Next Article in Journal
Reduced Collatz Dynamics Data Reveals Properties for the Future Proof of Collatz Conjecture
Next Article in Special Issue
Geometrical Platform of Big Database Computing for Modeling of Complex Physical Phenomena in Electric Current Treatment of Liquid Metals
Previous Article in Journal
Dataset on Substrate-Borne Vibrations of Constrictotermes cyphergaster (Blattodea: Isoptera) Termites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Data Descriptor

The Ionic Liquid Property Explorer: An Extensive Library of Task-Specific Solvents

by
Vishwesh Venkatraman
*,
Sigvart Evjen
and
Kallidanthiyil Chellappan Lethesh
Department of Chemistry, Norwegian University of Science and Technology, 7491 Trondheim, Norway
*
Author to whom correspondence should be addressed.
Submission received: 25 April 2019 / Revised: 18 June 2019 / Accepted: 21 June 2019 / Published: 21 June 2019
(This article belongs to the Special Issue Machine Learning and Materials Informatics)

Abstract

:
Ionic liquids have a broad spectrum of applications ranging from gas separation to sensors and pharmaceuticals. Rational selection of the constituent ions is key to achieving tailor-made materials with functional properties. To facilitate the discovery of new ionic liquids for sustainable applications, we have created a virtual library of over 8 million synthetically feasible ionic liquids. Each structure has been evaluated for their-task suitability using data-driven statistical models calculated for 12 highly relevant properties: melting point, thermal decomposition, glass transition, heat capacity, viscosity, density, cytotoxicity, CO 2 solubility, surface tension, and electrical and thermal conductivity. For comparison, values of six properties computed using quantum chemistry based equilibrium thermodynamics COSMO-RS methods are also provided. We believe the data set will be useful for future efforts directed towards targeted synthesis and optimization.
Data Set: The datasets used for machine learning can be accessed at dx.doi.org/10.5281/zenodo.3251643. The SQLite database containing computed properties and a graphical user interface for querying, are available from dx.doi.org/10.5281/zenodo.3251661.
Data Set License: The data set is made available under the Creative Commons License CC BY 4.0.

Graphical Abstract

1. Introduction

Ionic liquids (ILs) comprised of cations (mostly organic) and anions (both organic and inorganic) provide a widely applicable set of building blocks for advanced functional materials. Demonstrated applications include coatings and lubricants [1], pharmaceuticals [2], fuel cells [3], and catalysis [4]. Desirable properties such as high thermal and electrochemical stabilities, together with a negligible volatility, make them well-suited for developing novel and innovative materials. By making simple changes to the structure of the constituent ions, the chemical makeup can be altered to create the optimum solvent for a given application. The challenge, however, is to identify optimal task-specific ILs from the near-infinite combinations of the constitutive ions and functional groups [5,6]. Experimental work is limited to a small area of the ionic liquid chemical space, leaving many potentially promising compounds unexplored. Given the need to minimize experimental cost and time, the rational selection of suitable ILs from the available choices becomes paramount.
Currently, most ILs are discovered through laborious trial-and-error experiments. Given the diversity of the IL applications, the custom design of the solvents requires general knowledge of the properties, to carry out even preliminary studies. Chemical intuition and experience play key roles in the selection. In a number of studies, approaches based on electronic structure theory have been used to understand the structure–property mechanisms [7,8]. The use of electronic structure methods [9] are, however, confined to a few systems, and despite advances in computer hardware, the associated computational costs are still prohibitive for rapid large scale screening. In recent times, machine learning (ML) based virtual screening have emerged as a powerful approach facilitating in silico searches over millions of compounds [10,11,12]. With the availability of data repositories such as ILThermo [13], there has been a steep rise in the use of such approaches for modelling ionic liquid properties as evidenced by recent publications [14,15,16,17,18].
With a view to expedite task-specific ionic liquid discovery, we have assembled a large library of ionic liquids spanning nine different cation scaffolds: ammonium, imidazolium, phosphonium, piperidinium, pyridinium, pyrrolidinium, morpholinium, azepanium and sulphonium. These are combined with a diverse set of anions (alkylsulphonates, phenolates, phosphates, triazolides, HF 6 , BF 4 ) yielding over 8 million compounds. For each IL, twelve properties of interest have been predicted using machine learning. The conductor-like screening model for real solvents (COSMO-RS) developed by Klamt and Eckert [19] has been shown to be a relatively robust predictive method for properties such as activity coefficients of molecular solutes in ionic liquids [20], gas separation capacity [21] and cellulose solubilities [22]. The COSMO-RS approach, however, requires computationally expensive density functional theory (DFT) calculations and has been shown to fail as much as it succeeds [23]. It nevertheless remains a popular choice for ionic liquid-based screening [24,25,26] and is likely to improve with better parameterization [27]. We have therefore included COSMO-RS estimates for selected properties. The database can be expanded by way of progressively adding new structures and associated properties. We expect this repository to be a playground for future work in this active research area.

