Next Article in Journal
Magnetic Solid-Phase Extraction of Organic Compounds Based on Graphene Oxide Nanocomposites
Previous Article in Journal
Zinc and Calcium Cations Combination in the Production of Floating Alginate Beads as Prednisolone Delivery Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Calculation of the Isobaric Heat Capacities of the Liquid and Solid Phase of Organic Compounds at 298.15K by Means of the Group-Additivity Method

Department of Chemistry, University of Basel, 4003 Basel, Switzerland
Molecules 2020, 25(5), 1147; https://doi.org/10.3390/molecules25051147
Submission received: 5 February 2020 / Revised: 1 March 2020 / Accepted: 2 March 2020 / Published: 4 March 2020

Abstract

:
The calculation of the isobaric heat capacities of the liquid and solid phase of molecules at 298.15 K is presented, applying a universal computer algorithm based on the atom-groups additivity method, using refined atom groups. The atom groups are defined as the molecules’ constituting atoms and their immediate neighbourhood. In addition, the hydroxy group of alcohols are further subdivided to take account of the different intermolecular interactions of primary, secondary, and tertiary alcohols. The evaluation of the groups’ contributions has been carried out by solving a matrix of simultaneous linear equations by means of the iterative Gauss–Seidel balancing calculus using experimental data from literature. Plausibility has been tested immediately after each fitting calculation using a 10-fold cross-validation procedure. For the heat capacity of liquids, the respective goodness of fit of the direct (r2) and the cross-validation calculations (q2) of 0.998 and 0.9975, and the respective standard deviations of 8.24 and 9.19 J/mol/K, together with a mean absolute percentage deviation (MAPD) of 2.66%, based on the experimental data of 1111 compounds, proves the excellent predictive applicability of the present method. The statistical values for the heat capacity of solids are only slightly inferior: for r2 and q2, the respective values are 0.9915 and 0.9874, the respective standard deviations are 12.21 and 14.23 J/mol/K, and the MAPD is 4.74%, based on 734 solids. The predicted heat capacities for a series of liquid and solid compounds have been directly compared to those received by a complementary method based on the "true" molecular volume and their deviations have been elucidated.

1. Introduction

Most experimental measurements of thermodynamic properties, such as vaporization, sublimation, solvation, or fusion enthalpies, are usually carried out at temperatures that differ from the standard temperature, which has generally been accepted as being 298.15 K. These temperature differences lead to experimental values for the temperature-dependent properties that prevent a direct comparison of the results between various compounds or between scientific teams examining the same molecule, a deficiency which, however, can be corrected, provided that the heat capacity of the molecules under examination is known. Instead of measuring this property for a specific molecule, the large amount of experimental heat-capacity data for all kinds of compounds, such as inorganic and organic salts, liquid crystals, or ionic liquids, enabled its prediction by means of a large number of mathematical methods, a comprehensive overview of which has been given in a recent publication by the present author [1]. The majority of these prediction methods are based on the group-additivity (GA) approach, whereby the group notations vary from complete polyatomic ions as, e.g., applied by Gardas and Coutinho [2] to single atoms and their immediate neighbour atoms and ligands, as described by Benson and Buss [3]. Generally, the GA methods’ range of applicability for the prediction of any kind of descriptors varies over a large scope of molecular structures, depending on the complexity and number of the group notations as well as the number of experimental data upon which the group parameters are based. Similarly, the reliability of the predictions is highly dependent on the range of application. Zàbransky and Ruzicka [4], e.g., defined 130 functional groups including cis, trans, as well as ortho and meta corrections in the parametrization of their second-order polynomial GA model for the prediction of the liquid heat capacity and its temperature dependence, based on more than 1800 experimental data points. For the majority of compounds they reported an average deviation of below 2%. For alkanols, acids and aldehydes, however, the error was larger than 3% and rose with increasing temperature. A further limit to the use of their model was the observation that the prediction accuracy deteriorated further if the compounds contained functional groups from different families, such as N,N’-diethanolamine or 1-chloro-2-propanol. Another example of a GA method, provided by Chickos et al. [5], used 47 functional groups for the prediction calculation of the heat capacity of 810 liquids and 446 solids, reporting standard errors of 19.5 J/molK for the liquids and 26.9 J/mol/K for the solids. The authors compared these errors with the experimental uncertainties of 8.12 and 23.4 J/mol/K, respectively, which they estimated from the experimental data variations for each of 219 liquids and 102 solids published by independent sources.
A common deficiency of the GA and all the other approaches cited in [1] is that none of them enables the prediction of any specific descriptor for each and any molecular structure in the chemical realm. In the case of the heat capacities of the solid and liquid phase of molecules, however, this deficiency has been overcome in that their prediction values are determined via the "true" molecular volume (Vm) outlined in detail in [1]. Nevertheless, this approach has encountered several other shortcomings which could not all be addressed specifically, as it is based on one single number, the molecular volume. The three most important deficiencies are 1) the general influence of the hydroxy group of alcohols and carboxylic acids, 2) the specific effects of primary, secondary, and tertiary alcohols and 3) the impact of saturated cyclic rings vs. open-chained systems on the heat capacities. Accordingly, a first attempt of a linear correlation calculation in [1], which included the molecular volume and the experimental liquid heat capacity Cp(liq) of the complete set of compounds, for which both data were available, and which neglected the mentioned shortcomings, yielded a rather large standard deviation of 27.84 J/mol/K and a mean absolute percentage deviation (MAPD) of 8.23%. The neglect of the hydroxy-group effect on Cp(liq) was immediately manifest in that the predicted values for all those compounds carrying at least one OH group were systematically well below the experimental ones by up to ca. 130 J/mol/K. This general deviation, obviously caused by the formation of intermolecular hydrogen bridges between the OH groups, has been considered in subsequent calculations in that the complete set of compounds was separated by means of a few simple steps in the computer algorithm into three subsets, i.e., one encompassing all molecules lacking any OH group, a second one consisting of those carrying one OH group and a third one comprising those having more than one OH group. For each of these subsets, a separate linear correlation calculation had to be carried out yielding three sets of linear parameters for the prediction of the liquid and three for that of the solid heat capacities. In this way, the first one of the mentioned shortcomings has been eliminated, which correspondingly resulted in significantly better compliance of the predictions with the experimental data. The corresponding statistical results will be discussed and used for comparison in a later section. The remaining two deficiencies concerning the various alcohol classes as well as that of cyclic vs. open-chained structures in a saturated system, which exhibit a minor but still systematically negative influence on the prediction quality, has been plausibly explained, but a reasonably straightforward treatment within the context of the Vm method was not feasible.
Therefore, the question arose as to whether and how well a GA approach would overcome the remaining shortcomings of the Vm method and enable a more accurate and reliable prediction of the heat capacities of molecules in their liquid and solid phase at the standard temperature, in awareness of the disadvantage that it would not be able to cover each and every possible compound. A particularly versatile GA method, outlined in [6], enabling in a single sweep the calculation of 14 thermodynamic [6,7], solubility- [6,7,8], optics- [6], charge- [6], environment-related [6], and physical [8,9] properties of a nearly unlimited scope and size of molecular structures should best serve this purpose, all the more so as in most cases it in principle also opened a simple means for their reliable calculation on a sheet of paper. Accordingly, the present work puts a special focus on the effects of the hydroxy groups and the cyclization of saturated molecular parts on the heat capacities and how to deal with them. The statistical results of the present GA method will be put in relation to those of the Vm method but also to those of the GA approach of Chickos et al. [5], as this approach can be viewed as most closely related to the present one.

2. Method

The present study is founded on a project-owned and regularly updated, object-oriented database of more than 32000 molecules encompassing pharmaceuticals, plant protection, dyes, ionic liquids, liquid crystals, metal-organics, lab intermediates, and many more, all of which are stored as geometry-optimized 3-dimensional structures, including—besides several further descriptors—a set of 1176 experimental heat capacities of liquids and a corresponding set of 802 heat capacities of solids.
The details of the present atom-group additivity method and the evaluation of its group contributions have been outlined in an earlier paper [6]. Accordingly, its group notations have the same meaning as that exemplified in Table 1 of [6]. However, in order to include ionic liquids for which the experimental heat capacities are known, the list of group notations has been extended by ionic atom groups representing their charged fragments, as listed in the present Table 1. These special atom groups have already successfully been utilized in the calculation of the molecules’ viscosity [8] and surface tension [9], applied in the same way as the remaining groups. For the interpretation of the ionic atom groups of Table 1, the reader is invited to read section 2 of papers [8] and [9].
In the course of the first preliminary group-contribution calculations, whereby tentatively certain “standard” atom groups have been replaced by refined ones and special groups, which will be described in the following, have been added or omitted, their statistical results quickly revealed significant improvement of the predictive quality if the groups listed in Table 2 are included in the prediction of both the liquid and solid heat capacities.
In the discussion of the shortcomings of the molecular volume-based calculations of the heat capacities outlined in the introductory section, the hydroxy group appeared to be the most accountable group for large deviations between experimental and predicted heat-capacity values, even within the restricted set of OH-containing compounds, i.e., after their separation from the remaining ones. It turned out that the definition of the OH group on saturated carbon as in ordinary alcohols by the simple atom type “O” and its neighbours “HC” was inadequate for heat-capacity calculations, in contrast to the calculations of all the other descriptors mentioned in our earlier papers [6,7,8,9]. As a consequence, an additional procedure had to be integrated in the general GA algorithm outlined in [1], which redefined the atom type “O” into “O(prim)”, “O(sec)”, or “O(tert)”, depending on the number of carbon atoms attached to the C atom neighbouring the O atom, according to the definition of primary, secondary, and tertiary alcohols, as shown in Table 2. (Consequently, the definition of their neighbourhood “HC” was no longer relevant and was thus not examined.) This redefinition procedure is only invoked if the redefined atom types appear in the group-parameters table, as a consequence of the algorithmic procedure determining that it is the content of the group-parameter tables that defines which group parameters are to be evaluated for the corresponding descriptors calculations (as explained in subsection 2.2 of [1]), and since none of the other descriptors in [6,7,8,9] requires this redefinition, this procedure is only called up for the evaluation of the group parameters of present Table 3 and Table 7 and the subsequent heat-capacity predictions. The remaining hydroxy groups attached to unsaturated carbon found in carboxylic acids and phenols are notated separately by the atom type “O” and the neighbourhood “HC(pi)”, as defined in [6].
Another point of weakness discussed in the introductory section rested in the observation that the Vm approach systematically scored badly in the prediction of the heat capacities of molecules with cyclic saturated moieties. This deficiency has been resolved in the present GA method in that the endocyclic single bonds in a molecule are counted and their sum multiplied by the contribution value of the special group “Endocyclic bonds” to yield the effect of the cyclic moieties in a molecule on its heat capacity. The groups “Angle60”, “Angle90”, and “Angle102” serve as corrective elements for small rings. Not surprisingly, these special groups, which take account of an effect influencing the freedom of intramolecular motion, have also successfully been applied in the prediction of the entropy of fusion [7].
The special group “(COH)n” had to be introduced in the Cp calculations in order to compensate for deviations found for polyols and polyacids. This special group has played its useful part already in the calculation of the surface tension [9]. The test calculations also revealed a very strong influence of intramolecular hydrogen bonds on the liquid heat capacity, which had to be taken into account by the introduction of the special group “H/H Acceptor”, a group that has also been used successfully used in the prediction of the toxicity [6], the heats of solvation, and the sublimation, vaporization, and entropy of fusion [7].
The procedure for the evaluation of the atom-group contributions, as explained in [6], is identical for the two group-parameter sets for the prediction of the heat capacities of both the liquid and solid phases and may be summarized as follows: in a first step, a list of all the compounds, for which the experimental Cp values are known, is extracted from the database. In the next step, each “backbone” atom (i.e., each atom bound to at least two immediate neighbours) within each molecule has an atom type and its neighbourhood assigned to by means of two character strings defining an atom group, following the rules defined in [6] (e.g., “C sp3” and “H2CO” for the C1 atom in ethanol) and then this group’s occurrence in the molecule is counted. The list of M molecules and their N atom groups plus their experimental values are then entered into an M × (N + 1) matrix, wherein each matrix element (i,j) receives the number of occurrences of the jth atomic or special group in the ith molecule. The normalization of this matrix into an Ax = B matrix and its balancing by means of the Gauss–Seidel calculus, e.g., according to E. Hardtwig [10], yields the atom-group contributions. This mathematical approach is based on the assumption that the prediction value of a molecule’s descriptor in question can be evaluated by simply summing up all the group contributions in the molecule. For the evaluation of the heat capacities in this study, Equation 1 has been adopted, wherein Cp is the heat capacity at 298.15 K, ai and bj are the group contributions, Ai is the number of occurrences of the ith atom group, and Bj is the number of occurrences of the jth special group.
C p = i a i A i + j b j B j
The reliance of this procedure is immediately examined by a subsequent 10-fold cross-validation plausibility test, carried out in a way to ensure that each compound has been entered into the calculation as a test as well as a training sample. All the group contribution values and the statistical results of both the direct equalization and the cross-validation calculation of the liquid heat capacity Cp(liq,298) are then collected in Table 3 and for the solid heat capacity Cp(sol,298) in Table 7. However, for the evaluation of the statistical results, only those group contributions are considered as valid for use that have been represented by at least three independent molecules in the equalization calculation. The number of molecules responsible for the respective group contribution is listed in the rightmost column of Table 3 and Table 7. Evidently, for several atom groups, this number falls short of the validity requirement. Nevertheless, as this work is part of a continuous project, these groups have deliberately remained in the parameters’ tables for future use. They might also motivate readers working in this area to contribute corresponding experimental data. In order to achieve reliable contribution values for the atom and special groups, it was necessary to filter out compounds with Cp values that deviated too far from the predicted results. In the present work, the limit was defined as three times the cross-validated standard deviation q2. The corresponding outliers have been excluded from the parameters’ calculations and are collected in an outliers list. The present calculations are generally restricted to molecules containing the elements H, B, C, N, O, P, S, Si, and/or halogen.

