only overcome the shortcomings of organic solvents, such as volatility and high melting point of ionic liquids, but also maintain the advantages of these two types of reagents. DESs are widely used in many fields such as electrochemistry, separation technology, organic synthesis, and so on because of their simple preparation, low price, non-toxicity, low vapor pressure, and wide electrochemical window. It is confirmed that DESs have a very broad application prospect [11,12,13,14,15].
DESs are homogeneous mixtures that are made by blending two or three hydrogen bond donors (HBD) and one hydrogen bond acceptor (HBA) with the melting point lower than each of the components. Ammonium salt, amino acids, and metal chloride are usually used as HBA, whereas amines, amides, carboxylic acids, alcohols, and amino acids are employed as HBD. In the early studies, cholinium chloride (choline) was the most commonly used compound to prepare DESs [10,11,16]. To expand the application of DESs, it is necessary to research new HBAs, which can be substituted for choline chloride, that are cheaper, more readily available, and non-toxic. In this context, betaine has gradually entered the scope of research scholars. Betaine is a kind of inner salt consisting of an anionic carboxylic acid group and a positively charged quaternary nitrogen [17]. In theory, betaine cannot be used to prepare DES because the betaine molecule itself should establish a strong anion-cation interaction. After all, the anion-cation interaction is much stronger than the hydrogen bond formed by HBA and HBD. However, in practice, the positively charged quaternary nitrogen in betaine is highly shielded by methyl groups, thereby inhibiting the electrostatic interaction with anions [18]. Therefore, betaine has been considered a suitable HBA for the synthesis of DES in recent years due to its own structural characteristics and its sustainability.
The betaine-based DESs are being used in the fields of extraction, separation, CO2 adsorption, catalysts, nanomaterials, and so on. For example, Li et al. [19] synthesized six kinds of betaine-based DESs and successfully applied them in the extraction of protein from calf blood, finding that salt concentration played an important role in the optimal extraction conditions. Fanali et al. [20] reported that nutraceutical compounds were extracted from coffee grounds more efficiently using betaine-based DESs than the DESs using choline chloride as HBD. Additionally, three deep eutectic solvents based on betaine, glycerol, ethylene glycol, and propylene glycol were prepared by Kučan’s research group and successfully used as selective solvents for extracting only nitrogen compounds from gasoline [21]. Mojtabavi et al. [22] used a betaine-based DES to increase the enzyme stability at various temperatures and pH levels as an appropriate alternative strategy to improve the Laccase stability. Ribeiro et al. [23] found that DESs formed from betaine hydrochloride, organic acids, polyols, and amides were an alternative as non-aqueous media for enzymatic reactions of lipases. Additionally, lipase showed great thermostability and relative activity in the presence of betaine hydrochloride and urea (molar ratio 1:4) relative to aqueous solution. Moreover, betaine-based DES has a strong ability to absorb carbon dioxide [24], and another report by Crescenzo’s group showed that gold nanoparticles could be synthesized from betaine-based DESs without any surfactant or capping agent [25]. Although a growing number of researchers have reported the applications of some betaine-based DESs, the physical properties of betaine-based DESs are still poorly studied. However, especially in the design of the optimal system for a given application, it is important to strengthen the exploration of the physical properties of DESs, such as density, surface tension, and viscosity, among others [26,27,28,29,30].
Within this background, in this work, four DESs mixed using betaine as the HBA and 1,2-propanediol as the HBD in different molar ratios have been synthesized, and their physical properties were measured, including density (from 293.15 K to 363.15 K), surface tension (from 293.15 K to 353.15 K), and viscosity (from 293.15 K to 353.15 K). The changes in density, surface tension, and viscosity of DESs in relation to temperature were obtained, and the corresponding dependencies were fitted. Meanwhile, the effects of different molar ratios on the density, surface tension, and viscosity of DESs were summarized.

2. Materials and Methods

2.1. Materials

Betaine (CAS: 107-43-7) and 1,2-propanediol (CAS: 57-55-6) were purchased from Aladdin Chemistry Co. Ltd. (Shanghai, China) and were used directly without further purification with the mass fraction purity better than 99% from supplier (Table 1).