2. Data Description

2.1. Workflow

Figure 1 provides a schematic overview of the database. Using SmiLib [28] as the combinatorial enumeration engine, 219,216 cations spanning nine different cation families were created. These were combined with 38 anions to yield a total of 8,333,096 ionic liquids. For each IL, values for twelve properties of interest are reported: melting point ( T m ), glass transition temperature ( T g ), thermal decomposition temperature ( T d ), viscosity ( η ), density ( ρ ), heat capacity ( C p ), CO 2 capacity ( x C O 2 ), electrical ( κ ) and thermal conductivity ( λ ), cytotoxicity towards the leukemia rat cell line IPC-81 ( log 10 EC 50 ) , surface tension ( σ ) and refractive index ( n D ). For each property, an estimate of the uncertainty associated (±1 standard deviation) with the prediction is included as a way to assess when the model is likely to be more accurate. Large standard deviations are typically associated with those that cannot be trusted [17,29].

2.2. Ionic Liquid Library

The library spans a wide range of functionalities, from simple alkyl-functionalized cations to more exotic structures. Moieties were selected based on ease of synthesis i.e., those that can be readily prepared, such as the availability of halide precursors for S N 2-reaction preparation. Amine, ether and alcohol groups are able to coordinate to metals. Alcohols and amines can engage in H-bonding, allowing for selective interactions with hydrophilic compounds. Amines with different levels of basicity were incorporated into the database, facilitating more application-specific tuning of IL p K a , e.g., selective probes for acid gases [30].
The customized cation libraries were built using combinatorial enumeration (using the software SmiLib [28]) of building blocks attached to different scaffolds (see Table 1). The counterions were selected among common anion groups for ILs, as well as organic anions, which possess several interesting properties. For instance, acetates have a high cellulose solubility and phenolates can efficiently be used for extraction of acids [31,32]. Different carboxylates allow for the tuning of the IL properties. Acetylacetonate-based ILs could be used for selective metal extraction [33,34]. Sulfonate anions allow for the extraction of hydrocarbons [35], and have potential applications in batteries [36].

2.3. Graphical Summary of the Data Set

The variation in the values for the ML-predicted properties are shown in Figure 2, Figure 3 and Figure 4. Values for C p , η , ρ , κ , λ , x C O 2 are estimated at standard room temperature and pressure. For many properties, the estimated values mirror the trends observed in literature. For instance, the T m for phosphonium- and ammonium-based ILs are higher than those based on the imidazolium scaffold [37,38,39]. These differences are related to the highly symmetric nature of the tetraalkylammonium and tetraalkylphosphonium cations. The compounds in the library are thermally stable, with phosphonium-based ILs in particular showing higher values of T d compared to the other cationic cores (see Figure 2). This is attributed to the difference in the thermal degradation mechanisms [40]. In contrast, sulphonium based ILs show poorer stabilities owing to the unstable nature of the sulphur atom [41].
Low melting ILs are expected to have low viscosities. Analysis of the data suggests that less than 2% of the ILs have η < 300 cP and T m < 30 ° C. It has been reported that the viscosities of imidazolium and sulphonium ILs are on the lower side because of the asymmetric nature of the imidazolium cation and bulky nature of the sulphur atom [41,42]. The predicted viscosities show similar trends (shown in Figure 3). High refractive index ILs ( n D > 1.60 ) are much needed in optical microscopy studies of minerals [43]. While the predictions of n D , in particular for imidazolium and pyridinium ILs, are in accordance with experimentally observed data for similar compounds [44,45], none of the predictions exceed 1.60. Imidazolium-based ILs generally have higher electrical conductivities compared to other ILs based on other cationic cores, owing to their lower viscosity [46]. In the predicted data, over 100 low-viscosity imidazolium ILs show promise as electrolyte materials in battery applications.
In all cases, anions play a crucial role in determining the behaviour of the ionic liquids. Fluorinated anions (PF 6 , NTf2, hfac), for instance, show higher CO 2 solubilities [47] (see Figure 4). Although considered environmentally friendly, many ILs are soluble in water which can be hazardous to aquatic organisms, if released into the aqueous system [48,49]. It has been shown that toxicity increases with the length of the alkyl chain (varying between 2 and 12 carbon atoms) attached to the cationic cores. The effect of anions has not been investigated broadly to allow a conclusive analysis. The graph for cytotoxicity shown in Figure 4 suggests that, of the 8.33 million combinations examined, only 0.2% of the ILs showed log 10 EC 50 > 3.4 μ M) [50].

