Next Article in Journal
Conformational Aspects of the O-acetylation of C-tetra(phenyl)calixpyrogallol[4]arene
Previous Article in Journal
d-Amino Acid Peptide Residualizing Agents for Protein Radioiodination: Effect of Aspartate for Glutamate Substitution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Calculation of the Surface Tension of Ordinary Organic and Ionic Liquids by Means of a Generally Applicable Computer Algorithm Based on the Group-Additivity Method

1
Department of Chemistry, University of Basel, 4003 Basel, Switzerland
2
Department of Chemistry, University of North Texas, Denton, TX 76203, USA
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(5), 1224; https://doi.org/10.3390/molecules23051224
Submission received: 16 April 2018 / Revised: 16 May 2018 / Accepted: 17 May 2018 / Published: 20 May 2018

Abstract

:
The calculation of the surface tension of ordinary organic and ionic liquids, based on a computer algorithm applying a refined group-additivity method, is presented. The refinement consists of the complete breakdown of the molecules into their constituting atoms, further distinguishing them by their immediate neighbour atoms and bond constitution. The evaluation of the atom-groups’ contributions was carried out by means of a fast Gauss-Seidel fitting method, founded upon the experimental data of 1893 compounds from literature. The result has been tested for plausibility using a 10-fold cross-validation (cv) procedure. The direct calculation and the cv test proved the applicability of the present method by the close similarity and excellent goodness of fit R2 and Q2 of 0.9039 and 0.8823, respectively. The respective standard deviations are ±1.99 and ±2.16 dyn/cm. Some correlation peculiarities have been observed in a series of ordinary and ionic liquids with homologous alkyl chains, as well as with di- and trihydroxy-groups-containing liquids, which have been discussed in detail, exhibiting the limit of the present method.

1. Introduction

Surface tension has received increasing interest in recent years due to its significance in material and environmental science, as well as in chemical separation processes, where it plays a key role in the dispersion and emulsion of immiscible solvent compositions, and adsorption at solid surfaces. A detailed discussion of the forces acting on the “interfacial region” (as it was named in order to include contributions of the second or third layer below the actual surface layer) was given by Fowkes [1]. He explained the surface tension as a result of the attractive forces from the underlying molecules net perpendicular to the surface, which causes a reduction in the number of molecules on the surface, which again increases their intermolecular distance. This increase requires work, which is the intrinsic reason for the tension, and is expressed as surface free energy upon its relaxation. The intermolecular attractive forces are separable into the London dispersive, polar, and Lewis acid-base forces, of which the former two are additive and the latter is non-additive, as has been outlined by van Oss et al. [2]. Based on these findings, Freitas et al. [3] developed a linear free energy relationship (LSER) equation, which—besides a constant—encompasses parameters representing the dipolarity, the excess molar refraction, the hydrogen bond acidity and basicity, and the molar volume of an organic liquid. They concluded from the weights of these parameters in the LSER equation that the dominant factors contributing to the surface tension are the dipolarity, the excess molar refraction, and the constant. The latter was interpreted as representing the loss of the dispersive interaction which any molecule suffers upon transfer to the surface. A further term, Nc, was introduced, restricted to n-alkanes, in order to take account of the anisotropic properties found with linear alkanes larger than n-hexane, which was termed by Fowkes [4] as correlated molecular orientation (CMO), leading to enhanced adhesion on the liquid surface. The final result of the LSER equation for 299 compounds yielded a correlation coefficient of 0.936 and a standard error of 2.16 dyn/cm. A recent publication [5] has shown, with a homologous series of n-alkyl-substituted ionic liquids, that this CMO behaviour is not limited to the n-alkanes alone; in fact, the present paper will reveal further examples of normal compounds exhibiting anisotropism on the surface.
A number of further concepts have been developed, targeting a simple method for the reliable prediction of the surface tension and its temperature dependence. They essentially have in common that they do not claim to provide an explanation as to what kinds of intermolecular interactions are effective, and what influences the molecular orientations may exhibit at the surface of the liquids. In 1923, MacLeod [6], and later Sugden [7], observed that the fourth root of the experimental values of the surface tension correlated well with the product of the density difference between the liquid and the vapour state and a term called parachor, which has a nearly constant ratio to the critical volume. The fact that the parachor of a compound can be approximated as the sum of its constituting atomic and structural parachors enabled a direct estimation of the surface tension from the molecular structure. Sastri and Rao [8] found a correlation of the surface tension at the compound’s boiling point with the critical pressure, critical temperature, and the reduced boiling point, and extended this value to other temperatures. While in the previous examples the surface tension was evaluated from other known physical descriptors, a molecular mechanics force-field approach simulating the liquid state was applied, based on the molecule’s partial charges [9], yielding—besides numerous further properties—for 155 compounds (in the case of the surface tension) a root mean square deviation of 7.3 dyn/cm and a correlation coefficient of 0.89. An interesting concept applied a combination of an artificial neural network (ANN) with a group contribution method, in that 151 pre-defined chemical functional groups have been used as input and experimental surface-tension data as target on the output side in a three-layer neural network [10]. After training the ANN with 752 compounds at various temperatures and pressures, the result for the test set was surprisingly good, as it yielded an absolute average deviation of only 1.7% and a correlation coefficient of 0.995. In another approach, a multiple linear regression model was used to predict the surface tension of a total of 166 of alkanes, esters, and alcohols [11], founded upon a set of 10 descriptors with the highest correlation coefficients with the experimental surface tension, selected from of a total of 145 topological, geometrical, and electronic molecular descriptors. Some descriptors which exhibited a nonlinear relationship on a graphical plot were “linearized” by means of a suitable mathematical transform. An interesting point to be mentioned in view of the approach of the present paper is that 6 of the 10 descriptors are of topologic, 2 of electronic, and 2 of H-bonding nature. The final result for the 146 “training” compounds yielded a correlation coefficient of 0.983 and a standard deviation of 0.4 dyn/cm. An analogous concept to the one being presented here rested upon the well-known linear atom-group contribution method [12]. Due to the small number of only 12 special functional groups, some of them representing extended fragments, it was limited to a fairly narrow structural range of compounds.
The availability of a large number of experimental surface-tension data from many sources made it appealing to try to apply the highly versatile atom-group contribution method described in [13] for the calculation of this molecular descriptor, as this method has proven its outstanding success in the prediction of numerous thermodynamic [13,14], as well as solubility- [13,14,15], optics- [13], charge- [13] and environment-related [13], and physical [15] descriptors of an enormously diverse scope of molecular structures, without the requirement of any modification of the basic calculation algorithm. The goal was to provide a simple yet reliable means to predict the surface tension of a molecule, easily extendable to any kind of compound, e.g., also including the ionic liquids, of which this property is of particular importance in connection with their extraction and solvation capability for a large range of solutes.