3. Sources of Heat-Capacity Data

The present work is essentially based on the comprehensive list of experimental heat-capacities collected in [1], used to substantiate the feasibility of the Vm approach. However, a recent scan of the literature brought forth a number of further publications, which either confirmed previous data or even improved the conformance with prediction, but also enabled an extension of the applicability of the present GA approach. A number of new experimental Cp data have been published for saturated and unsaturated hydrocarbons, especially for bicyclo[2.2.2]octane and bicyclo[2.2.2]octene [11], 1-octyne and 4-octyne [12], biphenyl [13], benzo[b]fluoranthene, benzo[k]fluoranthene, indeno[1,2,3-cd]pyrene [14], and adamantane [15]. In addition, further Cp values have been found for the amines hexamethylenetetramine [15], tetra-N-phenylbenzidine and 4,4′-bis (N-carbazolyl)-1,1′-biphenyl [16], 1,3,5-triazine [17], the ethers 1,3,5-trioxane [17], diethylene glycol n-pentyl ether [18], triethylene glycol monopentyl ether [19] and diphenyl ether [20], several alcohols and aldehydes [21], derivatives of glycidol [22], carboxylic acids [23,24], aliphatic esters [24,25,26,27], benzoates [28], haloalkanes [29,30], haloaromatics [31,32], thio ethers, sulfones and sulfoxides [33,34], alkanolamines, [35,36,37] and nitriles [38] For several compounds, more recent publications have been found the following: 4-ethylmorpholine [39], methionine [40], theophylline, and caffeine [41], as well as for (-)-menthone, (+)-pulegone, and (-)-isopulegol [42]. Beyond these, experimental Cp values of a number of new compounds have been published: namely 3-amino-4-amidoximinofurazan [43], 3-fluoro-5-(3-pyridinyloxy)- benzenamine and N-[3-fluoro-5-(3-pyridinyloxy)phenyl]-N’-3-pyridinyl urea [44], 7-Methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene [45], 2-[(4-nitro-benzoyl)-hydrazone]-propionic acid [46], eugenol and (+)-carvone [47], indapamide [48], and the explosives EDNH and DNTA [49]. For some silicon-containing compounds, new Cp data have been found in [50]. Finally, a few further heat capacities data of ionic liquids [51,52,53] have been included in the present dataset.

4. Results

4.1. General Remarks

  • In the subsequent figures, the results of the cross-validation calculations have been superimposed in red over the training data drawn in black.
  • The complete lists of compounds with known heat capacities used in this study are available as SDF files in the Supplementary Materials, downloadable by external chemistry software. In addition, the Supplementary Materials provides the results lists containing the molecules’ names and experimental, training, and cross-validation data. Beyond this, the lists of outliers of both heat-capacity calculations are also available in the Supplementary Materials.

4.2. Heat Capacity of Liquids

In Table 3, the atom groups and their contribution for the prediction of the heat capacity of liquids are collected, together with the number of molecules and occurrences upon which each of them is based.
In rows A to H, at the bottom of the table, the statistical data of this table have been gathered. As shown in row A, the group contributions have been evaluated on the basis of 1176 compounds yielding the data for 211 atom groups, of which, however, only 134 are considered as valid, i.e., that are supported by at least three compounds. Accordingly, since only valid groups have been used for the statistical evaluations, the numbers of compounds entered in the calculations of the trained and cross-validated correlation coefficients (“goodness of fit”) r2 and q2 (rows B and F) are lower with 1111 and 1060, respectively. Both the standard deviations of the complete data set (row D) as well as that of the combined cross-validation sets (row H) reveal excellently low values (in J/mol/K), not only in relation to the large range of experimental values of between 81.92 (methanol) and 1849 J/mol/K (trimethylpropane trioleate), but particularly also in comparison with the standard error of 19.5 J/mol/K reported by Chickos et al. [5] for 810 liquids. The result is a very low scatter along the correlation line, as is shown in Figure 1. Accordingly, the error distributions of both the training and the cross-validated sets fairly well follow the Gaussian distribution function, as demonstrated in the histogram (Figure 2). The MAPD for the complete set of 1111 liquid compounds was 2.66%, clearly by far better than the 8.23% for the entire set of 1303 liquids resulting from the Vm method [1], and still much better in comparison with the 6.51% for the OH-free subset of 1102 liquids reported in Figure 2 of [1].
The distinctly better conformance of the predicted with the experimental Cp(liq) values in comparison with earlier literature references is essentially based on three primary reasons. The first one is the refinement of the molecules’ description itself by the most detailed classification of group notations, which is precluded on principal to the Vm method [1], but requires a large number of atom groups and consequently a large amount of experimental data for their parametrization. The second reason originated from an observation made in [1], namely that the heat capacities of primary, and less so, of secondary alcohols have notoriously been overestimated by the Vm approach. These systematic deviations can be seen in Table 4, where the experimental Cp(liq,298) data and the predicted values of both the present GA and Vm method of the corresponding alkanols, encompassing saturated alkyl mono-, di- and polyols, are compiled for comparison. In order to overcome this deficiency, the alcohols have therefore been subdivided as described in Section 2 into the three subclasses primary, secondary, and tertiary alcohols. This additional separation indeed had a dramatically positive effect on the entire alcohols class, demonstrated by the comparison of the correlation diagrams of Figure 3. The MAPD values shown at the bottom of Table 4 confirm that the GA method on average produces distinctly lower deviations from experimental values than the Vm approach.
A quick review of the contributions of the corresponding atom groups representing the primary, secondary, and tertiary alcohols (group numbers 166 to 168) in Table 3 reveals the large influence of the immediate neighbourhood of the OH group. Evidently, with its growing bulkiness, the contribution to the heat capacity of the OH group increases due to its progressively hampered accessibility to build a hydrogen bridge. This effect has been plausibly explained by Huelsekopf and Ludwig [54], who discovered, upon applying theoretical calculations based on the quantum cluster equilibrium theory (QCE) on two primary (ethanol and benzyl alcohol) and a tertiary alcohol (2,2-dimethyl-3-ethyl-3-pentanol), that primary alcohols on principle form cyclic tetramers and pentamers in the liquid phase, while tertiary alcohols under the same conditions only consist of monomers and dimers. Following this reasoning, the higher liquid heat capacity of secondary and tertiary alcohols over that of their primary counterparts having the same molecular formula is the result of their formation of smaller clusters, which inherently exhibit a higher number of rotational and translational degrees of freedom.
The third reason for the good compliance of the present Cp predictions with experimental values is the consideration of the cyclization effect in the present GA method. Table 5 presents a selection of some linear alkanes and their closely related cycloalkanes and compares their experimental Cp(liq) values with predicted data calculated by means of the present GA method and the Vm [1] approach. Scanning the table’s fifth column immediately reveals that the Vm approach systematically overestimates the liquid heat capacity of the cycloalkanes, whereas those of the linear alkanes are excellently well predicted. The reason is obvious: cyclization reduces the number of rotational degrees of freedom, an effect which is categorically excluded from consideration by the Vm method. The present GA method, however, includes this effect in that the number of endocyclic single bonds is counted and their count is multiplied by the assigned special group contribution, in this case by the value of −3.92 J/mol/K of group 212 in Table 3. The result of this inclusion is evident in column 3 of Table 5, proving that the overestimation of the Cp(liq) values of the cycloalkanes on average is completely lifted.
An exemplary implementation of these findings may be provided in the calculation according to Equation 1 of the Cp(liq,298) value of a cyclic alcohol, such as cyclohexanemethanol:
5 x [C sp3/H2C2] + [C sp3/HC3] + [C sp3/H2CO] + [O(prim)/HC] + 6 × [endocyclic bonds] = 5 × 30.06 + 21.11 + 73.79 + 14.41 + 6 × (−3.92) = 236.09 J/mol/K (experiment: 236.5 J/mol/K).
In this context, it is worth mentioning that Chickos et al. [5] took great care about the parametrization of the "cyclic tertiary sp3 carbons" (as they called them) and their neighbourhood, but only reserved a single atom group for all the alcohol classes including phenols.
Since, in recent years, the class of ionic liquids (IL) has received increasing interest as a group of new polar solvents, their heat capacity as an important property has come into focus. It was therefore interesting to examine how well the present GA method would cope in comparison with the Vm method of [1]. In Table 6, the experimental Cp(liq,298) data of 122 ILs have been collected and compared to the prediction data calculated by the present GA method and by the Vm method. A comparison of the MAPD values at the bottom of the table clearly demonstrate a substantial improvement of the present GA approach over the Vm method.
The present calculation of the atom-group parameters for the prediction of Cp(liq,298) revealed ca. 170 compounds with experimental values exceeding the deviation limit, as defined in Section 2, which have been removed from parameters calculations and are collected in an outliers list. A comparison of this list with that resulting from calculations by means of the Vm method [1] showed very high overlap, indicating that the exclusion of these compounds was indeed justified. After the removal of these outliers, a limited number of 1202 compounds with usable experimental data remained, supporting the contributions of 134 atom and special groups valid for prediction calculations, as is shown in row A of Table 3. Despite this fairly low number of atom groups, the range of applicability of the present GA method is considerably high: for nearly 62% of ChemBrain’s database of the more than 32000 compounds, the liquid heat capacity has been evaluable.