3. Methods

3.1. Machine Learning

Each structure was subjected to geometry optimization at the semi-empirical PM6 level using MOPAC [51]. Although fragment/group contribution descriptors have been popular, models trained on such variables often fail when presented with new fragments for which they were not trained. We have therefore chosen molecular descriptors that focus on charge distributions and geometrical indices which have been shown to yield good predictive performances for IL properties such as melting points [17], thermal decomposition temperatures [15], refractive indices [18] and CO 2 solubilities [52]. The variables were calculated independently for each cation and anion using the software KrakenX [53,54]. The top ranked variables (ranked according to the contribution of the variable to the response) in the models included the charged partial surface area descriptors (summarize the charge distribution in the ion), chemical reactivity parameters such as the HOMO/LUMO energies that are closely related to electrophilic/nucleophilic attack and the charge distribution in the ion, and softness (inverse of the HOMO-LUMO gap) which are indicative of the cation-anion electrostatic (nucleophilic-electrophilic) interactions. Experimental data for the properties was taken from various literature sources [13,17,18,52,55]. Machine learning models were evaluated for 12 different properties: melting points ( T m ), glass transition temperatures ( T g ), thermal decomposition temperatures ( T d ), viscosities ( η ), densities ( ρ ), heat capacities ( C p ), CO 2 capacity ( x C O 2 ), electrical ( κ ) and thermal conductivities ( λ ), cytotoxicities towards the leukemia rat cell line IPC-81 ( log 10 ( E C 50 ) , surface tension ( σ ) and refractive indices ( n D ). The models were trained on 67% of the available data and a 5-fold cross validation technique was used to obtain performance statistics. Three different ML models were employed: generalized boosted regression models (GBM), random forests (RF) and Cubist methods. For each property, the best performing model across both calibration and test data was determined based on standard evaluation metrics and used for further predictions.
For the obtained models, performance metrics including the squared coefficient of correlation ( R 2 ), root mean square error ( R M S E ) and the mean absolute error ( M A E ) are reported in Table 2. The supplementary material lists the performances of all the ML models used for prediction. For most IL properties, the experimental and predicted values are in agreement with reported studies [14,15,17,18,52]. The larger deviations for T m , T g and η can be attributed to experimental variations, presence of impurities and water [56]. Model applicability domain methods often rely on the chemical similarity of a test set structure to members of the training set [57]. Here, we have chosen to associate each prediction with a bootstrapped uncertainty estimate [17,18]. In bootstrapping [58], for example, the training set is randomly sampled with replacement, and a model is built for each bootstrapped sample. For computational expediency, a total of 100 models were built, and each model was applied to a given test set compound to obtain a distribution of predictions. The uncertainty associated with the prediction was then calculated as the standard deviation of the distribution [29]. Working on the assumption that ILs with small uncertainties are likely to have small prediction errors, one can exclude compounds with moderate to high prediction uncertainties.

3.2. COSMO-RS Evaluation

COSMO-RS, a quantum chemistry based method, was used to evaluate selected properties [59]. For each cation and anion, geometry optimizations using the density functional theory (DFT) functional B88-PW86 with a triple zeta valence polarized basis set [60] (TZVP) and the resolution of identity standard approximation were performed. Values of η , ρ , C p , x C O 2 , κ and T m were calculated using the COSMOtherm software with the parameterization set BP_TZVP_C30_01601) [61].

4. Database Exploration and Use

An interactive graphical user interface written in the Java programming language is provided as a way to query the library (see Figure 5). All data including machine learning and COSMO-RS predictions have been compiled into an SQLite database [62]. Structures can be searched (exact or substructure) as SMILES strings and IUPAC names of moieties. Alternatively, structure files in MOL format can be uploaded. For conversion of IUPAC names to SMILES representations, the functionality in the OPSIN [63] library has been used. Filtering of compounds according to select property cut-offs allows for a targeted search. Query results may be additionally saved as a sortable HTML table for future inspection.