2.2. Preparation of DESs

Four DESs were composed in the four molar ratios (1:3.5, 1:4, 1:5, 1:6) using betaine and 1,2-propanediol (as shown in Table 2). The weighed samples were shaken by an incubator at 300 rpm and 353.15 K for 2 h until a colorless liquid was obtained. Four DESs were prepared at atmospheric pressure and under tight control of moisture content. Then, the synthesized samples were put in a desiccator for at least 24 h. In addition, we tried to synthesize another two DESs with the molar ratios of betaine to 1,2-propanediol being 1:2 and 1:3. However, it failed to compose the homogeneous mixture in these ratios at atmospheric pressure.

2.3. Physical Properties Measurement

An Anton Paar digital densimeter (Model: DMA 5000M), which is based on the vibrating U-tube method, was used to measure the DESs’ density. The densimeter can be used to conduct density measurements in temperatures ranging from 273.15 K to 363.15 K, and its repeatability was 0.000001 g/cm³ from supplier. The density measurement’s relative standard uncertainty was specified as 0.001. Additionally, the densimeter must be well-cleaned using acetone or ethanol and dried before and after the measurements [31].
The measurement of surface tension was conducted by the experimental system that has been established in our laboratory based on the two capillaries rise method. The capillary rise method, which is one of the most accurate methods to determine the surface tension, not only has a relatively complete theory but also can strictly control the experimental conditions, and its principle has been introduced in details elsewhere [32,33]. In this work, the uncertainty of surface tension measurement was better than ±0.2 mN·m−1.
The viscosity of DESs were measured by an accurate viscosity measurement system that was composed of a Brookfield rheometer (Model: RST-CC), a Brookfield water thermostatic (Model: TC-550SD), and other accessories. Rheological evaluation through controlled stress and controlled rate measurements offers superior viscosity profiling, thixotropic response, yield stress determination, and creep analysis. The measurement system can be used to measure the viscosity in temperatures ranging from 253.15 K to 423.15 K with an uncertainty of ±0.04 K.

3. Results and Discussion

3.1. Density

Density is a key physical property of materials and has an impact on mass transfer or chemical processes, and it also can provide information about the intermolecular interactions in DESs. Generally, the densities of DESs reported in the literature are higher than that of water. In this work, four prepared DESs were measured in the temperature range (293.15–363.15) K. During the experimental measurement, the density in the same temperature was measured in triplicate to obtain the average data of the density. Table 3 gives the density of DESs and pure 1,2-propanediol in different temperatures. It should be noted that the density of the DES decreases with increasing the molar ratio of 1,2-propanediol. Densities of four DESs decreased as follows: DES1 > DES2 > DES3 > DES4, ranging from 1.075588 to 1.062906 kg·m3 at 293.15 K and from 1.02875 to 1.013525 kg·m3 at 363.15 K. In addition, the density of all DESs is higher than that of the pure 1,2-propanediol (for example, 1.0362 kg·m3 at 293.15 K) at the same temperature. As most betaine/polyol-based DESs are denser than water, as described by Zhang [12]. Such measurements were consistent with the density changes of DESs synthesized using betaine and levulinic acid in different ratios reported by Mulia [22], and the addition of salt increases the density of the isolated HBD. However, the density of the mixture of betaine and glycerol measured by Rodrigues was less than that of the isolated glycerol [34]. It is noted that in some cases, selecting the appropriate composition and molar ratio of HBA to HBD provides a method for changing the density of the eutectic mixture. This phenomenon may be explained in terms of free volume [35]. Figure 1 displays the relationship between the DESs’ density and temperature. Clearly, the density of all DESs linearly decreases with temperature rise. This is due to the higher molecular activity and the enhanced mobility of the molecules as the temperature increases, thereby increasing the molar volume of the solution and ultimately leading to a decrease in density [36].
The experimental data of the density were fitted by a linear function of temperature:
ρ = A + BT,
where ρ is the density, T is the temperature, and A and B are two constants. The values of A and B of different DESs are presented in Table 4.