2. General Procedure

The experimental surface-tension data are stored, together with the molecules in their 3D-geometry-optimized structure and further experimental and calculated descriptors, in an object-oriented knowledge database, at present encompassing more than 31,000 records of pharmaceuticals, plant protection products, dyes, ionic liquids, liquid crystals, metal-organics, lab intermediates, and more.
The details of the present atom-groups additivity method has been outlined in [13]. While the definitions and meanings of the atom groups in the following group-parameters table (Table 2) are to be interpreted in the same way as exemplified in Table 1 of [13], the inclusion of the ionic liquids required the addition of a number of further atom groups in order to represent their charged moieties, analogous to those given for the calculation of their viscosity in [15]. The exemplary list of these additional atom groups is collected in Table 1. These groups are treated just like the remaining ones by the computer algorithm.
For practical reasons—and following chemical conventions—the ion charges of the ionic liquids are centred on the atom types of the atom groups in Table 1 and Table 2. A certain deviation from this convention has been made for the imidazolium cations, where the conventional notation would imply an asymmetrcal charge distribution which, as e.g., the EHMO calculations indicate (visualized in Figure 1 in [15]), is not the case. Therefore, in this case, the positive charge has been positioned onto the carbon atom at position 2 between the two nitrogen atoms, which are bound to this central carbon atom by aromatic bonds. Accordingly, the carbon atom at position 2 and the nitrogen atoms in the imidazolium ions are represented in Table 1 by the atom groups 7 and 14, respectively. Atom types representing atoms that are immediate neighbours of charged atoms are distinguishable from those without charged neighbours by the added sign (in brackets) in their associated “Neigbours” definition (see examples 4, 6, 8–11, 14, 18, 19, 23, and 24 in Table 1).
The computer algorithm evaluating the atom-group parameters first collects from the database those molecules which fulfil the conditions for their inclusion into the parameters calculation, i.e., it checks the availability of an experimental surface-tension value and ensures that all atom groups in the molecule are present in the group-parameters table, and then carries out the parameters calculation using a fast Gauss-Seidel matrix-diagonalization procedure. Details of this entire algorithm have been outlined in [13]. Once the group parameters have been generated and stored in the parameters table, an immediate test of its predictive quality is carried out, first including all the compounds in the parameters evaluation, followed by a 10-fold cross-validation plausibility test, ensuring that each of the compounds has been introduced alternatively as both a test or training sample, as has been described in detail in Section 2.4 of [13]. These cross-validation calculations—and all the subsequent predictive descriptor calculations—are carried out using Equation (1), where ST is the surface-tension value, ai and bj are the contributions, Ai is the number of occurrences of the ith atom group, and Bj is the number of occurrences of the jth special group, and C is a constant. Yet, there is one further restriction beyond the ones mentioned above, in that for the predictive calculations of the surface tension of the training and test compounds, only those atom groups in the parameters table are considered valid which have been represented in the preceding parametrization process by at least three independent compounds with a known experimental surface-tension value.
S T = i a i A i + j b j B j + C
The results of the parameters evaluations and cross-validation calculations are summed up at the bottom of Table 2 (rows A to H). The rightmost column lists the number of compounds representing the respective atom group. For several atom groups, this number falls short of the required number to render the group valid. Although these atom groups are not applicable for surface-tension predictions, they have been left in the parameters table for potential future use in this continuous project (and may motivate scientists working in this area to focus on compounds carrying the corresponding atom groups). The calculations are generally restricted to molecules containing the elements H, B, C, N, O, P, S, Si, and/or halogen.
The simple example of anisole (methyl phenyl ether) may help in understanding the use of equation 1 and Table 2: Anisole contains the following atom groups (n × “atom type/neighbours”: Contribution): 1 × “C sp3/H3O”: 3.14; 1 × “O/C2(pi)”: −1.00; 1 × “C aromatic/:C2O”: 3.7; 5 × “C aromatic/H:C2”: 1.01. The sum of the contributions of these atom groups is added to that of the constant “Const” (24.34): 3.14 − 1.00 + 3.7 + (5 × 1.01) + 24.34 = 35.23 dyn/cm. The experimental value was published in [16] as 35.7 dyn/cm.