Supplementary Materials

The following are available online at https://www.mdpi.com/2306-5729/4/2/88/s1.

Author Contributions

V.V. performed the data curation, developed the models and carried out the computations. S.E. and K.C.L. proposed the molecular library and tested the software. V.V. wrote the paper with contributions from S.E. and K.C.L.

Funding

The Norwegian Research Council (NFR) is acknowledged for financial support through the CLIMIT program (Grant No. 233776).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ILIonic Liquids
MLMachine Learning
DFTDensity Functional Theory
RFRandom Forest
GBMGeneralized Boosted Models
COSMO-RSConductor like Screening Model for Real Solvents
hfacHexafluoroacetylacetonate
DCADicyanamide
NTf2Bis(trifluoromethanesulfonyl)imide
PF6Hexafluorophosphate
HOMOHighest Occupied Molecular Orbital
LUMOLowest Unoccupied Molecular Orbital

References

  1. Zhou, Y.; Qu, J. Ionic Liquids as Lubricant Additives: A Review. ACS Appl. Mater. Interfaces 2017, 9, 3209–3222. [Google Scholar] [CrossRef] [PubMed]
  2. Marrucho, I.; Branco, L.; Rebelo, L. Ionic Liquids in Pharmaceutical Applications. Annu. Rev. Chem. Biomol. Eng. 2014, 5, 527–546. [Google Scholar] [CrossRef] [PubMed]
  3. MacFarlane, D.R.; Tachikawa, N.; Forsyth, M.; Pringle, J.M.; Howlett, P.C.; Elliott, G.D.; Davis, J.H.; Watanabe, M.; Simon, P.; Angell, C.A. Energy applications of ionic liquids. Energy Environ. Sci. 2014, 7, 232–250. [Google Scholar] [CrossRef]
  4. Dai, C.; Zhang, J.; Huang, C.; Lei, Z. Ionic Liquids in Selective Oxidation: Catalysts and Solvents. Chem. Rev. 2017, 117, 6929–6983. [Google Scholar] [CrossRef] [PubMed]
  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] [PubMed]
  6. Niedermeyer, H.; Hallett, J.P.; Villar-Garcia, I.J.; Hunt, P.A.; Welton, T. Mixtures of ionic liquids. Chem. Soc. Rev. 2012, 41, 7780. [Google Scholar] [CrossRef]
  7. Ilawe, N.V.; Fu, J.; Ramanathan, S.; Wong, B.M.; Wu, J. Chemical and Radiation Stability of Ionic Liquids: A Computational Screening Study. J. Phys. Chem. C 2016, 120, 27757–27767. [Google Scholar] [CrossRef]
  8. Karu, K.; Ruzanov, A.; Ers, H.; Ivaništšev, V.; Lage-Estebanez, I.; de la Vega, J.G. Predictions of Physicochemical Properties of Ionic Liquids with DFT. Computation 2016, 4, 25. [Google Scholar] [CrossRef]
  9. Izgorodina, E.I. Towards large-scale fully ab initio calculations of ionic liquids. Phys. Chem. Chem. Phys. 2011, 13, 4189–4207. [Google Scholar] [CrossRef]
  10. Carpenter, K.A.; Huang, X. Machine Learning-based Virtual Screening and Its Applications to Alzheimer’s Drug Discovery: A Review. Curr. Pharm. Des. 2018, 24, 3347–3358. [Google Scholar] [CrossRef]
  11. Jørgensen, P.B.; Mesta, M.; Shil, S.; Lastra, J.M.G.; Jacobsen, K.W.; Thygesen, K.S.; Schmidt, M.N. Machine learning-based screening of complex molecules for polymer solar cells. J. Chem. Phys. 2018, 148, 241735. [Google Scholar] [CrossRef] [PubMed]
  12. Ichikawa, D.; Saito, T.; Ujita, W.; Oyama, H. How can machine-learning methods assist in virtual screening for hyperuricemia? A healthcare machine-learning approach. J. Biomed. Inform. 2016, 64, 20–24. [Google Scholar] [CrossRef] [PubMed]
  13. Dong, Q.; Muzny, C.D.; Kazakov, A.; Diky, V.; Magee, J.W.; Widegren, J.A.; Chirico, R.D.; Marsh, K.N.; Frenkel, M. ILThermo: A Free-Access Web Database for Thermodynamic Properties of Ionic Liquids. J. Chem. Eng. Data 2007, 52, 1151–1159. [Google Scholar] [CrossRef]
  14. Paduszyński, K.; Domańska, U. Viscosity of Ionic Liquids: An Extensive Database and a New Group Contribution Model Based on a Feed-Forward Artificial Neural Network. J. Chem. Inf. Model. 2014, 54, 1311–1324. [Google Scholar] [CrossRef] [PubMed]