4.3. Heat Capacity of Solids

While the measurement of the heat capacity of liquids principally implies a consistent isotropic phase, the corresponding examination of solids very often faces the question as to what type of association the particular compound has adopted in its solid phase. Many compounds precipitate in various crystalline forms, depending on the precipitation conditions, each of them having a different heat capacity, and many of these can change from one into another crystalline structure upon measurement, perhaps even switching from one tautomeric form into another one. In some cases, the apparent solid is merely a supercooled melt. The uncertainty of the actual structure of the solids appears to be the main cause of the larger scatter of the heat capacities of solids Cp(sol,298) as compared to that of the liquids, not only over the complete range of available compounds but also over particular compounds examined by several independent sources, as has been observed by Chickos et al. [5]. These uncertainties are expressed in the statistics data at the bottom of Table 7, which presents the list of atom and special groups and their contribution for the prediction of the heat capacity of solids. Based on the Cp data of 804 solids, the Gauss–Seidel calculus yielded 126 atom and special groups (row A in Table 7) valid for prediction calculations and a cross-validation standard deviation Q2 of 14.23 J/mol/K (row H). This standard deviation is clearly higher than that for the calculation of the liquid heat capacities, but much lower than the 26.9 J/mol/K of Chickos’ method [5] and even lower than the experimental variation of 23.4 J/mol/K for each of the 102 solids originating from independent sources [5]. The MAPD value for the complete set of solids was calculated to 4.74%, which is better than that of each of the subsets of compounds calculated by means of the Vm method [1]. Nevertheless, as is demonstrated in the corresponding diagram (Figure 4), the scatter around the correlation line is significantly larger compared to the one of Figure 1 for the liquid heat capacity. Analogously, the histogram (Figure 5) shows a wider "waist" than that of Figure 2.
Hydrogen bridges are known to play a crucial role in the formation of the crystalline structure of solids (think of snowflakes or water ice). Since the Vm approach of [1] is not able to include this effect directly, compounds containing OH groups were treated separately from the OH-free molecules. In analogy to the observation made with the liquid alcohols one would then have expected that the Vm approach again exhibited an unresolvable deficiency as concerning the deviations between experimental and predicted solid heat capacities of primary, secondary, and tertiary alcohols. Unfortunately, however, the enhanced extent of the scatter of the experimental Cp(sol) values in this compound’s class concealed these suspected deviations. The present GA method on the other hand provided an indirect proof of the influence of the immediate neighbourhood of the OH group in the alcohol subclasses: a comparison of the contributions of the atom groups 157, 158, and 159 in Table 7 (−23.36, −16.25 and −3.34 J/molK, respectively), assigned to the primary, secondary, and tertiary OH groups, immediately reveals that the primary alcohols exhibit the strongest hydrogen bridge effect, leading to the correspondingly largest decrease in the heat capacity due to the additional loss of freedoms of motion, followed by the secondary and the tertiary alcohols. The reason for this differentiation is the same as explained for that of the liquid alcohols in the prior section: the increase in the bulkiness around the OH group increasingly prevents hydrogen-bridge building. The separation of the alcohol subclasses in the present GA method also improved the reliability of the predicted Cp(sol,298) values. In Table 8, the results of the GA and the Vm method for 31 alkanols have been collected and compared with their experimental data. It is interesting to see that the largest deviations of the GA method coincide with large ones of the Vm method (i.e. for 2,2-dimethyl-1,3-propanediol and 1,15-pentadecanediol), indicating that their experimental values are probably incorrect. General experience suggests that, in cases where the Vm method exhibits a large deviation, it is the GA method that is more trustworthy.
While the intermolecular interactions of OH groups exhibit a large influence on the heat capacity of solids, a similar effect of saturated cyclic structures over non-cyclic ones should not be expected as their interactions merely result from the weak dispersive forces. Beyond this—and in contrast to the conditions in the liquid phase—in a solid crystal not only the translational but also the intramolecular freedoms of motion are largely restricted independent of cyclic or non-cylic molecular moieties. This seems to be confirmed by the smaller contribution of the saturated endocyclic bonds (special group 194 in Table 7) of −1.44 J/mol/K compared to that for the calculation of the liquid Cp of −3.92 J/mol/K. However, as has been demonstrated by the comparison of some structurally closely related examples in [1], e.g., o-, m-, and p-quinquephenyl, anthracene, phenanthrene, and various dimethylnaphthalenes, although aromatics, the chemical structure of a molecule itself has a very dominant effect on the crystalline structure, which again affects the experimental value of the solid heat capacity. In Table 9, a selection of saturated alkanes and cycloalkanes has been listed and their experimental solid heat capacities compared with the prediction values calculated by means of the present GA and the Vm method [1].
A quick scan of the deviations of the Vm-calculated Cp(sol,298) values (column 5 in Table 9) immediately shows that the Vm method systematically overestimates the solid heat capacity of the cycloalkanes (norbornane and the dicyclohexyldodecanes being exceptions), whereas that of the ring-free alkanes is systematically underestimated. Although the overestimation in the case of the cycloalkanes resembles the one found in the estimation of their liquid heat capacity, as demonstrated in Table 5, it does not seem far-fetched to assume that at least part of its extent lies in potentially more clearly defined crystalline structures as compared to the probably waxy consistence of the linear counterparts. The predicted Cp(sol,298) values resulting from the present GA method, on the other hand, yield excellent conformation with the experimental data. The largest deviations are interestingly found for norbornane, one of the exceptions in the Vm calculations, and bicyclo[3.3.3]undecane. For norbornane, the experimental value published by Steele [55] should be higher by ca. 8% to fit the respective deviations into the general picture of both prediction methods. For bicyclo[3.3.3]undecane, both prediction methods suggest a ca. 10% higher Cp(sol,298) value than reported by Parker et al. [56].
In conformance with the findings of the logP analysis in an earlier paper [6], the amino acids are assumed to exist in a zwitterionic form as solids (phenylglycine being an exception, as shown in Table 9 of [6], due to the lower basicity of the nitrogen atom conjugated to the phenyl ring). Accordingly, in Table 4, their carboxylate group is represented by entry 74, their alkyl- and dialkyl-ammonium functions by entries 148 and 149, respectively, and their immediate neighbours, the methyl and methylene groups, by the respective entries 4 and 11. Test calculations based on their non-ionic forms resulted in systematically and significantly overestimating the Cp(sol,298) values, indicating that their corresponding atom groups in Table 4, (i.e., entry 73, representing the carboxylic acid, and entries 122 and 125, representing the alkyl- and dialkyl-amino groups, respectively) are not applicable in the heat-capacity evaluation of amino acids. These results, however, should not be interpreted as a confirmation of the zwitterionic form of the amino acids as solids, because the basis for the parameters representing the neutral alkyl- and dialkyl-amino groups is at present too small, and a recalibration of the group parameters of Table 4 by applying the non-ionic instead of the zwitter-ionic forms could well lead to better-conforming GA-based results of the non-ionic forms with the experimental data.
In contrast, analogous comparative calculations based on the Vm method [1] revealed only minor prediction differences between the ionic and the non-ionic forms, which was to be expected as the "true" molecular volumes of both the prototropic forms are very similar. Typical examples listed in Table 10 demonstrate these observations. The MAPD between the experimental data and those calculated for the zwitterionic forms in Table 10 was 3.36 J/mol/K on applying the GA method and 4.02 J/mol/K when using the Vm approach.
As a consequence, and despite their excellent predictive quality, both the GA- and the Vm-based methods are not suitable to answer the question as to which form the amino acids exist as solids.
In the optimization process for the evaluation of the atom and special group contributions of Table 7, it turned out that 51 compounds had to be eliminated as outliers and have been collected in a separate list, available in the Supplementary Materials. This list again largely corresponded to the one resulting from the Vm optimization procedure. The remaining 800 compounds finally supported 126 atom and special groups valid for Cp(sol,298) predictions (row A in Table 7). Despite the smaller number of valid groups as compared to that of Table 3 for the liquid heat capacities, with 65% they cover an even slightly larger percentage of ChemBrain’s representative database.

5. Conclusions

The present paper is extending the series of publications [6,7,8,9] about the direct and indirect calculation of 14 molecular properties (enthalpy of combustion, formation, vaporization, sublimation and solvation, entropy of fusion, logPo/w, logS, logγinf, refractivity, polarizability, toxicity, liquid viscosity, and surface tension) by means of a single computer algorithm, adding two further molecular properties, the heat capacity for the liquid and solid phase of molecules. A comparison of the prediction quality of the present GA method for the heat capacities with that based on the "true" molecular volume [1] published recently proved a significantly higher accuracy over the latter. This was accomplished by directly addressing the deficiencies of the molecular volume approach, particularly its inevitable neglect of the intermolecular formation of hydrogen bridges of the OH groups as well as its non-consideration of the cyclization effect of saturated rings over ring-open forms on the heat capacity of both the liquid and solid phases. However, since the group additivity method in principle lacks the comprehensive range of the molecular volume approach, both prediction methods are beneficial in their own right—and they complement each other all the more, as in most cases they confirm each other’s result within explicable deviations. Therefore, in the present ongoing project ChemBrain IXL, version 5.9, available from Neuronix Software (www.neuronix.ch, Rudolf Naef, Lupsingen, Switzerland), the results of both methods are added to the database, the group-additivity result carrying the suffix "calc" and the volume-derived one the suffix "pred".

Supplementary Materials

Supplementary materials can be accessed at https://www.mdpi.com/1420-3049/25/5/1147/s1. The lists of 3D structures of the compounds and their experimental values used for the liquid and solid heat-capacities calculations are available online as standard SDF files under the names names “S01. Compounds List for Cp(liq,298) calculations.sdf” and “S02. Compounds List for Cp(sol,298) calculations.sdf”, respectively. The lists of compounds and their experimental and calculated Cp values are available as doc files under the names " S03. Experimental vs. calculated Cp(liq,298) Data Table.doc" and " S04. Experimental vs. calculated Cp(sol,298) Data Table.doc". The lists of outliers are available as excel files under the names "S05. Outliers of Cp(liq,298) by GA approach.xls" and " S06. Outliers of Cp(sol,298) by GA approach.xls". The figures are available as tif files and the tables as doc files under the names given in the text.

Funding

This research received no funding.

Acknowledgments

R. Naef is indebted to W.E. Acree of the University of North Texas for his support in the collection of literature references and for valuable discussions. He is also indebted to the library of the University of Basel for enabling full and free access to its electronic literature database.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