3. Results

Since the value of the surface tension is highly sensitive to the experimental temperature conditions, and since several authors applied different temperatures as their own standard, an overall temperature standard was required in order to ensure comparability. The decision to choose 293.15 K as standard resulted from the observation that the majority of the authors referred to this temperature, and that measurements of another molecular physical property, the liquid viscosity (see [15]), also rested upon this standard. Where possible, e.g., if experiments at a series of temperatures have been published, the experimental surface-tension value was either linearly inter- or extrapolated if necessary, provided that the experimental temperature conditions did not deviate too much from this standard. The most productive source for experimental surface-tension data for ordinary liquid compounds, Jasper’s comprehensive paper [16], collecting some 2200 data from the year 1874 until 1969, stated that besides the temperature, other aspects, such as the method of measurement, the purity of the compounds, and even the experience of the investigator, had a major influence on the accuracy of the data. Unfortunately, he did not elaborate on the extent of the data uncertainty resulting from these aspects. This collection has been complemented—and its data compared—by the more recent collective papers [11,12,17,18]. Additionally, surface-tension data have been provided for various alkanes [19,20,21,22,23,24], alkylbenzenes [25], haloalkanes [26,27], halogenated esters and ethers [28], sulfoxides [29], and siloxanes [30,31]. Of particular interest are surface-tension data for ionic liquids. A recent comprehensive collection of publications in the supplement of [32], accumulated for the development of a further method for the prediction of the surface tension, based on the density, molar mass, and anion type, provided the source of data for 222 ionic liquids which have been included in the present studies.
In Table 2, the result of the atom-group parameters calculations, based on 1895 molecules, has been collected, together with a summary of the statistics data at the bottom (rows A to H). Attempts to further improve the result, e.g., by the exclusion of one or both of the special groups “Alkane” and “Unsaturated HC” (olefins and aromatics), yielded slightly lower correlation coefficients and higher standard deviations.
According to the entries A to H in Table 2, 165 (of 221) atom and special groups are valid for predictive calculations, as they are based on at least three independent training molecules. Therefore, the result of the goodness of fit R2 of 0.9039 was based on 1833 of the 1893 training compounds, with a standard deviation σ of 1.99 dyn/cm. The average statistics data of the ten 10-fold cross-validation calculations (entries F–H) rested on a total of 1769 compounds, resulting in a cross-validated goodness of fit Q2 of 0.8823 and a standard deviation S of 2.16 dyn/cm. The standard deviations σ and S (entries D and H) have been calculated from the training set and the combined test sets of the cross-validation calculations, respectively, using the well-known Equation (2), where SD is the respective standard deviation, x the experimental, x ¯ the calculated surface tension of each molecule, and N the number of molecules. (The corresponding average deviations — entries C and G — are the sum of the absolute differences between the experimental and calculated surface tension of all involved compounds, divided by the number of these compounds. Since the standard deviation is more widely used in the examination of the reliability of predictive calculations, corresponding discussions in this paper refer to this value.)
S D = N ( x x ¯ ) 2 N 1
The excellent compliance, in most cases, between the black crosses of the training set with the affiliated red circles of the cross-validated set in Figure 1, as well as the close similarity of standard deviations R2 and Q2, confirm the applicability of the present surface-tension prediction method. The corresponding histogram in Figure 2 exhibits a fairly even Gaussian distribution for both the direct and the cross-validated deviations. A list of all the compounds used in this study, their experimental and calculated data and their 3D structures is available online in the supplementary material.
The relatively large standard deviation in relation to the overall data range, however, obscures the otherwise bright picture of the good correlation between the experimental and predicted surface-tension values, in that it hides three important observations. The first observation concerns the reliability of the experimental data for the ionic liquids, a point that has already been referred to in [32]. A typical example is 1,3-dimethylimidazolium bis(trifluoromethylsulfonyl)amide, for which in [5] a surface tension of 39 dyn/cm was given at 298.3 K, whereas in [33] a value 36.3 dyn/cm at the same temperature was published. In a further example, the surface tensions of each of the complete series of 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)amide (with alkyl being ethyl to decyl) varied by ca. 1.5 to 2.1 dyn/cm at 298.15 K between the two publications [33] and [34]. Due to their hygroscopicity and high viscosity, a higher uncertainty, and thus scatter, of the experimental values should be expected, as is reflected in Figure 3. As a further consequence, the number of ionic liquid outliers, i.e., compounds for which the values exceed three times the cross-validated standard deviation, are disproportionately higher (26.6%) than the 4.8% for the normal compounds (see the outliers list in the supplementary material). The small number of ionic liquids compared with the complete set of compounds, however, did not impede them from remaining included in the parameters calculations without undue deterioration of the result.
The second observation, disguised behind the range of the standard deviation, reveals an important shortcoming of the present prediction method. A small set of compound classes, characterized by the common feature of carrying a homologous sequence of linear methylene chains, exhibits an unexpected deviation of the experimental sequence of surface-tension data from the calculated values, whereas other analogous classes of homologues show fairly normal correlation between experiment and prediction. Typical examples of the latter normal correlation sequence are n-alkylbenzenes [3] (chart a in Figure 4), methylesters of long-chain carboxylic acids [35] (chart 4b), 1-alkanols [36], and 1-alkylthiols [37], which only deviate from the ideal correlation by slightly differing slopes. In contrast to this, the sequence of the experimental surface-tensions in the homologous n-alkane series [38] (chart 4c) is nonlinear and seems to aim at a constant maximum with increasing chain length. Analogous nonlinearity with increasing chain length was found for 1-alkenes [38] and 1-bromoalkanes [39] (chart 4d). A nearly linear but inverse correlation was found for a methylene chain homologue substituted at both ends by a nitrate group [40] (chart 4e). This characteristic feature was also found for the homologues of α,o-dibromo-n-alkane [41] and 1,4-Bis(n-alkylcarbonyloxy)-2-butyne [42] (chart 4f). Quite a bizarre surface-tension sequence was revealed by the symmetrical (chart a in Figure 5) and asymmetrical (chart 5b) homologues of the ionic liquids 1,3-Bis(n-alkyl)imidazolium and 1-n-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)amide [5,33,34], respectively. It is obvious that the present atom-group additivity approach is not able to treat these highly heterogeneous sequences. The reason behind these deviations has been described by Fowkes [4] as a result of anisotropism on the liquid surface caused by the extensive molecular directional orientation of these compounds, leading to a correlated molecular orientation (CMO). However, Fowkes only related his CMO thesis to linear n-alkanes; its extension to compounds with various substitutions inside or at the end of the methylene chains remains open to further studies. Since the CMO effect is generally small in relation to the other attractive forces on the liquid surface—Fowkes evaluated a range of between 0 for hexane and 2.89 dyn/cm for hexadecane, i.e., ca. 10% of the total force for the largest n-alkane in the series—the maximum range of the surface tension of all these homologous series remained within the deviation limits to allow all of their members to stay included in the parameters calculation. As a consequence, however, the present atom-group additivity method at best provides an average value for the surface tension of these homologues.
The third observation, another form of special intermolecular association, was apparent on comparing the experimental surface tension of di- and tri-hydroxy-group-containing compounds with their calculated value, as these systematically by far underestimated the measured values. (An analogous observation was made for hydrazine, ethanolamine, propanolamine, 2-(isopropylamino)ethanol, and ethylenediamine.) Evidently, the excessive increase of the experimental surface tension is caused by an effect that is not captured by the ordinary hydroxy-group parameter (entry 157 in Table 2) and is most probably best described as additional associative intermolecular H-O bond forces. Therefore, a special group (entry 219 in Table 2) has been introduced to take account of the surplus effect of each additional hydroxy group, which indeed improved conformance with the experimental values. Nevertheless, due to the large scatter of the experimental values, which did not indicate any systematic correlation with the corresponding molecular structure—compare, e.g., the experimental surface tensions of the two closely related outliers 1,2,3-propanetriol and 1,2,6-hexanetriol showing values of 63.3 and 44.14 dyn/cm [12], respectively, and on the other hand those of the two structurally very different compounds ethylene glycol and heptaethylene glycol exhibiting experimental values of 48.43 and 48.39 dyn/cm [16], respectively—11 of the 21 examples with available data still exceeded the deviation limits and had to remain in the outliers list.
Barring these special cases, the overwhelming majority of surface tension data have shown a normal statistical pattern in relation to the predictions, clearly proving the applicability of the present group-additivity approach. But how well does it compare with other published methods? Since to the best of our knowledge the present calculations are founded on the largest set of compounds with experimental surface-tension data, a direct comparison of their reliability with earlier papers, often focusing on only a limited number of closely related compounds, seems of little use. For instance, the most similar concept to the present group-additivity method, published in 2000 [12], yet only applying 12 functional groups, yielded a correlation coefficient R2 of 0.754, based on a training set of only 349 compounds of structurally limited extent. The correlation coefficient of 0.995 and average deviation of 1.7% of the ANN method [10] mentioned earlier, on the other hand, are surprisingly good—and questionable—insofar as for a number of compound examples the experimental values, which have been measured by various scientific groups, scatter by far more than 1.7%, as has been demonstrated in the comprehensive paper [16]. Beyond this, any prediction of a property by means of the ANN method is inevitably bound to the computer incorporating the trained artificial network. By contrast, the greatest advantage of the present approach lies in the fact that no computer is required: The prediction of the surface tension of a compound takes only a simple 2D drawing on a sheet of paper to help to find all the atom groups—and the parameters of Table 2 to sum up their contributions as exemplified at the bottom of Section 2. The large number of presently 165 valid atom groups in Table 2 enables the surface-tension prediction of a wide range of structurally varying molecules, which is evidenced by the surface-tension calculability of 55% of the currently 31,212 compounds in ChemBrain’s database, which can be viewed as representative for the entire structural coverage of chemicals.