  15. Venkatraman, V.; Alsberg, B.K. Quantitative structure-property relationship modelling of thermal decomposition temperatures of ionic liquids. J. Mol. Liq. 2016, 223, 60–67. [Google Scholar] [CrossRef]
  16. Rybinska-Fryca, A.; Sosnowska, A.; Puzyn, T. Prediction of dielectric constant of ionic liquids. J. Mol. Liq. 2018, 260, 57–64. [Google Scholar] [CrossRef]
  17. Venkatraman, V.; Evjen, S.; Knuutila, H.K.; Fiksdahl, A.; Alsberg, B.K. Predicting Ionic Liquid Melting Points using Machine Learning. J. Mol. Liq. 2018, 264, 318–326. [Google Scholar] [CrossRef]
  18. Venkatraman, V.; Raj, J.J.; Evjen, S.; Lethesh, K.C.; Fiksdahl, A. In silico prediction and experimental verification of ionic liquid refractive indices. J. Mol. Liq. 2018, 264, 563–570. [Google Scholar] [CrossRef]
  19. Klamt, A.; Eckert, F. COSMO-RS: A novel and efficient method for the a priori prediction of thermophysical data of liquids. Fluid Phase Equilib. 2000, 172, 43–72. [Google Scholar] [CrossRef]
  20. Paduszyński, K. An overview of the performance of the COSMO-RS approach in predicting the activity coefficients of molecular solutes in ionic liquids and derived properties at infinite dilution. Phys. Chem. Chem. Phys. 2017, 19, 11835–11850. [Google Scholar] [CrossRef]
  21. Liu, X.; Zhou, T.; Zhang, X.; Zhang, S.; Liang, X.; Gani, R.; Kontogeorgis, G.M. Application of COSMO-RS and UNIFAC for ionic liquids based gas separation. Chem. Eng. Sci. 2018, 192, 816–828. [Google Scholar] [CrossRef]
  22. Kahlen, J.; Masuch, K.; Leonhard, K. Modelling cellulose solubilities in ionic liquids using COSMO-RS. Green Chem. 2010, 12, 2172. [Google Scholar] [CrossRef]
  23. Izgorodina, E.I.; Seeger, Z.L.; Scarborough, D.L.A.; Tan, S.Y.S. Quantum Chemical Methods for the Prediction of Energetic, Physical, and Spectroscopic Properties of Ionic Liquids. Chem. Rev. 2017, 117, 6696–6754. [Google Scholar] [CrossRef] [PubMed]
  24. Anantharaj, R.; Banerjee, T. COSMO-RS-Based Screening of Ionic Liquids as Green Solvents in Denitrification Studies. Ind. Eng. Chem. Res. 2010, 49, 8705–8725. [Google Scholar] [CrossRef]
  25. Jeliński, T.; Cysewski, P. Screening of ionic liquids for efficient extraction of methylxanthines using COSMO-RS methodology. Chem. Eng. Res. Des. 2017, 122, 176–183. [Google Scholar] [CrossRef]
  26. Motlagh, S.R.; Harun, R.; Biak, D.A.; Hussain, S.; Ghani, W.W.A.K.; Khezri, R.; Wilfred, C.; Elgharbawy, A. Screening of Suitable Ionic Liquids as Green Solvents for Extraction of Eicosapentaenoic Acid (EPA) from Microalgae Biomass Using COSMO-RS Model. Molecules 2019, 24, 713. [Google Scholar] [CrossRef] [PubMed]
  27. Han, J.; Dai, C.; Yu, G.; Lei, Z. Parameterization of COSMO-RS model for ionic liquids. Green Energy Environ. 2018, 3, 247–265. [Google Scholar] [CrossRef]
  28. Schüller, A.; Hähnke, V.; Schneider, G. SmiLib v2.0: A Java-Based Tool for Rapid Combinatorial Library Enumeration. Mol. Inf. 2007, 26, 407–410. [Google Scholar] [CrossRef]
  29. Toplak, M.; Močnik, R.; Polajnar, M.; Bosnić, Z.; Carlsson, L.; Hasselgren, C.; Demšar, J.; Boyer, S.; Zupan, B.; Stålring, J. Assessment of Machine Learning Reliability Methods for Quantifying the Applicability Domain of QSAR Regression Models. J. Chem. Inf. Model. 2014, 54, 431–441. [Google Scholar] [CrossRef] [Green Version]
  30. Korotcenkov, G. Ionic Liquids in Gas Sensors. In Integrated Analytical Systems; Springer: New York, NY, USA, 2013; pp. 121–130. [Google Scholar] [CrossRef]