  • GA: atom-group additivity
  • Vm: "true" molecular volume

References

  1. Naef, R. Calculation of the isobaric heat capacities of the liquid and solid phase of organic compounds at and around 298.15 K based on their “True” molecular volume. Molecules 2019, 24, 1626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gardas, R.L.; Coutinho, J.A.P. A group contribution method for heat capacity estimation of ionic liquids. Ind. Eng. Chem. Res. 2008, 47, 5751–5757. [Google Scholar] [CrossRef]
  3. Benson, S.W.; Buss, J.H. Additivity rules for the estimation of molecular properties. Thermodynamic properties. J. Chem. Phys. 1958, 29, 546–572. [Google Scholar] [CrossRef]
  4. Zàbransky, M.; Ruzicka, V. Estimation of the heat capacities of organic liquids as a function of temperature using group additivity: An amendment. J. Phys. Chem. Ref. Data 2004, 33, 1071–1081. [Google Scholar]
  5. Chickos, J.S.; Hesse, D.G.; Liebman, J.F. A group additivity approach for the estimation of heat capacities of organic liquids and solids at 298 K. Struct. Chem. 1993, 4, 261–269. [Google Scholar] [CrossRef]
  6. 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] [CrossRef] [Green Version]
  7. 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 ordinary organic molecules and total phase-change entropy of liquid crystals. Molecules 2017, 22, 1059. [Google Scholar] [CrossRef] [Green Version]
  8. 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] [Green Version]
  9. Naef, R.; Acree, W.E., Jr. 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. [Google Scholar] [CrossRef] [Green Version]
  10. Hardtwig, E. Fehler-und Ausgleichsrechnung. Hochschultaschenbücher 262/262a; Bibliographisches Institut AG: Mannheim, Germany, 1968. [Google Scholar]
  11. Wong, W.-K.; Westrum, E.F., Jr. Thermodynamics of globular molecules. XVII. Heat capacities and transition behavior of bicyclo[2.2.2]octane and bicyclo[2.2.2]octene. J. Phys. Chem. 1970, 74, 1303–1308. [Google Scholar] [CrossRef]
  12. Figueroa-Gerstenmaier, S.; Cabanas, A.; Costas, M. Self-association and complex formation in alcohol-unsaturated hydrocarbon systems. Heat capacities of linear alcohols mixed with alkenes and alkynes. Phys. Chem. Chem. Phys. 1999, 1, 665–674. [Google Scholar] [CrossRef]
  13. Zaitsau, D.H.; Emel’yanenko, V.N.; Pimerzin, A.A.; Verevkin, S.P. Benchmark properties of biphenyl as a liquid organic hydrogen carrier: Evaluation of thermochemical data with complementary experimental and computational methods. J. Chem. Thermodyn. 2018, 122, 1–12. [Google Scholar] [CrossRef]
  14. Mahnel, T.; Pokorny, V.; Fulem, M.; Sedmidubskym, D.; Ruzicka, K. Measurement of low-temperature heat capacity by relaxation technique: Calorimeter performance testing and heat capacity of benzo[b]fluoranthene, benzo[k]fluoranthene, and indeno[1,2,3-cd]pyrene. J. Chem. Thermodyn. 2020, 142, 105964. [Google Scholar] [CrossRef]
  15. Chang, S.-S.; Westrum, E.F., Jr. Heat Capacities and Thermodynamic Properties of Globular Molecules. I. Adamantane and Hexamethylenetetramine. J. Phys. Chem. 1960, 64, 1547–1551. [Google Scholar] [CrossRef]
  16. Mendoza-Ruiz, E.A.; Mentado-Morales, J.; Flores-Segura, H. Standard molar enthalpies of formation and phase changes of Tetra-N-phenylbenzidine and 4,4′-Bis (N-carbazolyl)-1,1′-biphenyl. J. Therm. Anal. Calor. 2019, 135, 2337–2345. [Google Scholar] [CrossRef]
  17. Van Bommel, M.J.; Van Miltenburg, J.C.; Schuijff, A. Heat-capacity measurements and thermodynamic functions of 1,3,5-triazine and 1,3,5-trioxane. J. Chem. Thermodyn. 1988, 20, 397–403. [Google Scholar] [CrossRef]
  18. Piekarski, H.; Tkaczyk, M.; Tyczynska, M. Heat capacity and phase behavior of aqueous diethylene glycol n-pentyl ether by DSC. Thermochim. Acta 2012, 550, 19–26. [Google Scholar] [CrossRef]
  19. Wasiak, M.; Tkaczyk, M.; Piekarski, H. Heat capacity and phase behaviour of aqueous solutions of triethylene glycol monopentyl ether. Two point scaling analysis. Fluid Phase Equil. 2017, 431, 16–23. [Google Scholar] [CrossRef]
  20. Emel’yanenko, V.N.; Zaitsau, D.H.; Pimerzin, A.A.; Verevkin, S.P. Benchmark properties of diphenyl oxide as a potential liquid organic hydrogen carrier: Evaluation of thermochemical data with complementary experimental and computational methods. J. Chem. Thermodyn. 2018, 125, 149–158. [Google Scholar] [CrossRef]
  21. Ashrafi, F.; Saadati, R.; Behboodi, A. Modeling and theoretical calculation of liquid heat capacity of alcohols and aldehydes using QSPR. African J. Pure Appl. Chem. 2008, 2, 116–120. [Google Scholar]
  22. Campos, J.B.; Martinez-Gomez, A.J.; Orozco-Guareno, E. Design and building of an isoperibolic calorimeter: Measurements of enthalpy of formation for derivatives of glycidol. Meas. Sci. Technol. 2019, 30, 035902. [Google Scholar] [CrossRef]
  23. Buchholz, H.; Hylton, R.K.; Brandenburg, J.G.; Seidel-Morgenstern, A.; Lorenz, H.; Stein, M.; Price, S.L. The thermochemistry of racemic and enantiopure molecular crystals for predicting enantiomer separation. Cryst. Growth Des. 2017, 17, 4676–4686. [Google Scholar] [CrossRef]
  24. Nikitin, E.D.; Popov, A.P.; Bogatishcheva, N.S.; Faizullin, M.Z. Critical temperatures and pressures, heat capacities, and thermal diffusivities of levulinic acid and four n-alkyl levulinates. J. Chem. Thermodyn. 2019, 135, 233–240. [Google Scholar] [CrossRef]
  25. Su, C.; Zhu, C.; Ye, F.Y.Z.; Liu, X.; He, M. Isobaric molar heat capacity of ethyl octanoate and ethyl decanoate at pressures up to 24 Mpa. Chem. Eng. Data 2018, 63, 2252–2256. [Google Scholar] [CrossRef]
  26. Van Bommel, M.J.; Oonk, H.A.J.; Van Miltenburg, J.C. Heat Capacity measurements of 13 methyl esters of n-carboxylic acids from methyl octanoate to methyl eicosanoate between 5 K and 350 K. J. Chem. Eng. Data 2004, 49, 1036–1042. [Google Scholar] [CrossRef]
  27. Xu, J.; Li, S.; Zeng, Z.; Xue, W. Heat capacity, density, vapor pressure, and enthalpy of vaporization of isoamyl DL-lactate. J. Chem. Eng. Data 2019, 64, 3793–3798. [Google Scholar] [CrossRef]
  28. Neau, S.H.; Flynn, G.L. Solid and liquid heat capacities of n-alkyl para-aminobenzoates near the melting point. Pharmaceut. Res. 1990, 7, 1157–1162. [Google Scholar] [CrossRef]
  29. Chorazewski, M.; Goralski, P.; Tkaczyk, M. Heat capacities of 1-chloroalkanes and 1-bromoalkanes within the temperature range from 284.15 K to 353.15 K. A group additivity and molecular connectivity analysis. J. Chem. Eng. Data 2005, 50, 619–624. [Google Scholar] [CrossRef]
  30. Goralski, P.; Tkaczyk, M.; Chorazewski, M. Heat capacities of α,ω-dichloroalkanes at temperatures from 284.15 K to 353.15 K and a group additivity analysis. J. Chem. Eng. Data 2003, 48, 492–496. [Google Scholar] [CrossRef]
  31. Lipovska, M.; Schmidt, H.-G.; Rohac, V.; Ruzicka, V.; Wolf, G.; Zabransky, M. Heat capacities of three isomeric chlorobenzenes and of three isomeric chlorophenols. J. Therm. Anal. Calor. 2002, 68, 753–766. [Google Scholar] [CrossRef]
  32. Goralski, P.; Piekarski, H. Heat capacities and densities of some liquid chloro-, bromo-, and bromochloro-substituted benzenes. J. Chem. Eng. Data 2007, 52, 655–659. [Google Scholar] [CrossRef]
  33. Michalski, D.; Perry, R.T.; White, M.A. Additivity of guest and host properties in clathrates: A thermodynamic and Raman spectroscopic investigation of HPTB-based solids. J. Phys.: Condens. M. 1996, 8, 1647–1661. [Google Scholar] [CrossRef]
  34. Roux, M.V.; Temprado, M.; Jimenez, P.; Guzman-Mejia, R.; Juaristi, E.; Chickos, J.S. Heat capacities of thiane sulfones and thiane sulfoxide Refining of Cp group values for organosulfur compounds and their oxides. Thermochim. Acta 2003, 406, 9–16. [Google Scholar] [CrossRef]
  35. Zhang, Z.Y.; Yang, M.L. Heat capacity and phase transition of 2-amino-2-methyl-1,3-propanediol from 280K to the melting point. Thermochim. Acta 1990, 169, 263–269. [Google Scholar] [CrossRef]
  36. Chiu, L.-F.; Liu, H.-F.; Li, M.-H. Heat capacity of alkanolamines by differential scanning calorimetry. J. Chem. Eng. Data 1999, 44, 631–636. [Google Scholar] [CrossRef]
  37. Soares, B.P.; Stejfa, V.; Ferreira, O.; Pinho, S.P.; Ruzicka, K.; Fulem, M. Vapor pressures and thermophysical properties of selected ethanolamines. Fluid Phase Equil. 2018, 473, 245–254. [Google Scholar] [CrossRef] [Green Version]
  38. Perisanu, S.; Contineanu, I.; Neacsu, A.; Notario, R.; Roux, M.V.; Liebman, J.F.; Dodson, B.J. Thermochemistry and quantum chemical calculations of two dibenzocycloalkane nitriles. Struct. Chem. 2011, 22, 89–94. [Google Scholar] [CrossRef]
  39. Stejfa, V.; Fulem, M.; Ruzicka, K. Vapor pressure of 4-ethylmorpholine revisited: Thermodynamically consistent vapor pressure equation. J. Chem. Eng. Data 2019, 644, 1605–1610. [Google Scholar] [CrossRef]
  40. Tyunina, V.V.; Krasnov, A.V.; Tyunina, E.Y.; Badelin, V.G.; Rybkin, V.V. Enthalpies of sublimation of L-methionine and DL-methionine: Knudsen’s effusion mass spectrometric study. J. Chem. Thermodyn. 2019, 135, 287–295. [Google Scholar] [CrossRef]