3.2. Surface Tension

Surface tension is a useful physical property which is needed in a lot of areas, such as emulsions and surfactants applications. Measuring the surface tension of DES can be used as a method to understand the changes of DES molecular environment caused by the changes of composition and temperature. The vast majority of DES surface tension measurements reported so far are based on choline chloride and have found that the surface tension of choline-based DES decreases with the increase in temperature [29,37]. In addition, Gajardo-Parra [38] measured the surface tension of DES composed of choline chloride as HBA and ethylene glycol, phenol, and levulinic acid as HBD at 298.15 K and 101.3 kPa. The reported surface tension of choline chloride and ethylene glycol (45.66 mNm1 at 298.15 K and 101.3 kPa) was lower than that of the pure ethylene glycol (48.90 mNm1 measured). Very few studies reported the surface tension of betaine-based DESs. In this study, the measurement of surface tension of the studied betaine-based DESs in temperatures ranging from 293.15 K to 353.15 K was conducted. Table 5 lists the experimental surface tension results of solvents, and they are also higher than the pure 1,2-propanediol. Figure 2 and Figure 3 show these data graphically. It can be seen clearly that the surface tension of the four prepared DESs decreases as the temperature increases, which is consistent with the variation trend of the surface tension of choline-based DES. It is worth noting that the surface tensions of betaine-based DESs measured in this work are all higher than those of the pure 12-propanediol at the same temperature. In this study, a reproducible experiment was carried out in order to obtain the measurement results, indicating that the measurement results in this work have certain credibility. It can be concluded that the surface tension of betaine-based DES is higher than that of pure substance, whereas choline chloride-based DES has a lower surface tension than pure substance. These findings might depend on the intermolecular forces and hydrogen bonding between HBA and HBD; thus, further studies need to be conducted to investigate these from a molecular level.
From Figure 3 it can be seen that the surface tension of the four DESs first decreased and then increased from DES1 to DES4. DES1 has the highest surface tension, whereas DES3 has the lowest surface tension at the same temperature. The reason for this is that the HBD and HBA mixed at the appropriate molar ratio can generate the optimum hydrogen bonding interaction force, and the stronger the hydrogen bonding force appears, the smaller the DESs’ surface tension is [39].
The experimental results of the density and surface tension were fitted as a linear function of temperature as follows:
σ = a + bT,
where σ is the surface tension, T is the temperature, and a and b are two constants. The values of a and b of different DESs are presented in Table 6.

3.3. Viscosity

Viscosity is a physical quantity that measures the amount of frictional resistance between adjacent fluid layers when a fluid flows. It is well known that DESs generally have a higher viscosity, similar to most of the ionic liquids. The high viscosity of DES has brought inconvenience to its practical application and is regarded as the main obstacle to its widespread use. Therefore, viscosity is another important property of solvents that must be addressed. In this work, the viscosity of prepared DESs and pure 1,2-propanediol in the temperature range (293.15–353.15 K) was measured, and the experimental results are listed in Table 7. Figure 4 shows the trend of the viscosity with increasing temperature. It can be seen that the viscosity of DESs is greatly influenced by the temperature. The viscosity decreased significantly with increasing temperature, especially in the low temperature range. In general, the increase in temperature leads to the weakening of the intermolecular force in the liquid and the acceleration of the average movement speed of the molecules, thereby improving the kinetic energy of the molecules, promoting the flow between the molecules, and resulting in the decrease in dynamic viscosity of the liquid.
Figure 4 also shows that for a given temperature, the viscosity of the four prepared DESs are in the following descending order: DES1 > DES2 > DES3 > DES4. Although the viscosity of DES decreased with increasing molar ratio of 12-propanediol, the viscosity of all DESs increased compared with that of the isolated 1,2-propanediol. The main reasons are that, on the one hand, there are a large number of hydrogen bond networks between each component of HBA and HBD, which leads to the lower mobility of free species within DES [12]; on the other hand, it is due to the thickening potential of betaine. Consequently, in some practical applications that require lower viscosity, selecting the appropriate ratio of HBA to HBD or preheating and increasing the temperature are very simple and effective techniques to reduce viscosity.
The Arrhenius model and Vogel-Fulcher-Tammann (VFT) behavior are most commonly used to describe the temperature dependence of DES viscosity [40]. Therefore, the Arrhenius model is considered to be more accurate and simple. The measured viscosities could be described by the Arrhenius model over the studied temperature range as shown:
ln η = ln η 0 + E η R T ,
where η stands for the viscosity, η0 is a constant, Eη refers to the activation energy, R is the gas constant, and T is the temperature. Figure 5 indicates that the lnηT−1 relationship tends to be linear. According to the linear fitting, the values of η0 and Eη are given in Table 8.
Then we calculated the viscosity of DESs according to the Arrhenius model at different temperatures and compared them with the experimental data. The deviation values were also calculated, as listed in Table 9. It can be seen that the deviation is significant between the calculated and experimental data. The Arrhenius model is famous for expressing the viscous behavior of DESs, but it is more suitable for predicting the viscosity of liquids at high temperatures or for measuring viscosity over a narrow temperature range [40]. The temperature range (293.15–353.15 K) we used to measure the viscosity of DESs is a little large, resulting in a large deviation between the experimental values and the calculated values.