  31. Shah, S.N.; Mutalib, M.A.; Ismail, M.F.; Suleman, H.; Lethesh, K.C.; Pilus, R.B.M. Thermodynamic modelling of liquid-liquid extraction of naphthenic acid from dodecane using imidazolium based phenolate ionic liquids. J. Mol. Liq. 2016, 219, 513–525. [Google Scholar] [CrossRef]
  32. Zhang, J.; Wu, J.; Yu, J.; Zhang, X.; He, J.; Zhang, J. Application of ionic liquids for dissolving cellulose and fabricating cellulose-based materials: State of the art and future trends. Mater. Chem. Front. 2017, 1, 1273–1290. [Google Scholar] [CrossRef]
  33. Yoshii, K.; Oshino, Y.; Tachikawa, N.; Toshima, K.; Katayama, Y. Electrodeposition of palladium from palladium(II) acetylacetonate in an amide-type ionic liquid. Electrochem. Commun. 2015, 52, 21–24. [Google Scholar] [CrossRef]
  34. Nunes, P.; Nagy, N.V.; Alegria, E.C.; Pombeiro, A.J.; Correia, I. The solvation and electrochemical behavior of copper acetylacetonate complexes in ionic liquids. J. Mol. Struct. 2014, 1060, 142–149. [Google Scholar] [CrossRef] [Green Version]
  35. García, S.; Garciá, J.; Larriba, M.; Torrecilla, J.S.; Rodríguez, F. Sulfonate-Based Ionic Liquids in the Liquid–Liquid Extraction of Aromatic Hydrocarbons. J. Chem. Eng. Data 2011, 56, 3188–3193. [Google Scholar] [CrossRef]
  36. Dupont, D.; Raiguel, S.; Binnemans, K. Sulfonic acid functionalized ionic liquids for dissolution of metal oxides and solvent extraction of metal ions. Chem. Comm. 2015, 51, 9006–9009. [Google Scholar] [CrossRef] [PubMed]
  37. Tsunashima, K.; Kodama, S.; Sugiya, M.; Kunugi, Y. Physical and electrochemical properties of room-temperature dicyanamide ionic liquids based on quaternary phosphonium cations. Electrochim. Acta 2010, 56, 762–766. [Google Scholar] [CrossRef]
  38. Deive, F.J.; Rivas, M.A.; Rodríguez, A. Study of thermodynamic and transport properties of phosphonium-based ionic liquids. J. Chem. Thermodyn. 2013, 62, 98–103. [Google Scholar] [CrossRef]
  39. Kulkarni, P.S.; Branco, L.C.; Crespo, J.G.; Nunes, M.C.; Raymundo, A.; Afonso, C.A.M. Comparison of Physicochemical Properties of New Ionic Liquids Based on Imidazolium, Quaternary Ammonium, and Guanidinium Cations. Chem. Eur. J. 2007, 13, 8478–8488. [Google Scholar] [CrossRef]
  40. Maton, C.; Vos, N.D.; Stevens, C.V. Ionic liquid thermal stabilities: Decomposition mechanisms and analysis tools. Chem. Soc. Rev. 2013, 42, 5963. [Google Scholar] [CrossRef]
  41. Zhang, Q.; Liu, S.; Li, Z.; Li, J.; Chen, Z.; Wang, R.; Lu, L.; Deng, Y. Novel Cyclic Sulfonium-Based Ionic Liquids: Synthesis, Characterization, and Physicochemical Properties. Chem. Eur. J. 2009, 15, 765–778. [Google Scholar] [CrossRef]
  42. Sánchez, L.G.; Espel, J.R.; Onink, F.; Meindersma, G.W.; de Haan, A.B. Density, Viscosity, and Surface Tension of Synthesis Grade Imidazolium, Pyridinium, and Pyrrolidinium Based Room Temperature Ionic Liquids. J. Chem. Eng. Data 2009, 54, 2803–2812. [Google Scholar] [CrossRef]
  43. Deetlefs, M.; Shara, M.; Seddon, K.R. Refractive Indices of Ionic Liquids. In ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2005; pp. 219–233. [Google Scholar] [CrossRef]
  44. Tariq, M.; Forte, P.; Gomes, M.C.; Lopes, J.C.; Rebelo, L. Densities and refractive indices of imidazolium- and phosphonium-based ionic liquids: Effect of temperature, alkyl chain length, and anion. J. Chem. Thermodyn. 2009, 41, 790–798. [Google Scholar] [CrossRef]
  45. Yunus, N.M.; Mutalib, M.A.; Man, Z.; Bustam, M.A.; Murugesan, T. Thermophysical properties of 1-alkylpyridinum bis(trifluoromethylsulfonyl)imide ionic liquids. J. Chem. Thermodyn. 2010, 42, 491–495. [Google Scholar] [CrossRef]