  41. Abdelaziz, A.; Zaitsau, D.H.; Buzyurov, A.; Minakov, A.A.; Verevkin, S.P.; Schick, C. Fast scanning calorimetry: Sublimation thermodynamics of low volatile and thermally unstable compounds. Thermochim. Acta 2019, 676, 249–262. [Google Scholar] [CrossRef]
  42. Stejfa, V.; Fulem, M.; Ruzicka, K. Thermodynamic study of selected monoterpenes IV. J. Chem. Thermodyn. 2019. [Google Scholar] [CrossRef]
  43. Zhang, G.-Y.; Jin, S.-H.; Li, L.-J.; Li, Z.-H.; Shu, Q.-H.; Wang, D.-Q.; Zhang, B.; Li, Y.-K. Evaluation of thermal hazards and thermokinetic parameters of 3-amino-4-amidoximinofurazan by ARC and TG. J. Therm. Anal. Calor. 2016, 126, 1223–1230. [Google Scholar] [CrossRef]
  44. Li, R.; Hua, Y.; Tan, Z.; Shi, Q. Low temperature calorimetry of 3-fluoro-5-(3-pyridinyloxy) benzenamine and N-[3-fluoro-5-(3-Pyridinyloxy)phenyl]-N’-3-pyridinyl urea. Intern. J. Thermodyn. 2017, 20, 153–157. [Google Scholar] [CrossRef]
  45. Baird, Z.S.; Dahlberg, A.; Uusi-Kyyny, P.; Osmanbegovic, N.; Witos, J.; Helminen, J.; Cederkrantz, D.; Hyväri, P.; Alopaeus, V.; Kilpeläinen, I.; et al. Physical properties of 7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene (mTBD). Intern. J. Thermophys. 2019, 40, 71. [Google Scholar] [CrossRef] [Green Version]
  46. Liu, Y.; Di, Y.; Di, Y.; Qiao, C.; Che, F.; Yuan, F.; Yue, K.; Zhou, C. The studies of structure, thermodynamic properties and theoretical analyses of 2-[(4-nitro-benzoyl)-hydrazone]-propionic acid. J. Mol. Struct. 2019, 1184, 532–537. [Google Scholar] [CrossRef]
  47. Vilas-Boas, S.M.; Pokorny, V.; Stejfa, V.; Ferreira, O.; Pinho, S.P.; Ruzicka, K.; Fulem, M. Vapor pressure and thermophysical properties of eugenol and (+)-carvone. Fluid Phase Equil. 2019, 499, 112248. [Google Scholar] [CrossRef] [Green Version]
  48. Skotnicki, M.; Drogon, A.; Calvin, J.J.; Rosen, P.F.; Woodfield, B.F.; Pyda, M. Heat capacity and enthalpy of indapamide. Thermochim. Acta 2019, 674, 36–43. [Google Scholar] [CrossRef]
  49. Cai, M.; Zhou, T.-H.; Li, Y.-N.; Xu, K.-Z. Two interesting derivatives of 1-amino-1-hydrazino-2,2-dinitroethylene (AHDNE): Synthesis and thermal properties. J. Therm. Anal. Calor. 2019. [Google Scholar] [CrossRef]
  50. Doncaster, A.M.; Walsh, R. Thermochemistry of silicon-containing compounds. J. Chem. Soc. Faraday Trans. 1986, 82, 707–717. [Google Scholar] [CrossRef]
  51. Waliszewski, D.; Stepniak, I.; Piekarski, H.; Lewandowski, A. Heat capacities of ionic liquids and their heats of solution in molecular liquids. Thermochim. Acta 2005, 433, 149–152. [Google Scholar] [CrossRef]
  52. Lopez-Bueno, C.; Bugallo, D.; Leboran, V.; Rivadulla, F. Sub-μL measurements of the thermal conductivity and heat capacity of liquids. Phys.Chem.Chem.Phys. 2018, 20, 7277. [Google Scholar] [CrossRef] [PubMed]
  53. Zaitsau, D.H.; Schmitz, A.; Janiak, C.; Verevkin, S.P. Heat capacities of ionic liquids based on tetrahydrothiophenium cation and NTf2 anion. Thermochim. Acta 2020, 686, 178547. [Google Scholar] [CrossRef]
  54. Huelsekopf, M.; Ludwig, R. Temperature dependence of hydrogen bonding. J. Mol. Liq. 2000, 85, 105–125. [Google Scholar] [CrossRef]
  55. Steele, W.V. The standard enthalpies of formation of a series of C7 bridged-ring hydrocarbons: Norbornane, norbornene, nortricyclene, norbornadiene, and quadricyclane. J. Chem. Thermodyn. 1978, 10, 919–927. [Google Scholar] [CrossRef]
  56. Parker, W.; Steele, W.V.; Watt, I. A high-precision aneroid static-bomb combustion calorimeter for samples of about 20 mg. The standard enthalpy of formation of bicyclo[3.3.3]undecane. J. Chem. Thermodyn. 1975, 7, 795–802. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available.
Figure 1. Correlation diagram of the Cp(liq,298) data (in J/mol/K). The cross-validation data are superimposed as red circles. (n = 1111; r2 = 0.998; q2 = 0.9975; regression line: intercept = 0.7993; slope = 0.9977).
Figure 1. Correlation diagram of the Cp(liq,298) data (in J/mol/K). The cross-validation data are superimposed as red circles. (n = 1111; r2 = 0.998; q2 = 0.9975; regression line: intercept = 0.7993; slope = 0.9977).
Molecules 25 01147 g001
Figure 2. Histogram of the liquid heat-capacity data. The deviations are in J/mol/K. The cross-validation data are superimposed as red bars. (S = ± 9.19; Experimental value range: 81.92–1849 J/molK).
Figure 2. Histogram of the liquid heat-capacity data. The deviations are in J/mol/K. The cross-validation data are superimposed as red bars. (S = ± 9.19; Experimental value range: 81.92–1849 J/molK).
Molecules 25 01147 g002
Figure 3. Correlation diagrams of the calculated vs. experimental Cp(liq,298) data of saturated alcohols (in J/mol/K) based on A: the present GA; B: the Vm method.
Figure 3. Correlation diagrams of the calculated vs. experimental Cp(liq,298) data of saturated alcohols (in J/mol/K) based on A: the present GA; B: the Vm method.
Molecules 25 01147 g003
Figure 4. Correlation diagram of the Cp(sol,298) data (in J/mol/K). The cross-validation data are superimposed as red circles. (n = 734; r2 = 0.9915; q2 = 0.9874; regression line: intercept = −0.0999; slope = 0.9984).
Figure 4. Correlation diagram of the Cp(sol,298) data (in J/mol/K). The cross-validation data are superimposed as red circles. (n = 734; r2 = 0.9915; q2 = 0.9874; regression line: intercept = −0.0999; slope = 0.9984).
Molecules 25 01147 g004
Figure 5. Histogram of the solid heat-capacity data. The deviations are in J/mol/K. The cross-validation data are superimposed as red bars. (s = ±14.23; Experimental value range: 78.7–1129 J/molK).
Figure 5. Histogram of the solid heat-capacity data. The deviations are in J/mol/K. The cross-validation data are superimposed as red bars. (s = ±14.23; Experimental value range: 78.7–1129 J/molK).
Molecules 25 01147 g005
Table 1. Atom-group examples for ionic liquids and their meaning.
Table 1. Atom-group examples for ionic liquids and their meaning.
Atom TypeNeighboursMeaningaExample
B(-)F4BF4-tetrafluoroborate
C sp3H2CN(+)CCH2N(+)C1 in tetraalkylammonium
C sp3H2CP(+)CCH2P(+)C1 in tetraalkylphosphonium
C sp3H2CS(+)CCH2S(+)C1 in trialkylsulfonium
C(-) sp3C3C3C-central C- in tricyanocarbeniate
C aromaticH:C:N(+)C:CH:N+C2 in pyridinium
C(+) aromaticH:N2N:C+H:NC2 in imidazolium
C spB#N(-)B-(C#N)C in tetracyanoborate
C spC#N(-)C-(C#N)cyano-C in tricyanocarbeniate
C spN#N(-)N-(C#N)C in dicyanoamide
C sp=N=S(-)N=C=S-thiocyanate
N(+) sp3C4N+C4tetraalkylammonium
N(+) sp2O2=O(-)NO3-nitrate
N aromaticC2:C(+)C-N(C):C+N1 and N3 in 1,3-dialkylimidazolium
N(+) aromaticC:C2C:N+(C):CN in 1-alkylpyridinium
N(-)C2C-N--CN- in dicyanoamide
N(-)CSC-N--SN- in saccharinate
N(-)S2S-N--Sbis(trifluoromethanesulfonyl)amide
P4CO2=O(-)CPO3-alkylphosphonate
P(+)C4PC4+tetraalkylphosphonium
P(-)C3F3F3P-C3tris(pentafluoroethyl)trifluorophosphate
P(-)F6PF6-hexafluorophosphate
S(+)C3C3S+trialkylsulfonium
S4CN=O2(-)CS(O2)N-bis(trifluoromethanesulfonyl)amide
S4CO=O2(-)CSO3-alkylsulfonate
S4O2=O2(-)SO4-alkylsulfate
a The central atom defined by the atom type is indicated by a bold character.
Table 2. Refined atom and special groups and their meaning.
Table 2. Refined atom and special groups and their meaning.
Atom TypeNeighboursMeaning
O(prim)HCPrimary alcohol
O(sec)HCSecondary alcohol
O(tert)HCTertiary alcohol
Endocyclic bondsNo of single bondsCount single bonds in cyclic ring
Angle60 Bond angle < 60 deg
Angle 90 Bond angle between 60 and 90 deg
Angle102 Bond angle between 90 and 102 deg
(COH)nn > 1Molecule contains more than 1 OH group
HH AcceptorIntramolecular H bridge between acidic H (on O, N or S) and basic acceptor (O, N or F)
Table 3. Atom groups and their contributions for the heat-capacity calculation of liquids.
Table 3. Atom groups and their contributions for the heat-capacity calculation of liquids.
EntryAtom TypeNeighboursContributionOccurrencesMolecules
1BC324011
2B(-)C4698.6622
3B(-)F451.2166
4C sp3H3C37.031555790
5C sp3H3N100.02127101
6C sp3H3N(+)147.912018
7C sp3H3O81.298466
8C sp3H3S84.431713
9C sp3H3S(+)172.1711
10C sp3H3P217.5911
11C sp3H3Si717118
12C sp3H2BC–37.0331
13C sp3H2C230.063249696
14C sp3H2CN90.52222146
15C sp3H2CN(+)142.857852
16C sp3H2CO73.86477243
17C sp3H2CS75.253827
18C sp3H2CS(+)136.252910
19C sp3H2CP252.0821
20C sp3H2CP(+)71.8123
21C sp3H2CCl63.673830
22C sp3H2CBr63.932621
23C sp3H2CJ67.73109
24C sp3H2CSi60.71188
25C sp3H2N2151.5542
26C sp3H2NO157.511212
27C sp3H2O2111.4744
28C sp3H2S2128.4111
29C sp3HC321.11303196
30C sp3HC2N81.861414
31C sp3HC2N(+)159.633
32C sp3HC2O67.4510787
33C sp3HC2S67.46109
34C sp3HC2Si36.4811
35C sp3HC2Cl56.5699
36C sp3HC2Br56.8444
37C sp3HC2J62.2122
38C sp3HCNO(+)176.831
39C sp3HCO296.733
40C sp3HCF2157.2611
41C sp3HCFCl73.8911
42C sp3HCCl286.1598
43C sp3HCClBr89.6811
44C sp3HCBr282.8521
45C sp3C47.786251
46C sp3C3N81.3354
47C sp3C3N(+)55.2333
48C sp3C3O57.442321
49C sp3C3S57.4375
50C sp3C3F43.6653
51C sp3C3Cl50.9211
52C sp3C3Br54.6211
53C sp3C2N2(+)223.6622
54C sp3C2O299.7111
55C sp3C2F250.887813
56C sp3C2FCl64.3152
57C sp3C2Cl287.4222
58C sp3CNF2112.9931
59C sp3CF366.923123
60C sp3CSF2011
61C sp3CPF2(-)44.2362
62C sp3CF2Cl89.6744
63C sp3CF2Br86.4174
64C sp3CFCl288.1132
65C sp3CCl3102.5588
66C sp3SF3102.7815178
67C(-) sp3C3131.7611
68C sp2H2=C35.646159