4. Conclusions

The present results prove the reliable applicability of the atom-group additivity approach on the molecular surface-tension prediction by simply extending, by a few further lines of control code, the common computer algorithm outlined in [13], which has already demonstrated its extraordinary versatility with the trustworthy prediction of 13 further descriptors described in the previous papers [13,14,15] in a split second in one single sweep on a desktop computer: The heats of combustion, formation (indirectly), solvation, sublimation and vaporization, the entropy of fusion, the partition coefficient logPo/w, the solubility logSwater, the refractivity, the polarizability, the toxicity against the protozoan Tetrahymena pyriformis, the liquid viscosity, and the activity coefficient at infinite dilution. In addition, the present method has the advantage of enabling an easily generalizable computer algorithm for the definition of the atom groups, i.e., the atom types and their neighbours. The present work is part of an ongoing project called ChemBrain IXL available from Neuronix Software (www.neuronix.ch, Rudolf Naef, Lupsingen, Switzerland).

Supplementary Materials

Supplementary materials are available online at https://www.mdpi.com/1420-3049/23/5/1224/s1. The list of compounds, their experimental and calculated data and 3D structures of the surface-tension calculations are available online under the names of “S1. Experimental and Calculated Surface-Tension Data Table.doc” and “S2. Compounds List of Surface-Tension Calculations.sdf”. A list of their outliers has been added under the name of “S3. Compounds List of Surface-Tension Outliers.xls”. The figures are available as tif files and the tables as doc files under the names given in the text.

Author Contributions

R.N. developed project ChemBrain and its software upon which this paper is based, and also fed the database, calculated and analysed the results and wrote the paper. W.E.A. suggested the extension of ChemBrain’s tools to include the presented descriptors and contributed the experimental data and the majority of the literature references. Beyond this, R.N. is deeply indebted to W.E.A. for the many valuable discussions.