  46. Ohno, H. (Ed.) Electrochemical Aspects of Ionic Liquids; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2005. [Google Scholar] [CrossRef]
  47. Ramdin, M.; de Loos, T.W.; Vlugt, T.J. State-of-the-Art of CO2 Capture with Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 8149–8177. [Google Scholar] [CrossRef]
  48. 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] [PubMed]
  49. Hartmann, D.O.; Pereira, C.S. Toxicity of Ionic Liquids. In Ionic Liquids in Lipid Processing and Analysis; Elsevier: Amsterdam, The Netherlands, 2016; pp. 403–421. [Google Scholar] [CrossRef]
  50. Torrecilla, J.S.; Palomar, J.; Lemus, J.; Rodríguez, F. A quantum-chemical-based guide to analyze/quantify the cytotoxicity of ionic liquids. Green Chem. 2010, 12, 123–134. [Google Scholar] [CrossRef]
  51. Stewart, J.J.P. MOPAC2016. Stewart Computational Chemistry: Colorado Springs, CO, USA, 2016. Available online: http://openmopac.net (accessed on 21 June 2019).
  52. Venkatraman, V.; Alsberg, B.K. Predicting CO 2 capture of ionic liquids using machine learning. J. CO2 Util. 2017, 21, 162–168. [Google Scholar] [CrossRef]
  53. Venkatraman, V.; Alsberg, B.K. KRAKENX: Software for the generation of alignment-independent 3D descriptors. J. Mol. Model. 2016, 22, 1–8. [Google Scholar] [CrossRef]
  54. Venkatraman, V. KrakenX. 2019. Available online: https://gitlab.com/vishsoft/krakenx (accessed on 21 June 2019).
  55. Zhang, S.; Lu, X.; Zhou, Q.; Li, X.; Zhang, X.; Li, S. Ionic Liquids Physicochemical Properties; Elsevier: Amsterdam, The Netherlands, 2009. [Google Scholar]
  56. Coutinho, J.A.P.; Carvalho, P.J.; Oliveira, N.M.C. Predictive methods for the estimation of thermophysical properties of ionic liquids. RSC Adv. 2012, 2, 7322. [Google Scholar] [CrossRef]
  57. Hanser, T.; Barber, C.; Marchaland, J.F.; Werner, S. Applicability domain: Towards a more formal definition$. SAR QSAR Environ. Res. 2016, 27, 865–881. [Google Scholar] [CrossRef]
  58. Efron, B. Better Bootstrap Confidence Intervals. J. Am. Stat. Assoc. 1987, 82, 171–185. [Google Scholar] [CrossRef]
  59. Klamt, A. The COSMO and COSMO-RS solvation models. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 699–709. [Google Scholar] [CrossRef]
  60. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [Google Scholar] [CrossRef] [PubMed]
  61. Eckert, F.; Klamt, A. COSMOtherm Version C3.0, Release 16.01; COSMOlogic GmbH & Co. KG: Leverkusen, Germany, 2015. [Google Scholar]
  62. Team, S.D. SQLite Version 3.27.2. 2019. Available online: https://www.sqlite.org/releaselog/3_27_2.html (accessed on 21 June 2019).
  63. Lowe, D.M.; Corbett, P.T.; Murray-Rust, P.; Glen, R.C. Chemical Name to Structure: OPSIN, an Open Source Solution. J. Chem. Inf. Model. 2011, 51, 739–753. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Fragments and other functional moieties used in the construction of the ionic liquid library.
Figure 1. Fragments and other functional moieties used in the construction of the ionic liquid library.
Data 04 00088 g001
Figure 2. Violin plots showing the distribution of the ML predictions for thermophysical properties. Predicted values for C p and λ are calculated at standard room temperature and pressure.
Figure 2. Violin plots showing the distribution of the ML predictions for thermophysical properties. Predicted values for C p and λ are calculated at standard room temperature and pressure.
Data 04 00088 g002
Figure 3. Violin plots showing the distribution of the ML-predictions for volumetric properties. The predicted values are calculated at standard room temperature and pressure.