69C sp2HC=C22.79195107
70C sp2HC=N95.9444
71C sp2HC=O54.972323
72C sp2H=CN92.2516687
73C sp2H=CO42.81110
74C sp2H=CS87.4155
75C sp2H=CCl56.4153
76C sp2H=CSi43.6944
77C sp2HN=N32.9133
78C sp2HN=O100.6333
79C sp2HO=O58.8277
80C sp2H=NS15.4822
81C sp2C2=C16.225444
82C sp2C2=N333.4811
83C sp2C=CN89.6132
84C sp2C2=O50.284949
85C sp2C=CO36.1655
86C sp2C=CS74.3154
87C sp2C=CCl160.2811
88C sp2CN=O87.091212
89C sp2CN=O(-)87.7811
90C sp2C=NS8.0611
91C sp2CO=O43.25216158
92C sp2CO=O(-)27.6387
93C sp2C=OS011
94C sp2C=OCl70.5376
95C sp2=CF256.521
96C sp2=CCl277.2654
97C sp2N2=N56.8811
98C sp2N2=O131.8333
99C sp2NO=O98.3211
100C sp2O2=O50.0455
101C aromaticH:C2221115238
102C aromaticH:C:N42.321913
103C aromaticH:C:N(+)–9.455332
104C aromaticH:N20 0
105C aromatic:C39.571911
106C aromaticC:C211.58251152
107C aromaticC:C:N30.5987
108C aromaticC:C:N(+)–2.691111
109C aromatic:C2N71.343129
110C aromatic:C2N(+)118.05118
111C aromatic:C2:N31.7233
112C aromatic:C2O33.824628
113C aromatic:C2S88.6877
114C aromatic:C2Si37.2107
115C aromatic:C2F37.065417
116C aromatic:C2Cl41.281915
117C aromatic:C2Br52.57118
118C aromatic:C2J43.4333
119C(+) aromaticH:N2–155.067474
120C spB#N(-)–130.9382
121C spH#C38.7865
122C spC#C23.99107
123C sp=C225.1644
124C spC#N48.783531
125C spC#N(-)–9.8131
126C sp#CSi49.5821
127C spN#N(-)–2.88126
128C sp=N2–89.6911
129C sp=N=O–20.7585
130C sp=N=S(-)43.6333
131N sp3H2C–4.353328
132N sp3H2C(pi)0.3199
133N sp3H2N48.6754
134N sp3HC2–71.792120
135N sp3HC2(pi)–72.441414
136N sp3HC2(2pi)–103.2866
137N sp3HCN–15.7443
138N sp3HCN(pi)–1311
139N sp3HCS(pi)–21.5211
140N sp3C3–160.233328
141N sp3C3(pi)–149.651714
142N sp3C3(2pi)–180.0233
143N sp3C3(3pi)–165.4611
144N sp3C2N–91.622
145N sp3C2N(2pi)–143.3722
146N sp3C2N(3pi)–160.7111
147N sp2H=C–243.1311
148N sp2C=C15.821713
149N sp2C=N–20.2421
150N sp2C=N(+)–42.2211
151N sp2=CN033
152N sp2=CO–53.4311
153N aromaticC2:C(+)–0.3114874
154N aromatic:C2−16.751515
155N(+) sp3H3C−44.3311
156N(+) sp3H2C2−140.4144
157N(+) sp3HC3−291.6411
158N(+) sp3C4−372.931313
159N(+) sp2C=NO(-)011
160N(+) sp2CO=O(-)−45.712517
161N(+) sp2O2=O(-)5.9944
162N(+) aromaticC:C214.223232
163N(-)C262.3666
164N(-)CS−32.5711
165N(-)S233.367373
166O(prim)HC14.3510289
167O(sec)HC36.174747
168O(tert)HC581111
169OHC(pi)48.395746
170OHP−119.3411
171OHS39.1111
172OC2−59.3217098
173OC2(pi)−26.57191149
174OC2(2pi)−15.472212
175OCN(+)(pi)55.5533
176OCN(2pi)011
177OCS16.0388
178OCP(pi)22.2531
179OCSi−22.04205
180OSi2−21.83197
181P4C2O=O(-)−344.9611
182P4CO2=O(-)011
183P4O3=O011
184P(+)C4−95.0633
185P(-)C3F333.1222
186P(-)F696.5399
187S2HC0.941919
188S2HC(pi)−25.4411
189S2C2−53.941919
190S2C2(pi)−77.0722
191S2C2(2pi)−89.8677
192S2CS−11.1184
193S4C2=O−23.4522
194S4C2=O2−18.8611
195S4CN=O2011
196S4CN=O2(-)5.6914774
197S4CO=O2(-)4.6199
198S4O2=O2(-)099
199S(+)C3−203.281010
200SiC4−89.091110
201SiC3O−49.4163
202SiC3Cl−25.4911
203SiC2O29.5166
204SiC2Cl225.4433
205SiCCl386.8333
206SiO4055
207(COH)nn > 1−27.732019
208HH Acceptor−21.4333
209Endocyclic bondsNo of single bds−3.921341243
210Angle60 4.136919
211Angle90 1.76319
ABased onValid groups134 1176
BGoodness of fitR20.998 1111
CDeviationAverage6.09 1111
DDeviationStandard8.24 1111
EK-fold cvK10 1060
FGoodness of fitQ20.9975 1060
GDeviationAverage (cv)6.85 1060
HDeviationStandard (cv)9.19 1060
Table 4. Experimental Cp(liq,298) data of 66 alkanols, compared with prediction values calculated by the present the group-additivity (GA) and the Vm [1] method (in J/mol/K).
Table 4. Experimental Cp(liq,298) data of 66 alkanols, compared with prediction values calculated by the present the group-additivity (GA) and the Vm [1] method (in J/mol/K).
Molecule NameCp(liq,298) calc. (GA)Dev. (%)Cp(liq,298) exp.Dev. (%)Cp(liq,298) calc. (Vm)
1-Propanol155.30−5.73146.88−8.12158.80
2-Propanol177.70−15.07154.43−2.83158.80
2-Methyl-1-propanol183.40−1.30181.05−5.05190.20
1-Butanol185.40−4.65177.16−7.81191.00
Cyclopentanol204.20−10.14185.40−9.76203.50
2-Butanol207.70−5.61196.673.34190.10
Isopentyl alcohol213.50−1.86209.60−5.92222.00
1-Pentanol215.40−3.49208.14−7.19223.10
2-Methyl-2-propanol226.50−3.61218.6012.76190.70
Cyclohexanol230.40−7.97213.40−9.89234.50
3-Methyl-2-butanol235.804.11245.909.68222.10
Cyclohexanemethanol236.100.17236.50−12.05265.00
3,3-Dimethyl-1-butanol237.10−0.43236.08−7.51253.80
3-Pentanol237.800.79239.706.76223.50
2-Ethyl-1-butanol243.501.28246.65−2.74253.40
2-Methyl-1-pentanol243.501.97248.40−2.62254.90
1-Hexanol245.50−1.15242.70−5.44255.90
Cycloheptanol256.50−2.51250.22−6.03265.30
2-Methyl-2-butanol256.60−3.76247.3010.31221.80
trans-2-Methylcyclohexanol258.501.70262.98−1.30266.40
cis-2-Methylcyclohexanol258.503.89268.951.51264.90
4-Methyl-2-pentanol265.902.36272.346.66254.20
3-Methyl-2-pentanol265.903.62275.897.93254.00
Cyclohexaneethanol266.20−0.08266.00−12.03298.00
2-Hexanol267.90−4.52256.310.20255.80
3-Hexanol267.900.51269.275.49254.50
1-Heptanol275.50−0.25274.81−4.84288.10
1-Methylcyclohexanol279.20−0.05279.055.03265.00
2-Methyl-2-pentanol286.700.81289.0311.84254.80
3-Methyl-3-pentanol286.702.25293.3013.57253.50
2,4-Dimethyl-3-pentanol294.005.77312.008.53285.40
5-Methyl-2-hexanol296.00−0.27295.202.74287.10
Cyclohexanepropanol296.20−1.09293.00−12.70330.20
2-Heptanol297.900.24298.633.59287.90
3-Heptanol297.905.19314.208.37287.90
4-Heptanol297.902.89306.776.15287.90
2-Ethyl-1-hexanol303.604.38317.50−0.28318.40
2-Methyl-1-heptanol303.603.00313.00−2.01319.30
5-Methyl-1-heptanol303.600.20304.20−4.73318.60
1-Octanol305.602.08312.10−2.40319.60
2-Methyl-2-hexanol316.70−1.01313.548.50286.90
2,5-Dimethyl-3-hexanol324.004.54339.406.25318.20
2-Methyl-4-heptanol326.001.75331.804.01318.50
4-Methyl-2-heptanol326.00−4.32312.50−1.70317.80
4-Methyl-3-heptanol326.00−5.43309.20−2.81317.90
6-Methyl-2-heptanol326.00−3.46315.10−1.30319.20
6-Methyl-3-heptanol326.00−4.99310.50−2.77319.10
2-Octanol328.000.64330.103.00320.20
3-Octanol328.003.10338.505.44320.10
4-Octanol328.001.23332.093.64320.00
1-Nonanol335.701.55341.00−3.37352.50
2-Methyl-2-heptanol346.80−2.73337.605.45319.20
4-Methyl-4-heptanol346.805.48366.9013.25318.30
2-Nonanol358.00−0.47356.321.10352.40
3-Nonanol358.004.18373.635.71352.30
4-Nonanol358.002.68367.864.26352.20
5-Nonanol358.003.44370.754.98352.30
3,7-Dimethyl-1-octanol361.801.47367.21−4.05382.10
n-Decyl alcohol365.703.00377.00−1.83383.90
5-Decanol388.104.35405.775.24384.50
1-Undecanol395.802.59406.34−2.60416.90
1-Dodecanol425.802.88438.42−2.44449.10
1-Tridecanol455.904.22476.00−1.11481.30
Myristyl alcohol486.003.91505.80−1.52513.50
1-Pentadecanol516.003.57535.10−1.98545.70
1-Hexadecanol546.10−4.26523.80−10.33577.90
MAPD 3.02 5.51
Table 5. Experimental Cp(liq,298) data of four linear alkanes and four related cycloalkanes, compared with the prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Table 5. Experimental Cp(liq,298) data of four linear alkanes and four related cycloalkanes, compared with the prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Molecule NameCp(liq,298)
calc. (GA)
Dev. (%)Cp(liq,298) Exp.Dev. (%)Cp(liq,298)
Calc.(Vm)
Cyclopentane130.70−1.48128.80−15.30148.50
Pentane164.301.73167.19−0.01167.20
Cyclohexane156.800.82158.10−12.14177.30
Hexane194.301.70197.660.08197.50
Cycloheptane183.00−1.32180.61−13.23204.50
Heptane224.400.41225.33−1.10227.80
Cyclooctane209.102.98215.53−10.29237.70
Octane254.400.50255.68−0.99258.20
Table 6. Experimental Cp(liq,298) data of 122 ionic liquids, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Table 6. Experimental Cp(liq,298) data of 122 ionic liquids, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Molecule NameCp(liq,298) GA-Calc.Dev. (%)Cp(liq,298) Exp.Dev. (%)Cp(liq,298) Vm-Calc.
1-Ethyl-3-methylimidazolium bromide256.403.17264.803.29256.10
1-Propyl-3-methylimidazolium bromide286.50−1.81281.40−1.92286.80
1-Ethyl-3-methylimidazolium thiocyanate300.00−6.59281.45−4.96295.40
1-Ethyl-3-methylimidazolium acetate321.100.25321.907.86296.60
1-Ethyl-3-methylimidazolium tetrafluoroborate307.600.16308.101.95302.10
1,3-Dimethylimidazolium methosulfate326.204.34341.0010.50305.20
1-Butyl-3-methylimidazolium chloride316.500.16317.003.63305.50
1-Ethyl-3-methylimidazolium dicyanamide313.000.52314.64−0.08314.90
1-Butyl-3-methylimidazolium bromide316.500.06316.70−0.38317.90
N-Methyl-2-hydroxyethylammonium propionate333.30−1.62328.002.10321.10
1-Ethyl-3-methylimidazolium methanesulfonate345.400.03345.506.80322.00
1-Ethyltetrahydrothiophenium dicyanamide339.60−1.26335.381.34330.90
1-Ethyl-3-methylimidazolium methylsulfate353.70−3.72341.000.67338.70
1-Ethyl-3-methylimidazolium hexafluorophosphate353.00−2.74343.600.67341.30
1-Butyl-3-methylimidazolium iodide316.50−0.80314.00−8.82341.70
1-Benzyl-3-methylimidazolium chloride341.00−0.47339.40−1.30343.80
1-Ethylpyridinium triflate348.700.88351.800.71349.30
1-Ethyl-3-methylimidazolium trifluoromethylsulfonate363.80−0.28362.803.64349.60
N-Methyl-2-hydroxyethylammonium butanoate363.40−0.66361.002.13353.30
1-Butyl-3-methylimidazolium thiocyanate360.206.44385.007.53356.00
1-Butyl-3-methylimidazolium acetate381.200.52383.206.16359.60
1-Butyl-3-methylimidazolium tetrafluoroborate367.70−0.79364.800.63362.50