Acknowledgments

R. Naef is indebted to the library of the University of Basel for allowing him full and free access to the electronic literature database.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fowkes, F.M. Attractive Forces at Interfaces. Ind. Eng. Chem. 1964, 12, 40–52. [Google Scholar] [CrossRef]
  2. Van Oss, C.J.; Good, R.J.; Chaudhury, M.K. Additive and Nonadditive Surface Tension Components and the Interpretation of Contact Angles. Langmuir 1988, 4, 884–891. [Google Scholar] [CrossRef]
  3. Freitas, A.A.; Quina, F.H.; Carroll, F.A. A Linear Free Energy Analysis of the Surface Tension of Organic Liquids. Langmuir 2000, 16, 6689–6692. [Google Scholar] [CrossRef]
  4. Fowkes, F.M. Surface Effects of Anisotropic London Dispersion Forces in n-Alkanes. J. Phys. Chem. 1980, 84, 510–512. [Google Scholar] [CrossRef]
  5. Almeida, H.F.D.; Freire, M.G.; Fernandes, A.M.; Lopes-da-Silva, J.A.; Morgado, P.; Shimizu, K.; Filipe, E.J.M.; Lopes, J.N.C.; Santos, L.M.N.B.F.; Coutinho, J.A.P. Cation Alkyl Side Chain Length and Symmetry Effects on the Surface Tension of Ionic Liquids. Langmuir 2014, 30, 6408–6418. [Google Scholar] [CrossRef] [PubMed]
  6. MacLeod, D.B. On A Relation between Surface Tension and Density. Trans. Faraday Soc. 1923, 19, 38–42. [Google Scholar] [CrossRef]
  7. Sugden, S. A Relation between Surface Tension, Density and Chemical Composition. J. Chem. Soc. 1924, 125, 32–41. [Google Scholar] [CrossRef]
  8. Sastri, S.R.S.; Rao, K.K. A simple method to predict surface tension of organic liquids. Chem. Eng. J. 1995, 59, 181–186. [Google Scholar] [CrossRef]
  9. Caleman, C.; van Maaren, P.J.; Hong, M.; Hub, J.S.; Costa, L.T.; van der Spoel, D. Force Field Benchmark of Organic Liquids: Density, Enthalpy of Vaporization, Heat Capacities, Surface Tension, Isothermal Compressibility, Volumetric Expansion Coefficient, and Dielectric Constant. J. Chem. Theory Comput. 2012, 8, 61–74. [Google Scholar] [CrossRef] [PubMed]
  10. Gharagheizi, F.; Eslamimanesh, A.; Mohammadi, A.H.; Richon, D. Use of Artificial Neural Network-Group Contribution Method to Determine Surface Tension of Pure Compounds. J. Chem. Eng. Data 2011, 56, 2587–2601. [Google Scholar] [CrossRef]
  11. Stanton, D.T.; Jurs, P.C. Computer-Assisted Study of the Relationship between Molecular Structure and Surface Tension of Organic Compounds. J. Chem. Inf. Comput. Sci. 1992, 32, 109–115. [Google Scholar] [CrossRef]
  12. Egemen, E.; Nirmalakhandan, N.; Trevizo, C. Predicting Surface Tension of Liquid Organic Solvents. Environ. Sci. Technol. 2000, 34, 2596–2600. [Google Scholar] [CrossRef]
  13. Naef, R. A Generally Applicable Computer Algorithm Based on the Group Additivity Method for the Calculation of Seven Molecular Descriptors: Heat of Combustion, LogPO/W, LogS, Refractivity, Polarizability, Toxicity and LogBB of Organic Compounds; Scope and Limits of Applicability. Molecules 2015, 20, 18279–18351. [Google Scholar] [PubMed]
  14. Naef, R.; Acree, W.E. Calculation of Five Thermodynamic Molecular Descriptors by Means of a General Computer Algorithm Based on the Group-Additivity Method: Standard Enthalpies of Vaporization, Sublimation and Solvation, and Entropy of Fusion of OrdinaryOrganic Molecules and Total Phase-Change Entropy of Liquid Crystals. Molecules 2017, 22, 1059. [Google Scholar] [CrossRef]
  15. Naef, R.; Acree, W.E. Application of a General Computer Algorithm Based on the Group-Additivity Method for the Calculation of two Molecular Descriptors at both Ends of Dilution: Liquid Viscosity and Activity Coefficient in Water at infinite Dilution. Molecules 2018, 23, 5. [Google Scholar] [CrossRef] [PubMed]
  16. Jasper, J.J. The Surface Tension of Pure Liquid Compounds. J. Phys. Chem. Ref. Data 1972, 1, 841–1009. [Google Scholar] [CrossRef]
  17. Körösi, G.; Kovats, E.S. Density and Surface Tension of 83 Organic Liquids. J. Chem. Eng. Data 1981, 26, 323–332. [Google Scholar] [CrossRef]
  18. Zubillaga, R.A.; Labastida, A.; Cruz, B.; Martinez, J.C.; Sanchez, E.; Alejandre, J. Surface Tension of Organic Liquids Using the OPLS/AA Force Field. J. Chem. Theory Comput. 2013, 9, 1611–1615. [Google Scholar] [CrossRef] [PubMed]
  19. Luning Prak, D.J.; Cowart, J.S.; Trulove, P.C. Density and Viscosity from 293.15 to 373.15 K, Speed of Sound and Bulk Modulus from 293.15 to 343.15 K, Surface Tension, and Flash Point of Binary Mixtures of Bicyclohexyl and 1,2,3,4-Tetrahydronaphthalene or Trans-decahydronaphthalene at 0.1 Mpa. J. Chem. Eng. Data 2016, 61, 650–661. [Google Scholar] [CrossRef]