4. Conclusions

DESs have been considered to be green solvents due to their remarkable solubilizing power and have been recommended for industrial applications. In this study, four DESs were successfully prepared by using betaine as the HBA and 1,2-propanediol as the HBD in different molar ratios. The physical properties (density, surface tension, and viscosity) of these four DESs were determined and reported in the large temperature range. The results show that the density, surface tension, and viscosity of these DESs decrease with the temperature increasing and are higher than the isolated 1,2-propanediol. The density and viscosity of the DESs showed a linear relationship with temperature, whereas the dependence of viscosity and temperature can be obtained using the Arrhenius model. In addition, with the increase in the molar ratio of 1,2-propanediol, the density and viscosity of DES decreased, and the surface tension showed a trend of first decreasing and then increasing. The molar ratio 1:5 of betaine and 1,2-propanediol had the lowest surface tension. It is believed that the results in this study would be important for utilization of DESs in practical industrial applications.
Although some problems need to be further studied, such as the issue that homogeneous mixtures could not be formed for HBA: HBD ratios 1:2 and 1:3, the effect of betaine on the surface tension of DESs is unclear. Molecular dynamic simulation provides a useful tool to investigate the molecular interaction between different HBA and HBD. Additionally, molecular perspective investigations conducted by MD simulations might be helpful to understand the formation mechanism of DESs and the mechanism behind the optimal hydrogen bonding.