1-Ethyl-3-methylimidazolium ethosulfate383.30−1.40378.002.99366.70
1-Propyl-3-methylimidazolium hexafluorophosphate383.00−2.30374.400.64372.00
1-Butyl-3-methylimidazolium dicyanoamide373.10−2.22365.00−2.19373.00
1-Butyl-3-methylimidazolium trifluoroacetate411.10−0.71408.205.66385.10
N-Methyl-2-hydroxyethylammonium pentanoate393.401.90401.003.87385.50
1-Butyltetrahydrothiophenium dicyanamide399.70−1.14395.190.96391.40
1-Butyl-3-methylpyridinium tetrafluoroborate379.202.27388.00−1.26392.90
1-Ethyl-3-methylpyridinium ethylsulfate394.70−1.47389.00−2.08397.10
1-Butyl-3-methylimidazolium methosulfate413.800.53416.004.50397.30
1-Benzyl-3-methylimidazolium tetrafluoroborate392.20−1.21387.50−3.23400.00
1-Butyl-3-methylimidazolium hexafluorophosphate413.10−1.32407.701.37402.10
1-Butyl-1-methylpyrrolidinium dicyanamide397.803.68413.001.86405.30
1-Butyl-3-methylimidazolium trifluoromethylsulfonate423.90−1.65417.001.75409.70
1-Hexyl-3-methylimidazolium tetrafluoroborate427.90−2.86416.00−1.32421.50
1-Pentyl-3-methylimidazolium hexafluorophosphate443.10−1.30437.401.03432.90
1-Butyl-1-methylpyrrolidinium trifluoromethanesulfonate448.70−3.15435.00−0.55437.40
1-Ethyl-3-methylimidazolium toluenesulfonate486.30−0.43484.206.84451.10
1-Hexyl-3-methylimidazolium trifluoromethylsulfonate484.003.64502.306.43470.00
1-Octyl-3-methylimidazolium tetrafluoroborate488.002.01498.002.99483.10
1-Ethyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide506.70−1.34500.003.36483.20
N-Ethylpyridinium bis(trifluoromethylsulfonyl)amide491.502.12502.152.98487.20
1-Ethyl-3-methylimidazolium 2-(2-methoxyethoxy)ethylsulfate530.50−0.86526.006.29492.90
1-Heptyl-3-methylimidazolium hexafluorophosphate503.30−0.54500.601.34493.90
1-Isopropyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide535.20−1.00529.903.25512.70
1-Propyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide536.70−0.34534.904.08513.10
1-Butyl-3-methylimidazolium toluenesulfonate546.400.36548.406.20514.40
N-Ethyl-2-methylpyridinium bis(trifluoromethylsulfonyl)amide535.30−0.15534.503.55515.50
1-Octyl-3-methylimidazolium hexafluorophosphate533.300.52536.102.16524.50
1-Cyclopropylmethyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide551.50−2.30539.101.21532.60
Trimethyl butylammonium bis(trifluoromethylsulfonyl)amide561.10−0.34559.204.02536.70
1,2-Diethylpyridinium bis(trifluoromethanesulfonyl) amide565.400.12566.104.82538.80
1-Butyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide566.80−0.16565.904.22542.00
1-sec-Butyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide565.30−1.47557.102.64542.40
1-Methyl-1-propylpyrrolidinium bis(trifluoromethanesulfonyl) amide561.50−1.35554.001.93543.30
1-Isobutyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide564.80−1.38557.102.48543.30
N-Propyl-2-methylpyridinium bis(trifluoromethylsulfonyl)amide565.40−1.33557.962.18545.80
N-Butylpyridinium bis(trifluoromethanesulfonyl) amide551.602.63566.523.50546.70
1-Nonyl-3-methylimidazolium hexafluorophosphate563.401.05569.402.79553.50
N-Octylisoquinolinium thiocyanate528.30−1.21522.00−6.88557.90
1-Butyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide593.400.44596.006.12559.50
N-Ethyl-4-dimethylaminopyridinium bis(trifluoromethanesulfonyl) amide591.200.52594.304.37568.30
1-Pentyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide596.90−0.22595.604.08571.30
1-Ethyl-2-propylpyridinium bis(trifluoromethanesulfonyl) amide595.40−0.25593.903.25574.60
1-Isobutyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)amide589.60−1.27582.201.03576.20
1-Isobutyl-3-methylpyridinium bis(trifluoromethylsulfonyl)amide576.300.47579.000.40576.70
N-Butyl-3-methylpyridinium bis(trifluoromethylsulfonyl)amide578.20−0.02578.100.14577.30
1-Butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)amide591.50−3.41572.00−1.05578.00
1-Butyl-3-cyanopyridinium bis(trifluoromethylsulfonyl)amide590.00−0.68586.001.11579.50
1-Benzyl-3-methylimidazolium bis(trifluoromethylsulfonyl) amide591.202.73607.804.34581.40
1-Decyl-3-methylimidazolium hexafluorophosphate593.401.64603.303.03585.00
1-Pentyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide623.500.56627.006.19588.20
1-Hexyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide626.900.37629.204.04603.80
1-Butyl-1-methylpiperidinium bis(trifluoromethylsulfonyl)amide617.70−1.68607.500.23606.10
1-Ethyl-2-butylpyridinium bis(trifluoromethanesulfonyl) amide625.50−0.30623.602.79606.20
N-Hexylpyridinium bis(trifluoromethanesulfonyl) amide611.800.03612.000.83606.90
1-Methyl-1-pentylpyrrolidinium bis(trifluoromethanesulfonyl) amide621.600.16622.602.28608.40
1-Butyl-3-methylimidazolium octylsulfate623.801.76635.003.92610.10
1-Cyclohexylmethyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide617.50−0.06617.100.26615.50
1-Hexyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide653.60−1.18646.004.30618.20
N-Butyl-4-dimethylaminopyridinium bis(trifluoromethanesulfonyl) amide651.400.96657.714.71626.70
1-Heptyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide657.000.33659.204.31630.80
1-Ethyl-2-pentylpyridinium bis(trifluoromethanesulfonyl) amide655.60−0.44652.702.47636.60
1-Hexyl-3-methylpyridinium bis(trifluoromethylsulfonyl)amide638.30−2.29624.00−2.20637.70
1-Hexyl-1-methylpyrrolidinium bis(trifluoromethanesulfonyl) amide651.700.52655.102.49638.80
1-Hexyl-4-cyanopyridinium bis(trifluoromethylsulfonyl)amide650.10−2.70633.00−1.04639.60
1-Hexyl-3-cyanopyridinium bis(trifluoromethylsulfonyl)amide650.101.20658.002.78639.70
1-Dodecyl-3-methylimidazolium hexafluorophosphate653.601.91666.302.97646.50
1-Heptyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide683.600.20685.005.37648.20
1-Octyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide687.100.45690.203.80664.00
1-Methyl-1-heptylpyrrolidinium bis(trifluoromethanesulfonyl) amide681.700.50685.102.96664.80
1-Octylpyridinium bis(trifluoromethylsulfonyl)amide671.902.06686.002.93665.90
1-Ethyl-2-hexylpyridinium bis(trifluoromethanesulfonyl) amide685.60−0.01685.502.68667.10
1-Octyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide713.700.74719.005.63678.50
1-(3,4,5,6-Perfluorohexyl)-3-methylimidazolium-3-methylimidazolium bis(trifluoromethanesulfonyl) amide719.300.79725.006.40678.60
4-Dimethylamino-1-hexylpyridinium bis(trifluoromethanesulfonyl) amide711.502.67731.005.99687.20
1-Ethyl-2-heptylpyridinium bis(trifluoromethanesulfonyl) amide715.700.25717.502.80697.40
1-Methyl-1-octylpyrrolidinium bis(trifluoromethanesulfonyl) amide711.800.63716.302.36699.40
1-Octyl-3-cyanopyridinium bis(trifluoromethylsulfonyl)amide710.20−0.17709.001.26700.10
N-Hexyl-3-methyl-4-dimethylaminopyridinium bis(trifluoromethanesulfonyl) amide738.10−1.81725.001.94710.90
1-Nonyltetrahydrothiophenium bis(trifluoromethylsulfonyl) amide743.70−0.36741.004.01711.30
Butyl 1-butylnicotinate bis(trifluoromethylsulfonyl)amide728.90−3.10707.00−0.91713.40
1-Decyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide747.201.01754.804.03724.40
1-Ethyl-2-octylpyridinium bis(trifluoromethanesulfonyl) amide745.700.49749.403.26725.00
1-Methyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide785.90−0.58781.404.20748.60
1-Ethyl-2-nonylpyridinium bis(trifluoromethanesulfonyl) amide775.800.35778.503.48751.40
1-Methyl-1-decylpyrrolidinium bis(trifluoromethanesulfonyl) amide771.900.90778.902.40760.20
1-Ethyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide813.400.61818.404.55781.20
1-Dodecyl-3-methylimidazolium bis(trifluoromethanesulfonyl) amide807.301.57820.204.04787.10
1-Ethyl-2-decylpyridinium bis(trifluoromethanesulfonyl) amide805.800.67811.202.79788.60
1-Propyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide843.50−0.49839.403.71808.30
1-Butyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide873.50–0.84866.203.20838.50
1-Pentyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide903.600.15905.003.55872.90
1-Hexyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide933.701.15944.604.74899.80
1-Heptyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide963.70–0.33960.503.25929.30
1-Octyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide993.800.12995.003.59959.30
1-Nonyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide1023.80−0.021023.602.83994.60
1-Decyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide1053.90–0.761045.902.311021.70
1-Undecyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide1084.000.401088.303.551049.70
1-Dodecyl-3-menthyloxymethylimidazolium bis(trifluoromethylsulfonyl)imide1114.000.351117.902.851086.00
Tetradecyl trihexylphosphonium bis(trifluoromethylsulfonyl)amide1312.10−1.021298.80−1.351316.30
MAPD 1.13 3.08
Table 7. Atom groups and their contributions for the heat-capacity calculation of solids.
Table 7. Atom groups and their contributions for the heat-capacity calculation of solids.
EntryAtom TypeNeighboursContributionOccurrencesMolecules
1B(-)F42.9911
2C sp3H3C37.12569246
3C sp3H3N101.285034
4C sp3H3N(+)99.5974
5C sp3H3O68.065135
6C sp3H3S45.8533
7C sp3H3P131.1211
8C sp3H3Si59.18145
9C sp3H2C225.451427249
10C sp3H2CN82.728956
11C sp3H2CN(+)81.351915
12C sp3H2CO64.67210108
13C sp3H2CS72.272212