  20. Luning Prak, D.J. Density, Viscosity, Speed of Sound, Bulk Modulus, Surface Tension, and Flash Point of Binary Mixtures of Butylcyclohexane with Toluene or n-Hexadecane. J. Chem. Eng. Data 2016, 61, 3595–3606. [Google Scholar] [CrossRef]
  21. Luning Prak, D.J.; Lee, B.G. Density, Viscosity, Speed of Sound, Bulk Modulus, Surface Tension, and Flash Point of Binary Mixtures of 1,2,3,4-Tetrahydronaphthalene and Trans-decahydronaphthalene. J. Chem. Eng. Data 2016, 61, 2371–2379. [Google Scholar] [CrossRef]
  22. Zhang, C.; Li, G.; Yue, L.; Guo, Y.; Fang, W. Densities, Viscosities, Refractive Indices, and Surface Tensions of Binary Mixtures of 2,2,4-Trimethylpentane with Several Alkylated Cyclohexanes from (293.15 to 343.15) K. J. Chem. Eng. Data 2015, 60, 2541–2548. [Google Scholar] [CrossRef]
  23. Luning Prak, D.J.; Lee, B.G.; Cowart, J.S.; Trulove, P.C. Density, Viscosity, Speed of Sound, Bulk Modulus, Surface Tension, and Flash Point of Binary Mixtures of Butylbenzene + Linear Alkanes (n-Decane, n-Dodecane, n-Tetradecane, n-Hexadecane, or n-Heptadecane) at 0.1 Mpa. J. Chem. Eng. Data 2017, 62, 169–187. [Google Scholar] [CrossRef]
  24. Luning Prak, D.J.; Jones, M.H.; Cowart, J.S.; Trulove, P.C. Density, Viscosity, Speed of Sound, Bulk Modulus, Surface Tension, and Flash Point of Binary Mixtures of 2,2,4,6,6-Pentamethylheptane and 2,2,4,4,6,8,8-Heptamethylnonane at (293.15 to 373.15) K and 0.1 MPa and Comparisons with Alcohol-to-Jet Fuel. J. Chem. Eng. Data 2015, 60, 1157–1165. [Google Scholar] [CrossRef]
  25. Luning Prak, D.J.; Luning Prak, P.J.; Cowart, J.S.; Trulove, P.C. Densities and Viscosities at 293.15−373.15 K, Speeds of Sound and Bulk Moduli at 293.15−333.15 K, Surface Tensions, and Flash Points of Binary Mixtures of n-Hexadecane and Alkylbenzenes at 0.1 Mpa. J. Chem. Eng. Data 2017, 62, 1673–1688. [Google Scholar] [CrossRef]
  26. Tahery, R. Surface Tension Measurements for Binary Mixtures of Chlorobenzene or Chlorocyclohexane + Tetrahydrofuran at 298.15 K. J. Solut. Chem. 2017, 46, 1152–1164. [Google Scholar] [CrossRef]
  27. Heide, R. The surface tension of HFC refrigerants and mixtures. Int. J. Refrig. 1997, 20, 496–503. [Google Scholar] [CrossRef]
  28. Faurote, P.D.; Henderson, C.M.; Murphy, C.M.; O’Rear, J.G.; Ravner, H. Partiallv Fluorinated Esters and Ethers as Temperature-Stable Liquids. Ind. Eng. Chem. 1956, 48, 445–454. [Google Scholar] [CrossRef]
  29. Markarian, S.A.; Terzyan, A.M. Surface and bulk behavior of (dialkylsulfoxides + carbon tetrachloride) mixtures. J. Chem. Thermodyn. 2009, 41, 1413–1418. [Google Scholar] [CrossRef]
  30. Romagosa, E.E.; Gaines, G.L., Jr. The Surface Tension of Hexamethyldisiloxane-Toluene Mixtures. J. Phys. Chem. 1969, 3150–3152. [Google Scholar] [CrossRef]
  31. Myers, R.S.; Clever, H.L. Surface Tension of Octamethylcyclotetrasiloxane and Hexamethyldisilazane and Their Solutions with Carbon Tetrachloride and n-Hexadecane. J. Chem. Eng. Data 1969, 14, 161–164. [Google Scholar] [CrossRef]
  32. Koller, T.M.; Steininger, C.; Rausch, M.H.; Fröba, A.P. A Simple Prediction Method for the Surface Tension of Ionic Liquids as a Function of Temperature. Int. J. Thermophys. 2017, 38, 167. [Google Scholar] [CrossRef]
  33. Kolbeck, C.; Lehmann, J.; Lovelock, K.R.J.; Cremer, T.; Paape, N.; Wasserscheid, P.; Fröba, A.P.; Maier, F.; Steinrück, H.-P. Density and Surface Tension of Ionic Liquids. J. Phys. Chem. B 2010, 114, 17025–17036. [Google Scholar] [CrossRef] [PubMed]
  34. Carvalho, P.J.; Freire, M.G.; Marrucho, I.M.; Queimada, A.J.; Coutinho, J.A.P. Surface Tensions for the 1-Alkyl-3-methylimidazolium Bis(trifluoromethylsulfonyl)imide Ionic Liquids. J. Chem. Eng. Data 2008, 53, 1346–1350. [Google Scholar] [CrossRef]
  35. Vogel, A.I. Physical Properties and Chemical Constitution. Part XIII. Aliphatic Carboxylic Esters. J. Chem. Soc. 1948, 624–644. [Google Scholar] [CrossRef]
  36. Vogel, A.I. Physical Properties and Chemical Constitution. Part XX. Aliphatic Alcohols and Acids. J. Chem. Soc. 1948, 1814–1819. [Google Scholar] [CrossRef]
  37. Vogel, A.I. Physical Properties and Chemical Constitution. Part XXI. Aliphatic Thiols, Sulphides, and Disulphides. J. Chem. Soc. 1948, 1820–1825. [Google Scholar] [CrossRef]
  38. Jasper, J.J.; Kring, E.V. The Isobaric Surface Tension and Thermodynamic Properties of the Surfaces of a Series of n-Alkanes, C5 to C18, 1-Alkenes, C6 to C16, and of n-Decylcyclopentane, n-Decylcyclohexane and n-Decylbenzene. J. Phys. Chem. 1955, 59, 1019–1021. [Google Scholar] [CrossRef]
  39. Vogel, A.I. Physical Properties and Chemical Constitution. Part VIII. Alkyl Chlorides, Bromides, and Iodides. J. Chem. Soc. 1943, 636–647. [Google Scholar] [CrossRef]