Author Contributions

Investigation, N.H.; writing—original draft preparation, J.F.; writing—review and editing, F.S.; supervision, Q.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Jilin Province (China), grant number 20200201223JC.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the financial support by the Natural Science Foundation of Jilin Province (China) with the project number is 20200201223JC.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Chernikova, E.A.; Glukhov, L.M.; Krasovskiy, V.G.; Kustov, L.M.; Vorobyeva, M.G.; Koroteev, A.A. Ionic liquids as heat transfer fluids: Comparison with known systems, possible applications, advantages and disadvantages. Russ. Chem. Rev. 2015, 84, 875–890. [Google Scholar] [CrossRef]
  2. Minea, A.A. Overview of ionic liquids as candidates for new heat transfer fluids. Int. J. Thermophys. 2020, 41, 151. [Google Scholar] [CrossRef]
  3. Buzzeo, M.C.; Evans, R.G.; Compton, R.G. Non-haloaluminate room-temperature ionic liquids in electrochemistry-A review. Chemphyschem 2004, 5, 1106–1120. [Google Scholar] [CrossRef] [PubMed]
  4. Krishnan, A.; Gopinath, K.P.; Vo, D.N.; Malolan, R.; Nagarajan, V.M.; Arun, J. Ionic liquids, deep eutectic solvents and liquid polymers as green solvents in carbon capture technologies: A review. Environ. Chem. Lett. 2020, 18, 2031–2054. [Google Scholar] [CrossRef]
  5. Plechkova, N.V.; Seddon, K.R. Applications of ionic liquids in the chemical industry. Chem. Soc. Rev. 2008, 37, 123–150. [Google Scholar] [CrossRef]
  6. Fernandez, M.D.; Boiteux, J.; Espino, M.; Gomez, F.J.; Silva, M.F. Natural deep eutectic solvents-mediated extractions: The way forward for sustainable analytical developments. Anal. Chim. Acta 2018, 1038, 1–10. [Google Scholar] [CrossRef]
  7. Alonso, D.A.; Baeza, A.; Chinchilla, R.; Guillena, G.; Pastor, I.M.; Ramon, D.J. Deep eutectic solvents: The organic reaction medium of the century. Eur. J. Org. Chem. 2016, 2016, 612–632. [Google Scholar] [CrossRef] [Green Version]
  8. Mjalli, F.S.; Mousa, H. Viscosity of aqueous ionic liquids analogues as a function of water content and temperature. Chin. J. Chem. Eng. 2017, 25, 1877–1883. [Google Scholar] [CrossRef]
  9. Yan, Y.C.; Rashmi, W.; Khalid, M.; Shahbaz, K.; Gupta, T.C.; Mase, N. Potential application of deep eutectic solvents in heat transfer application. J. Eng. Sci. Technol. 2017, 12, 1–14. [Google Scholar]
  10. Abbott, A.P.; Capper, G.; Davies, D.L.; Rasheed, R.K.; Tambyrajah, V. Novel solvent properties of choline chloride/urea mixtures. Chem. Commun. 2003, 39, 70–71. [Google Scholar] [CrossRef] [Green Version]
  11. Abbott, A.P.; Capper, G.; Davies, D.L.; McKenzie, K.J.; Obi, S.U. Solubility of metal oxides in deep eutectic solvents based on choline chloride. J. Chem. Eng. Data 2006, 51, 1280–1282. [Google Scholar] [CrossRef]
  12. Zhang, Q.H.; De Oliveira Vigier, K.; Royer, S.; Jerome, F. Deep eutectic solvents: Syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 7108–7146. [Google Scholar] [CrossRef] [PubMed]
  13. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep eutectic solvents (DESs) and their applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Tran, K.T.; Le, L.T.; Phan, A.L.; Tran, P.H.; Vo, T.D.; Truong, T.T.; Nguyen, N.T.; Garg, A.; Le, P.M.; Tran, M.V. New deep eutectic solvents based on ethylene glycol—LiTFSI and their application as an electrolyte in electrochemical double layer capacitor (EDLC). J. Mol. Liq. 2020, 320, 114495. [Google Scholar] [CrossRef]
  15. Rahman, M.S.; Roy, R.; Jadhav, B.; Hossain, M.N.; Halim, M.A.; Raynie, D.E. Formulation, structure, and applications of therapeutic and amino acid-based deep eutectic solvents: An overview. J. Mol. Liq. 2021, 321, 114745. [Google Scholar] [CrossRef]
  16. Abbott, A.P.; Boothby, D.; Capper, G.; Davies, D.L.; Rasheed, R.K. Deep eutectic solvents formed between choline chloride and carboxylic acids: Versatile alternatives to ionic liquids. J. Am. Chem. Soc. 2004, 126, 9142–9147. [Google Scholar] [CrossRef] [PubMed]