Figure 3. Violin plots showing the distribution of the ML-predictions for volumetric properties. The predicted values are calculated at standard room temperature and pressure.
Data 04 00088 g003
Figure 4. Distribution of the predicted CO 2 mole fractions and log 10 EC 50 with respect to the anion groups.
Figure 4. Distribution of the predicted CO 2 mole fractions and log 10 EC 50 with respect to the anion groups.
Data 04 00088 g004
Figure 5. Interface to the Ionic Liquid Property Explorer is through a graphical user interface that allows for searching the database. Where available, COSMO-RS values are displayed as tool tips. The results can also be exported to a sortable HTML table.
Figure 5. Interface to the Ionic Liquid Property Explorer is through a graphical user interface that allows for searching the database. Where available, COSMO-RS values are displayed as tool tips. The results can also be exported to a sortable HTML table.
Data 04 00088 g005
Table 1. The second column shows the number of cations obtained using combinatorial library enumeration. The final column gives the number of ionic liquids obtained.
Table 1. The second column shows the number of cations obtained using combinatorial library enumeration. The final column gives the number of ionic liquids obtained.
Cation#Molecules# Ionic Liquids
ammonium1794666819708
azepanium5460207480
imidazolium5460207480
morpholinium5460207480
phosphonium7914300732
piperidinium5460207480
pyridinium104039520
pyrrolidinium5460207480
sulphonium3576135888
Table 2. Summary of the ML model performances for 12 different properties. In each case, the number of data points available for calibration and testing are provided. The metrics are reported the best performing model. EC 50 values correspond to Rat cell line toxicities. N C and N A are the number of cations and anions, respectively.
Table 2. Summary of the ML model performances for 12 different properties. In each case, the number of data points available for calibration and testing are provided. The metrics are reported the best performing model. EC 50 values correspond to Rat cell line toxicities. N C and N A are the number of cations and anions, respectively.
Property N C N A MLCalibrationValidation
N cal RMSE ( MAE ) R cv 2 N val ( MAE ) R cv 2
T m ( ° C)1369141RF148644 (15)0.6772645 (33)0.66
T g ( ° C)327109Cubist44221 (2)0.6220219 (12)0.62
T d ( ° C)538192RF83339 (12)0.7745535 (25)0.8
l o g 10 ( η ) (mPa·s)847227GBM46580.17 (0.06)0.9439940.35 (0.23)0.76
ρ (kg/m 3 )333120RF922512.23 (2.65)0.99773149.10 (28.34)0.93
ln ( C p ) (J/K/mol)11548GBM63200.042 (0.012)0.9927630.28 (0.19)0.91
γ (N/m)13168GBM18630.001 (0.0002)0.9711170.004 (0.0027)0.79
n D 23785GBM16460.006 (0.002)0.9714560.017 (0.011)0.83
x C O 2 7874GBM60840.03 (0.01)0.9848390.09 (0.06)0.86
l o g 10 ( E C 50 ) μ M11425Cubist1570.52 (0.05)0.79700.40 (0.30)0.86
κ (S/m)15880GBM14330.05 (0.01)0.9812510.15 (0.10)0.79
λ (W/m/K)2828GBM3260.005 (0.002)0.951470.009 (0.006)0.89

Share and Cite

MDPI and ACS Style

Venkatraman, V.; Evjen, S.; Chellappan Lethesh, K. The Ionic Liquid Property Explorer: An Extensive Library of Task-Specific Solvents. Data 2019, 4, 88. https://doi.org/10.3390/data4020088

AMA Style

Venkatraman V, Evjen S, Chellappan Lethesh K. The Ionic Liquid Property Explorer: An Extensive Library of Task-Specific Solvents. Data. 2019; 4(2):88. https://doi.org/10.3390/data4020088

Chicago/Turabian Style

Venkatraman, Vishwesh, Sigvart Evjen, and Kallidanthiyil Chellappan Lethesh. 2019. "The Ionic Liquid Property Explorer: An Extensive Library of Task-Specific Solvents" Data 4, no. 2: 88. https://doi.org/10.3390/data4020088

Article Metrics

Back to TopTop