  40. Zimmer, M.F.; Dix, S.; Lawrence, A.R. Thermodynamics of liquid Surfaces. Surface Tension of a Homologous Series of α,ι- Alkane Dinitrate Esters. J. Chem. Eng. Data 1965, 10, 288–290. [Google Scholar] [CrossRef]
  41. Grzeskowiak, R.; Jeffrey, G.H.; Vogel, A.I. Physical Properties and Chemical Constitution. Part XXXI. Polymethylene Dichlorides, Dibromides, Di-iodides, and Dicyanides. J. Chem. Soc 1960, 4728–4731. [Google Scholar] [CrossRef]
  42. Grzeskowiak, R.; Jeffrey, G.H.; Vogel, A.I. Physical Properties and Chemical Constitution. Part XXIX. Acetylenic Compounds. J. Chem. Soc. 1960, 4719–4722. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. Correlation diagram of the surface-tension data (in dyn/cm). Cross-validation data are added as red circles. (N = 1833; R2 = 0.9039; Q2 = 0.8823; regression line: intercept = 2.8653; slope = 0.9049).
Figure 1. Correlation diagram of the surface-tension data (in dyn/cm). Cross-validation data are added as red circles. (N = 1833; R2 = 0.9039; Q2 = 0.8823; regression line: intercept = 2.8653; slope = 0.9049).
Molecules 23 01224 g001
Figure 2. Histogram of the surface-tension data. Deviations are in dyn/cm. Cross-validation data are superpositioned as red bars. (S = 2.16; experimental values range: 9.89—53.5 dyn/cm).
Figure 2. Histogram of the surface-tension data. Deviations are in dyn/cm. Cross-validation data are superpositioned as red bars. (S = 2.16; experimental values range: 9.89—53.5 dyn/cm).
Molecules 23 01224 g002
Figure 3. Correlation diagram of the surface-tension data of the ionic liquids (in dyn/cm). (N = 154; R2 = 0.8579).
Figure 3. Correlation diagram of the surface-tension data of the ionic liquids (in dyn/cm). (N = 154; R2 = 0.8579).
Molecules 23 01224 g003
Figure 4. Correlation diagrams of the surface tension (in dyn/cm) of ordinary organic liquids with linear n-alkyl chains.
Figure 4. Correlation diagrams of the surface tension (in dyn/cm) of ordinary organic liquids with linear n-alkyl chains.
Molecules 23 01224 g004
Figure 5. Correlation diagrams of the surface tension (in dyn/cm) of ionic liquids with linear n-alkyl chains.
Figure 5. Correlation diagrams of the surface tension (in dyn/cm) of ionic liquids with linear n-alkyl chains.
Molecules 23 01224 g005
Table 1. Atom-group examples for ionic liquids and their meaning.
Table 1. Atom-group examples for ionic liquids and their meaning.
NoAtom TypeNeighboursMeaningExample
1B(−)C4 B in tetracyanoborate
2B(−)CF3CBF3alkyltrifluoroborate
3B(−)F4BF4tetrafluoroborate
4C sp3H3B(−)H3CBC in methyltrifluoroborate
5C(−) sp3C3 central C in tricyanocarbeniate
6C aromaticH:C:N(+)C:CH:N+C2 in pyridinium
7C(+) aromaticC:N2N:C(C):NC2 in 2-alkylimidazolium
8C spC#N(−)C(C#N)cyano-C in tricyanocarbeniate
9C spN#N(−)N(C#N)C in dicyanoamide
10C spB#N(−)B(C#N)C in tetracyanoborate
11C sp=N=S(−)N=C=Sthiocyanate
12N(+) sp3C4N+C4tetraalkylammonium
13N(+) sp2O2=O(−)NO3nitrate
14N aromaticC2:C(+)C-N(C):C+N1 in 1-alkylimidazolium
15N(+) aromaticC:C2C:N+(C):CN in 1-alkylpyridinium
16N(−)C2C-N-CN in dicyanoamide
17N(−)S2S-N-Sbis(trifluoromethanesulfonyl)amide
18P4CO2=O(−)CPO3alkylphosphonate
19P4O3=O(−)O=PO3dialkylphosphate
20P(+)C4PC4+tetraalkylphosphonium
21P(−)C3F3F3PC3tris(pentafluoroethyl)trifluorophosphate
22P(−)F6PF6hexafluorophosphate
23S4CO=O2(−)CSO3alkylsulfonate
24S4O2=O2(−)SO4alkylsulfate
Table 2. Atom groups and their contributions for surface-tension calculations.
Table 2. Atom groups and their contributions for surface-tension calculations.
EntryAtom TypeNeighboursContributionOccurrencesMolecules
1Const 24.3418931893
2BC3−4.511
3BO30.666
4B(−)C417.4955
5B(−)CF31.1155
6B(−)F41.671010
7C sp3H3B(−)1.2111
8C sp3H3C−2.2830481529
9C sp3H3C(+)29.4533
10C sp3H3N7.1111104
11C sp3H3N(+)8.273423
12C sp3H3O3.14194155
13C sp3H3S4.711311
14C sp3H3S(+)4.6611
15C sp3H3P5.4622
16C sp3H3Si−0.6511318
17C sp3H2BC2.2831
18C sp3H2BC(−)−6.7833
19C sp3H2C20.263591366
20C sp3H2C2(+)2.0477
21C sp3H2CN6.38271175
22C sp3H2CN(+)6.137446
23C sp3H2CO2.921277673
24C sp3H2CP7.041615
25C sp3H2CP(+)2.87246
26C sp3H2CS4.577655
27C sp3H2CS(+)6.6452
28C sp3H2CSi3.91125
29C sp3H2CF−1.6555
30C sp3H2CCl4.935139
31C sp3H2CBr6.414133
32C sp3H2CJ9.032821
33C sp3H2O26.1877
34C sp3HC31.56428304
35C sp3HC2N7.611512
36C sp3HC2N(+)8.466
37C sp3HC2O5.08150116
38C sp3HC2S5.6585
39C sp3HC2Cl4.681212
40C sp3HC2Br6.371010
41C sp3HC2J9.4644
42C sp3HCO27.11513
43C sp3HCF2−1.622917
44C sp3HCCl25.351211
45C sp3HCBr212.0532
46C sp3HO39.8733
47C sp3C43.1111104
48C sp3C3N4.5322
49C sp3C3O6.423636
50C sp3C3S4.7533