  17. Viertorinne, M.; Valkonen, J.; Pitknäen, I.; Mathlouthi, M.; Nurmi, J. Crystal and molecular structure of anhydrous betaine, (CH3)3NCH2CO2. J. Mol. Struct. 1999, 477, 23–29. [Google Scholar] [CrossRef]
  18. Abranches, D.O.; Silva, L.P.; Martins, M.A.; Pinho, S.P.; Coutinho, J.A. Understanding the formation of deep eutectic solvents: Betaine as a universal hydrogen bond acceptor. Chemsuschem 2020, 13, 4916–4921. [Google Scholar] [CrossRef]
  19. Li, N.; Wang, Y.Z.; Xu, K.J.; Huang, Y.H.; Wen, Q.; Ding, X.Q. Development of green betaine-based deep eutectic solvent aqueous two-phase system for the extraction of protein. Talanta 2016, 152, 23–32. [Google Scholar] [CrossRef]
  20. Fanali, C.; Della, P.S.; Dugo, L.; Gentili, A.; Mondello, L.; Gara, L.D. Choline-chloride and betaine-based deep eutectic solvents for green extraction of nutraceutical compounds from spent coffee ground. J. Pharmaceut. Biomed. 2020, 189, 113421. [Google Scholar] [CrossRef]
  21. Kučan, K.Z.; Perković, M.; Cmrk, K.; Nacinovic, D.; Rogosic, M. Betaine plus (glycerol or ethylene glycol or propylene glycol) deep eutectic solvents for extractive purification of gasoline. Chemistryselect 2018, 3, 12582–12590. [Google Scholar] [CrossRef]
  22. Mojtabavi, S.; Jafari, M.; Samadi, N.; Mehrnejad, F.; Faramarzi, M.A. Insights into the Molecular-Level details of betaine interactions with Laccase under various thermal conditions. J. Mol. Liq. 2021, 339, 116832. [Google Scholar] [CrossRef]
  23. Ribeiro, B.D.; de Carvalho Iff, L.; Coelho, M.A.Z.; Marrucho, I.M. Influence of betaine- and choline-based eutectic solvents on lipase activity. Curr. Biochem. Eng. 2019, 5, 57–68. [Google Scholar] [CrossRef]
  24. Mulia, K.; Krisanti, E.; Nasruddin; Libriandy, E. Betaine-based deep eutectic solvents with diol, acid and amine hydrogen bond donors for carbon dioxide absorption. In Proceedings of the 3rd International Conference of Chemical and Materials Engineering, Kanoa Resort Saipan, America, Banda Aceh, Indonesia, 20–21 September 2017. [Google Scholar]
  25. Crescenzo, A.D.; Tiecco, M.; Zappacosta, R.; Boncompagni, S.; Profio, P.D.; Ettorre, V.; Fontana, A.; Germani, R.; Siani, G. Novel zwitterionic natural deep eutectic solvents as environmentally friendly media for spontaneous self-assembly of gold nanoparticles. J. Mol. Liq. 2018, 268, 371–375. [Google Scholar] [CrossRef]
  26. Craveiro, R.; Aroso, I.; Flammia, V.; Carvalho, T.; Viciosa, M.T.; Dionisio, M.; Barreiros, S.; Reis, R.L.; Duarte, A.R.; Paiva, A. Properties and thermal behavior of natural deep eutectic solvents. J. Mol. Liq. 2016, 215, 534–540. [Google Scholar] [CrossRef]
  27. Sánchez, P.B.; González, B.; Salgado, J.; Parajó, J.J.; Domínguez, Á. Physical properties of seven deep eutectic solvents based on L-proline or betaine. J. Chem. Thermodyn. 2019, 131, 517–523. [Google Scholar] [CrossRef]
  28. Hayyan, A.; Mjalli, F.S.; AlNashef, I.M.; Al-Wahaibi, Y.M.; Al-Wahaibi, T.; Hashim, M.A. Glucose-based deep eutectic solvents: Physical properties. J. Mol. Liq. 2013, 178, 137–141. [Google Scholar] [CrossRef]
  29. AlOmar, M.K.; Hayyan, M.; Alsaadi, M.A.; Akib, S.; Hayyan, A.; Hashim, M.A. Glycerol-based deep eutectic solvents: Physical properties. J. Mol. Liq. 2016, 215, 98–103. [Google Scholar] [CrossRef]
  30. Chen, Y.; Fu, L.; Liu, Z.H.; Dai, F.C.; Dong, Z.K.; Li, D.; Liu, H.; Zhao, D.; Lou, Y.Y. Surface tension and surface thermodynamic properties of PEG-based deep eutectic solvents. J. Mol. Liq. 2020, 318, 114042. [Google Scholar] [CrossRef]
  31. Wang, X.; Wang, X.; Chen, J. Experimental investigations of density and dynamic viscosity of n-hexadecane with three fatty acid methyl esters. Fuel 2016, 166, 553–559. [Google Scholar] [CrossRef]