14C sp3H2CF53.6841
15C sp3H2CCl50.1241
16C sp3H2CBr54.7352
17C sp3H2CJ52.9441
18C sp3H2CSi63.1211
19C sp3H2N2134.39133
20C sp3H2O2108.47123
21C sp3H2S2−6.6733
22C sp3HC311.9216473
23C sp3HC2N72.942823
24C sp3HC2N(+)70.742928
25C sp3HC2O51.8416163
26C sp3HC2S47.2242
27C sp3HC2Si142.2711
28C sp3HCN2137.0411
29C sp3HCNO119.8975
30C sp3HCNS116.0821
31C sp3HCO2112.621714
32C sp3HCF2246.811
33C sp3HCBr268.6711
34C sp3C4−2.178148
35C sp3C3N62.09119
36C sp3C3N(+)19.2322
37C sp3C3O27.591010
38C sp3C3Cl78.1311
39C sp3C3Br44.3411
40C sp3C2NO89.8611
41C sp3C2O291.7565
42C sp3C2S243.8652
43C sp3CF369.8922
44C sp3CSF2011
45C sp3CCl392.2843
46C sp2H2=C39.9666
47C sp2HC=C16.9510965
48C sp2HC=N99.991313
49C sp2HC=O43.051412
50C sp2H=CN33.942517
51C sp2H=CO41.9933
52C sp2H=CS36.1975
53C sp2H=CCl19.4411
54C sp2HN=N102.561714
55C sp2HN=O28.9143
56C sp2H=NO112.5711
57C sp2C2=C5.772922
58C sp2C2=N95.441410
59C sp2C2=N(+)−9.6122
60C sp2C=CN18.41615
61C sp2C2=O27.684430
62C sp2C=CO25.281513
63C sp2C=CS29.0854
64C sp2C=CCl35.263
65C sp2=CN238.931414
66C sp2=CN2(+)78.3277
67C sp2CN=N87.151913
68C sp2CN=N(+)118.2321
69C sp2CN=O37.3313192
70C sp2=CNO55.0911
71C sp2CN=S46.5133
72C sp2CO=O51.25208155
73C sp2CO=O(-)15.254140
74C sp2C=OCl61.4621
75C sp2=CS245.53122
76C sp2N2=N113.9753
77C sp2N2=O55.724338
78C sp2N=NO91.5611
79C sp2N2=S67.0277
80C sp2N=NS105.9577
81C sp2NO=O63.4388
82C sp2NO=S64.8533
83C sp2=NOS108.0111
84C sp2NS=S62.4643
85C sp2O2=O58.4755
86C sp2OS=S63.2311
87C aromaticH:C217.963232437
88C aromaticH:C:N24.163720
89C aromaticH:C:N(+)21.4721
90C aromaticH:N27.0833
91C aromatic:C38.0417157
92C aromaticC:C26.56699307
93C aromaticC:C:N5.85139
94C aromatic:C2N24.29172107
95C aromatic:C2N(+)49.485742
96C aromatic:C2:N19.63137
97C aromatic:C2O29.21184113
98C aromatic:C2P15.162
99C aromatic:C2S27.434328
100C aromatic:C2Si58.515312
101C aromatic:C2F31.08259
102C aromatic:C2Cl34.355725
103C aromatic:C2Br37.48189
104C aromatic:C2J48.4953
105C aromaticC:N218.8131
106C aromatic:CN:N39.0375
107C aromatic:C:NO53.9211
108C aromatic:C:NCl49.0433
109C aromaticN:N236.2242
110C aromatic:N2O40.8131
111C(+) aromaticH:N2−39.6333
112C(+) aromatic:N3−20.9422
113C spH#C103.0321
114C spC#C14.9583
115C spC#N39.642820
116C spC#N(+)62.5611
117C sp#CSi021
118C sp#NO58.0721
119C sp=N=O110.9863
120N sp3H2C−18.6255
121N sp3H2C(pi)15.4912998
122N sp3H2N9.5143
123N sp3H2S45.9299
124N sp3HC2−104.0763
125N sp3HC2(pi)−56.777857
126N sp3HC2(2pi)−9.788261
127N sp3HCN(pi)21.2275
128N sp3HCN(2pi)36.0977
129N sp3C3−159.231510
130N sp3C3(pi)−123.38109
131N sp3C3(2pi)−83.832721
132N sp3C3(3pi)−43.15126
133N sp3C2N(pi)−75.0733
134N sp3C2N(+)(pi)−52.6972
135N sp3C2N(2pi)25.3933
136N sp3C2N(+)(2pi)−41.1822
137N sp2C=C−74.595444
138N sp2C=N2.4653
139N sp2=CN−98.0177
140N sp2=CN(+)−14.7611
141N sp2=CO−44.632411
142N sp2N=N17.4332
143N aromaticH2:C(+)4.6542
144N aromaticHC:C(+)40.4411
145N aromaticC2:C(+)−3.1174
146N aromatic:C2−0.165033
147N(+) sp3H3C−6.73534
148N(+) sp3H2C2−73.7944
149N(+) sp3HC3−156.7111
150N(+) sp3C4−233.9322
151N(+) sp2CO=O(-)9.717249
152N(+) sp2=CO2(-)−7.4222
153N(+) sp2NO=O(-)−2.08105
154N(+) aromaticC:C2011
155N(+) spC#C(-)25.3433
156N(+) sp#CO(-)011
157O(prim)HC−23.3610660
158O(sec)HC−16.2511750
159O(tert)HC−3.3466
160OHC(pi)4.78218157
161OHN(pi)16.3333
162OHSi16.9382
163OC2−70.396029
164OC2(pi)−39.7112588
165OC2(2pi)−23.914741
166OCN(2pi)−28.5155
167OCSi−30.4369
168ON2(2pi)1.174
169ON2(+)(2pi)−2.222
170OSi2−7.14398
171P3C3−2.2511
172P4C3=O2.2511
173P4C=OCl2011
174S2HC10.7311
175S2HC(pi)16.0755
176S2C2−15.43128
177S2C2(pi)−30.85146
178S2C2(2pi)−10.042417
179S2CS−20.6642
180S2CS(pi)6.9163
181S4C2=O5.3322
182S4C2=O215.5655
183S4CN=O22.7599
184S4CO=O2(-)−118.4111
185SiC4−197.5133
186SiC3O−100.7942
187SiC3Cl−107.3811
188SiC2O2−38.26113
189SiCO310.05203
190SiCCl357.8922
191SiO4099
192(COH)nn>13.4614559
193HH Acceptor1.196143
194Endocyclic bondsNo of single bds−1.44998149
195Angle60 0.97152
196Angle90 0.31126
197Angle102 2.1628483
ABased onValid groups126 802
BGoodness of fitr20.9915 734
CDeviationAverage9.36 734
DDeviationStandard12.21 734
EK-fold cvK10 663
FGoodness of fitq20.9874 663
GDeviationAverage (cv)11.1 663
HDeviationStandard (cv)14.23 663
Table 8. Experimental Cp(sol,298) data of 31 alkanols, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Table 8. Experimental Cp(sol,298) data of 31 alkanols, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Molecule NameCp(sol,298)
GA-calc
Dev. (%)Cp(sol,298) exp.Dev. (%)Cp(sol,298)
Vm−calc. [1]
2-Methyl-2-propanol135.906.99146.1118.28119.40
2,2-Dimethyl-1,3-propanediol158.0013.75183.1813.42158.60
Erythritol164.20−1.42161.900.99160.30
cis-1,2-Cyclohexanediol168.00−4.74160.40−4.68167.90
trans-1,2-Cyclohexanediol168.00−2.94163.20−3.00168.10
Pentaerythritol174.407.43188.403.24182.30
Hexamethyleneglycol187.901.11190.002.47185.30
Xylitol206.900.05207.0010.14186.00
Ethriol191.1010.62213.807.86197.00
Inositol223.60−2.57218.000.50216.90
2-Adamantanol193.306.71207.20−6.13219.90
1-Adamantanol195.600.56196.70−12.10220.50
Dulcose246.00−3.14238.503.31230.60
Isoborneol243.106.88261.0610.21234.40
Borneol243.106.88261.069.68235.80
1,8-Octanediol238.90−1.07236.36−0.23236.90
Sorbitol242.30−1.38239.000.50237.80
Menthol250.70−0.24250.10−0.72251.90
1,9-Nonanediol264.30−2.94256.74−2.36262.80
1,10-Decanediol289.80−3.77279.26−3.34288.60
1,11-Undecanediol315.20−5.85297.79−5.58314.40
Tri-t-butylmethanol351.80−0.34350.606.36328.30
1,12-Dodecanediol340.70−3.17330.23−3.05340.30
1-Tridecanol358.605.13378.008.41346.20
1,13-Tridecanediol366.200.19366.880.21366.10
Myristyl alcohol384.001.03388.004.28371.40
1,14-Tetradecanediol391.60−3.16379.61−3.24391.90
1-Pentadecanol409.50−2.38400.000.88396.50
1,15-Pentadecanediol417.10−10.50377.45−10.66417.70
1-Hexadecanol435.00−3.08422.000.09421.60
1,16-Hexadecanediol442.50−3.83426.18−4.06443.50
MAPD 4.00 5.16
Table 9. Experimental Cp(sol,298) data of 18 alkanes and cycloalkanes, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Table 9. Experimental Cp(sol,298) data of 18 alkanes and cycloalkanes, compared with prediction values calculated by the present GA and the Vm [1] method (in J/mol/K).
Molecule NameCp(sol,298)
GA-Calc.
Dev. (%)Cp(sol,298) exp.Dev. (%)Cp(sol,298)
Vm-Calc. [1]
Nortricyclene130.70−1.32129.00−9.15140.80
Norbornane163.80−8.48151.000.93149.60
Bicyclo[2.2.2]octane163.80−3.87157.69−8.44171.00
Adamantane183.303.53190.00−5.95201.30
Bicyclo[3.3.3]undecane236.10−10.74213.20−9.19232.80
Diamantane222.100.58223.40−17.32262.10
Perhydrophenanthrene279.603.42289.50−0.73291.60
Tri-t-butylmethane339.504.31354.8015.30300.50
Cetane431.002.44441.8013.26383.20
Octadecane481.900.77485.6411.70428.80
Docosane583.80−3.58563.607.59520.80
2,11-Dicyclohexyldodecane563.40−1.09557.304.65531.40
1,1-Dicyclohexyldodecane565.20−0.46562.605.44532.00
Hexacosane685.60−3.69661.207.43612.10
Triacontane787.402.65808.8013.03703.40
Dotriacontane838.40−4.02806.007.07749.00
Tetratriacontane889.30−0.21887.4010.45794.70
Pentatriacontane914.700.13915.9010.74817.50
MAPD 3.07 8.80
Table 10. Comparison of the Cp(sol,298) data of the ionic and the non-ionic forms of amino acids calculated by the present GA and the Vm [1] method (in J/mol/K).
Table 10. Comparison of the Cp(sol,298) data of the ionic and the non-ionic forms of amino acids calculated by the present GA and the Vm [1] method (in J/mol/K).
Molecule nameCp(sol,298) GA-calc.Cp(sol,298) Exp.Cp(sol,298) Vm-Calc. [1]
Non-IonicZwitter-Ionic Zwitter-IonicNon-Ionic
Glycine120.2089.9099.3095.9091.30
Alanine147.60116.50119.90118.50116.10
N-Methylglycine136.00122.40118.20119.10116.90
Serine152.70121.70135.60126.20126.50
Aminobutyric acid173.00142.00146.40140.90140.70
Proline162.20137.30150.40149.80150.90
Threonine183.00152.00155.31153.30154.70
Aspartic acid193.00162.00155.18155.10156.00
Asparagine189.80158.70159.80157.60159.30
Valine196.70165.60165.00162.90165.40
5-Aminopentanoic acid196.60166.30163.70164.30166.60
Ornithine225.40194.40191.20179.20183.60
Glutamine214.00183.00184.18182.30186.50
Leucine222.10191.10200.80183.90188.20
Isoleucine222.10191.10188.28184.70189.90
Methionine238.50207.40205.16189.10194.30
N-Phenylglycine196.10162.20177.40194.60198.60
Phenylalanine232.20201.10203.10215.50221.60
8-Aminooctanoic acid273.00242.70251.70232.80241.90
Tyrosine248.20217.10216.44236.20238.10
Tryptophane268.30237.20238.15252.00263.90

Share and Cite

MDPI and ACS Style

Naef, R. Calculation of the Isobaric Heat Capacities of the Liquid and Solid Phase of Organic Compounds at 298.15K by Means of the Group-Additivity Method. Molecules 2020, 25, 1147. https://doi.org/10.3390/molecules25051147

AMA Style

Naef R. Calculation of the Isobaric Heat Capacities of the Liquid and Solid Phase of Organic Compounds at 298.15K by Means of the Group-Additivity Method. Molecules. 2020; 25(5):1147. https://doi.org/10.3390/molecules25051147

Chicago/Turabian Style

Naef, Rudolf. 2020. "Calculation of the Isobaric Heat Capacities of the Liquid and Solid Phase of Organic Compounds at 298.15K by Means of the Group-Additivity Method" Molecules 25, no. 5: 1147. https://doi.org/10.3390/molecules25051147

Article Metrics

Back to TopTop