51C sp3C3F2.232111
52C sp3C3Cl7.324233
53C sp3C3Br3.7311
54C sp3C2F2−0.433357
55C sp3C2Cl2−0.1422
56C sp3CNF28.2693
57C sp3COF22.993418
58C sp3CF3−4.9811149
59C sp3CSF20.0521
60C sp3CF2Cl−0.721
61C sp3CPF2(−)2.753311
62C sp3CFCl2−0.7411
63C sp3CCl34.6109
64C sp3N3F(+)−4.311
65C sp3SF3−2.3111057
66C(−) sp3C39.3333
67C sp2H2=C−2.437877
68C sp2HB=C(−)2.7411
69C sp2HC=C1266174
70C sp2HC=O2.741313
71C sp2H=CN2.75216108
72C sp2H=CO0.2599
73C sp2H=CS4.093230
74C sp2H=CCl0.853
75C sp2H=CBr−1.8511
76C sp2HN=O10.3922
77C sp2HO=O1.281313
78C sp2C2=C3.156756
79C sp2C2=N5.543529
80C sp2C2=O6.27372
81C sp2C=CO1.7133
82C sp2C=CS5.292524
83C sp2C=CCl3.3395
84C sp2C=CBr7.4733
85C sp2CN=O7.2522
86C sp2CO=O2.25737528
87C sp2CO=O(−)−2.992323
88C sp2=COS7.3122
89C sp2C=OCl7.5311
90C sp2C=OBr11.4211
91C sp2=CSCl7.4332
92C sp2=CSBr10.3732
93C sp2=CSJ15.6911
94C sp2=CCl22.7464
95C sp2NO=O6.0674
96C sp2O2=O2.641212
97C sp2OS=S9.1855
98C aromaticH:C21.011614344
99C aromaticH:C:N4.0110663
100C aromaticH:C:N(+)8.323318
101C aromaticH:N22.2311
102C aromatic:C31.6511960
103C aromaticC:C22.25313254
104C aromaticC:C:N5.492120
105C aromaticC:C:N(+)13.4433
106C aromatic:C2N6.231919
107C aromatic:C2N(+)10.031010
108C aromatic:C2:N9.4511
109C aromatic:C2O3.73229
110C aromatic:C2S6.7499
111C aromatic:C2Si3.9843
112C aromatic:C2F−0.498
113C aromatic:C2Cl3.952117
114C aromatic:C2Br7.0944
115C aromatic:C2J9.6933
116C(+) aromaticH:N20.96104104
117C(+) aromaticC:N2−22.731010
118C spH#C1.52424
119C spB#N(−)−4.57205
120C spC#C1.745640
121C spC#N5.96362
122C spC#N(−)−2.1593
123C spN#N(−)0.69168
124C sp#NS9.6544
125C sp=N=S4.8744
126C sp=N=S(−)4.6877
127N sp3H2C−3.522828
128N sp3H2C(pi)4.3666
129N sp3H2N0.5655
130N sp3HC2−9.482020
131N sp3HC2(pi)−4.777
132N sp3HC2(2pi)5.7911
133N sp3HCN(pi)9.3922
134N sp3HSi2−2.2911
135N sp3C3−14.791818
136N sp3C3(pi)−11.877
137N sp3C2N−7.9244
138N sp3C2N(pi)0.5666
139N sp3C2N(2pi)4.4222
140N(+) sp3HC3−5.9511
141N(+) sp3C4−5.842121
142N aromaticHC:C(+)9.333
143N aromatic:C2−2.386463
144N aromaticC2:C(+)0.45225114
145N aromatic:C:N7.0521
146N(+) aromaticC:C2−2.461818
147N sp2C=C044
148N sp2=CN−0.22126
149N sp2C=N−1.9463
150N sp2=CO0.082323
151N sp2N=O−2.4799
152N sp2O=O3.0333
153N(+) sp2CO=O(−)1.72220
154N(+) sp2O2=O(−)5.72313
155N(−)C2−0.3688
156N(−)S2−6.685454
157OHC0.58161150
158OHC(pi)1.079090
159OHN(pi)3.266
160OHO25.7721
161OHP0.31137
162OHS933
163OBC−1.36186
164OC2−4.02288181
165OC2(pi)−1725561
166OC2(2pi)4.8155
167OCN(pi)−2.671616
168OCN(+)(pi)−0.782212
169OCN(2pi)5.8844
170OCP−0.4116875
171OCP(pi)−3.2531
172OCS1.663523
173OCSi−3.09235
174OSi21.66309
175P3O3−1.632020
176P4HO2=O2.981616
177P4C2O=O(−)−9.2411
178P4CO2=O−2.11515
179P4CO2=O(−)0.3411
180P4O3=O3.171313
181P4O3=O(−)−6.4666
182P4O2=OF0.5222
183P4O2=OCl5.4722
184P4O=OCl28.0322
185P(−)C3F3−4.021111
186P(−)F6−4.9266
187P(+)C40.3266
188S2HC−1.341111
189S2HC(pi)2.8911
190S2C2−2.442020
191S2C2(pi)−3.021515
192S2C2(2pi)−3.493333
193S2CS0.422012
194S4C2=O5.933
195S4CN=O2(−)0.0710854
196S4CO=O29.3433
197S4CO=O2(−)−5.0155
198S4C=O2F1.5211
199S4C=O2Cl7.2455
200S4O2=O−0.6588
201S4O2=O23.4244
202S4O2=O2(−)−6.4477
203S4O=O2S−0.444
204S(+)C38.9622
205SiHC3−8.5311
206SiHC2Cl−5.3111
207SiHCCl2−4.2711
208SiHO34.2411
209SiC4−7.9144
210SiC3N021
211SiC3O−3.02147
212SiC3Cl−4.6311
213SiC3Br−2.8133
214SiC2O2−0.04216
215SiC2Cl2−2.9411
216SiC2Br21.9811
217SiCCl3−3.3911
218SiO47.8464
219(COH)nCOH groups: n > 13.261110
220AlkaneNo of C atoms0.221263125
221Unsaturated HCNo of C atoms0.021314125
ABased onValid groups165 1893
BGoodness of fitR20.9039 1833
CDeviationAverage1.53 1833
DDeviationStandard1.99 1833
EK-fold cvK10 1769
FGoodness of fitQ20.8823 1769
GDeviationAverage (cv)1.66 1769
HDeviationStandard (cv)2.16 1769

Share and Cite

MDPI and ACS Style

Naef, R.; Acree, W.E. Calculation of the Surface Tension of Ordinary Organic and Ionic Liquids by Means of a Generally Applicable Computer Algorithm Based on the Group-Additivity Method. Molecules 2018, 23, 1224. https://doi.org/10.3390/molecules23051224

AMA Style

Naef R, Acree WE. Calculation of the Surface Tension of Ordinary Organic and Ionic Liquids by Means of a Generally Applicable Computer Algorithm Based on the Group-Additivity Method. Molecules. 2018; 23(5):1224. https://doi.org/10.3390/molecules23051224

Chicago/Turabian Style

Naef, Rudolf, and William E. Acree. 2018. "Calculation of the Surface Tension of Ordinary Organic and Ionic Liquids by Means of a Generally Applicable Computer Algorithm Based on the Group-Additivity Method" Molecules 23, no. 5: 1224. https://doi.org/10.3390/molecules23051224

Article Metrics

Back to TopTop