  32. Fan, J.; Zhao, X.; Guo, Z. Surface tension of ethyl fluoride (HFC161) from (233 to 373)K. Fluid. Phase. Equilibr. 2012, 316, 98–101. [Google Scholar] [CrossRef]
  33. Gao, W.; Zhao, X.; Liu, Z. Surface tension of 2,2-Dimethylbutane from (233 to 378) K. J. Chem. Eng. Data 2009, 54, 1761–1763. [Google Scholar] [CrossRef]
  34. Rodrigues, L.A.; Cardeira, M.; Leonardo, I.C.; Gaspar, F.B.; Redovnikovic, I.R.; Duarte, A.R.; Paiva, A.; Matias, A.A. Deep eutectic systems from betaine and polyols-Physicochemical and toxicological properties. J. Mol. Liq. 2021, 335, 116201. [Google Scholar] [CrossRef]
  35. Shahbaz, K.; Baroutian, S.; Mjalli, F.S.; Hashim, M.A.; AlNashef, I.M. Densities of ammonium and phosphonium based deep eutectic solvents: Prediction using artificial intelligence and group contribution techniques. Thermochim. Acta 2012, 527, 59–66. [Google Scholar] [CrossRef]
  36. Hayyan, A.; Mjalli, F.S.; AlNashef, I.M.; Al-Wahaibi, T.; Al-Wahaibi, Y.M.; Hashim, M.A. Fruit sugar-based deep eutectic solvents and their physical properties. Thermochim. Acta 2012, 541, 70–75. [Google Scholar] [CrossRef]
  37. Abbott, A.P.; Harris, R.C.; Ryder, K.S.; D’Agostino, C.; Gladden, L.F.; Mantle, M.D. Glycerol eutectics as sustainable solvent systems. Green Chem. 2011, 13, 82–90. [Google Scholar] [CrossRef]
  38. Gajardo-Parra, N.F.; Lubben, M.J.; Winnert, J.M.; Leiva, Á.; Brennecke, J.F.; Canales, R.I. Physicochemical properties of choline chloride-based deep eutectic solvents and excess properties of their pseudo-binary mixtures with 1-Butanol. J. Chem. Thermodyn. 2019, 133, 272–284. [Google Scholar] [CrossRef]
  39. Chen, Y.; Chen, W.; Fu, L.; Yang, Y.; Wang, Y.; Hu, X.; Wang, F.; Mu, T. The surface tension of 50 deep eutectic solvents: Effect of hydrogen-bonding donors, hydrogen-bonding acceptors, other solvents and temperature. Ind. Eng. Chem. Res. 2019, 58, 12741–12750. [Google Scholar] [CrossRef]
  40. Hansen, B.B.; Spittle, S.; Chen, B.; Poe, D.; Zhang, Y.; Klein, J.M.; Horton, A.; Adhikari, L.; Zelovich, T.; Doherty, B.W.; et al. Deep Eutectic Solvents: A Review of Fundamentals and Applications. Chem. Rev. 2021, 121, 1232–1285. [Google Scholar] [CrossRef]
Figure 1. Temperature dependence of density of 1,2-propanediol and DESs.
Figure 1. Temperature dependence of density of 1,2-propanediol and DESs.
Polymers 14 01783 g001
Figure 2. Plot of σ vs. T of 1,2-propanediol and four DESs.
Figure 2. Plot of σ vs. T of 1,2-propanediol and four DESs.
Polymers 14 01783 g002
Figure 3. Variations of surface tension with type of DESs in different temperatures.
Figure 3. Variations of surface tension with type of DESs in different temperatures.
Polymers 14 01783 g003
Figure 4. Variations of viscosity with temperature.
Figure 4. Variations of viscosity with temperature.
Polymers 14 01783 g004
Figure 5. Relationship between lnη and T−1 of different DESs.
Figure 5. Relationship between lnη and T−1 of different DESs.
Polymers 14 01783 g005
Table 1. Chemical specifications.
Table 1. Chemical specifications.
ChemicalsCAS NumberChemical FormulaSourceMass Fraction Purity (Supplier)
Betaine107-43-7C5H11NO2Aladdin>0.99
1,2-Propanediol57-55-6C3H8O2Aladdin>0.99
Table 2. Betaine-based DESs prepared in this work.
Table 2. Betaine-based DESs prepared in this work.
HBAHBDMole RatioSolutions
Betaine1,2-Propanediol1:3.5DES1
1:4DES2
1:5DES3
1:6DES4
Table 3. Density of 1,2-propanediol and DESs in different temperatures *.
Table 3. Density of 1,2-propanediol and DESs in different temperatures *.
T/(K)ρexp/(kg·m3)
1,2-PropanediolDES1DES2DES3DES4
293.151.03621.07561.07271.06701.0629
298.151.03261.072201.06941.06361.0595
303.151.02881.06871.06611.06021.0561
308.151.02511.06521.06281.05681.0527
313.151.02131.06171.05941.05341.0492
318.151.01751.05811.05611.05001.0457
323.151.01361.05461.05271.04661.0422
328.151.00981.05121.04931.04311.0387
333.151.00581.04781.04591.03961.0352
338.151.00181.04461.04251.03611.0316
343.150.99781.04151.03911.03261.0281
348.150.99371.03851.03571.02911.0245
353.150.98971.03541.03221.02551.0208
358.150.98551.03211.02871.02201.0172
363.150.98131.02861.02521.01841.0135
* The standard uncertainties (u) are u(T) = 0.01 K and ur(ρ) = 0.001.
Table 4. Parameters in equation (1) for 1,2-propanediol and different DESs.
Table 4. Parameters in equation (1) for 1,2-propanediol and different DESs.
Parameters1,2-PropanediolDES1DES2DES3DES4
A/(kg·m−3)1.266971.271211.271771.270621.26997
B × 10−4/(kg·m−3·K−1)−7.8514−6.68825−6.78264−6.93838−7.05260
Table 5. Surface tension of 1,2-propanediol and different DESs *.
Table 5. Surface tension of 1,2-propanediol and different DESs *.
T/(K)σexp/(mN·m−1)
1,2-PropanediolDES1DES2DES3DES4
293.1535.9251.2646.1041.6145.34
313.1535.1649.9844.7041.0944.01
333.1534.4049.6044.3840.6943.42
353.1533.6448.6742.6839.9842.82
* The combined expanded uncertainties Uc are Uc(T) = 0.02 K, Uc(σ) = 0.02 in the level of confidence 0.95.
Table 6. Fitted coefficients of surface tension of 1,2-propanediol and different DESs.
Table 6. Fitted coefficients of surface tension of 1,2-propanediol and different DESs.
Parameters1,2-PropanediolDES1DES2DES3DES4
a/(mN·m−1)47.059763.045961.559649.389857.0659
b/(mN·m−1·K−1)−0.0380−0.0408−0.0529−0.0265−0.0408
Table 7. Viscosity of 1,2-propanediol and different DESs *.
Table 7. Viscosity of 1,2-propanediol and different DESs *.
T/(K)ηexp/(mPa·s)
1,2-PropanediolDES1DES2DES3DES4
293.1556.876178.952156.418132.340116.689
303.1533.897103.47784.31171.55262.524
313.1519.98057.38349.71641.16936.191
323.1512.95634.60230.43125.16022.495
333.158.39923.35120.60517.33015.961
343.155.46616.86215.09612.92911.791
353.153.72112.78511.3739.9209.017
* The combined expanded uncertainties Uc are Uc(T) = 0.02 K, Uc(η) = 0.03 in the level of confidence 0.95.
Table 8. Parameters in Equation (3) for different DESs.
Table 8. Parameters in Equation (3) for different DESs.
DESsη0 /(mPa·s)Eη/(kJ·mol−1)r2
DES12.3098 × 10−538.490.9929
DES22.7029 × 10−537.700.9926
DES32.6080 × 10−537.350.9894
DES43.0196 × 10−536.660.9889
Table 9. Comparison between experimental and calculated viscosity of different DESs.
Table 9. Comparison between experimental and calculated viscosity of different DESs.
T/(K)293.15303.15313.15323.15333.15343.15353.15
DES1 Eη = 38.49 (kJ·mol−1) η0 = 2.3098 × 10−5 (mPa·s)
ηexp/(mPa·s)178.952103.47757.38334.60223.35116.86212.785
ηcal/(mPa·s) 166.733 99.033 60.812 38.486 25.035 16.698 11.396
Dev(%)6.824.295.9711.22 7.21 0.9710.86
DES2 Eη = 37.70 (kJ·mol−1) η0 = 2.7029 × 10−5 (mPa·s)
ηexp/(mPa·s)156.41884.31149.71630.43120.60515.09611.373
ηcal/(mPa·s) 141.128 84.726 52.549 33.571 22.031 14.817 10.192
Dev(%)9.77 0.49 5.70 10.32 6.92 1.85 10.38
DES3 Eη = 37.35 (kJ·mol−1) η0 = 2.6080 × 10−5 (mPa·s)
ηexp/(mPa·s)132.34071.55241.16925.16017.33012.9299.920
ηcal/(mPa·s) 117.957 71.151 44.326 28.436 18.734 12.647 8.729
Dev(%)10.87 0.56 7.67 13.02 8.10 2.18 12.00
DES4 Eη = 36.66 (kJ·mol−1) η0 = 3.0196 × 10−5 (mPa·s)
ηexp/(mPa·s)116.68962.52436.19122.49515.96111.7919.017
ηcal/(mPa·s) 102.901 62.652 39.374 25.467 16.908 11.497 7.990
Dev(%)11.82 0.20 8.80 13.21 5.93 2.49 11.39
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Q.; He, N.; Fan, J.; Song, F. Physical Properties of Betaine-1,2-Propanediol-Based Deep Eutectic Solvents. Polymers 2022, 14, 1783. https://0-doi-org.brum.beds.ac.uk/10.3390/polym14091783

AMA Style

Chen Q, He N, Fan J, Song F. Physical Properties of Betaine-1,2-Propanediol-Based Deep Eutectic Solvents. Polymers. 2022; 14(9):1783. https://0-doi-org.brum.beds.ac.uk/10.3390/polym14091783

Chicago/Turabian Style

Chen, Qicheng, Nan He, Jing Fan, and Fenhong Song. 2022. "Physical Properties of Betaine-1,2-Propanediol-Based Deep Eutectic Solvents" Polymers 14, no. 9: 1783. https://0-doi-org.brum.beds.ac.uk/10.3390/polym14091783

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

Article Metrics

Back to TopTop