Next Article in Journal
“Miswak” Based Green Synthesis of Silver Nanoparticles: Evaluation and Comparison of Their Microbicidal Activities with the Chemical Synthesis
Next Article in Special Issue
A Molecular Electron Density Theory Study of the Chemical Reactivity of Cis- and Trans-Resveratrol
Previous Article in Journal
Medium Optimization and Fermentation Kinetics for κ-Carrageenase Production by Thalassospira sp. Fjfst-332
Previous Article in Special Issue
A Theoretical Study of the Relationship between the Electrophilicity ω Index and Hammett Constant σp in [3+2] Cycloaddition Reactions of Aryl Azide/Alkyne Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Statistical Significance of the Maximum Hardness Principle Applied to Some Selected Chemical Reactions

Department of Chemistry and Centre for Theoretical Studies, Indian Institute of Technology Kharagpur, Kharagpur 721302, India
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(11), 1477; https://doi.org/10.3390/molecules21111477
Submission received: 16 September 2016 / Revised: 24 October 2016 / Accepted: 1 November 2016 / Published: 5 November 2016

Abstract

:
The validity of the maximum hardness principle (MHP) is tested in the cases of 50 chemical reactions, most of which are organic in nature and exhibit anomeric effect. To explore the effect of the level of theory on the validity of MHP in an exothermic reaction, B3LYP/6-311++G(2df,3pd) and LC-BLYP/6-311++G(2df,3pd) (def2-QZVP for iodine and mercury) levels are employed. Different approximations like the geometric mean of hardness and combined hardness are considered in case there are multiple reactants and/or products. It is observed that, based on the geometric mean of hardness, while 82% of the studied reactions obey the MHP at the B3LYP level, 84% of the reactions follow this rule at the LC-BLYP level. Most of the reactions possess the hardest species on the product side. A 50% null hypothesis is rejected at a 1% level of significance.

Graphical Abstract

1. Introduction

The conceptual density functional theory (CDFT) [1,2] has been shown to be useful in analyzing popular concepts in chemistry like electronegativity (χ) [3,4,5,6], chemical hardness (η) [7,8,9,10], electrophilicity (ω) [11,12,13,14], etc. The idea of hardness was first introduced by Pearson in the context of his famous hard-soft acids and bases (HSAB) principle [7,10,15,16,17,18], which states that “hard acids prefer to coordinate with hard bases and soft acids prefer to coordinate with soft bases”. In 1987, Pearson proposed the maximum hardness principle (MHP) [19] as “there seems to be a rule of nature that molecules arrange themselves so as to be as hard as possible”. Now, for the ground state of an N electronic system, hardness (η) is defined as:
η =   ( 2 E N 2 ) v ( r )
where E is the total energy of the system, and this equation is valid for constant external potential ( v ( r ) ).
A finite difference approximation to Equation (1) gives
η = ( I A )
where I and A are the ionization potential and electron affinity, respectively.
Now according to Koopmans’ theorem [20], I and A can be approximated in terms of the energies of the frontier molecular orbitals (I = −EHOMO and A = −ELUMO) and the Equation (2) can be modified as
η = ( E L U M O E H O M O )
where EHOMO and ELUMO are the energies of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), respectively.
Thus, the stability of molecules, atoms or ions is connected to their HOMO-LUMO energy difference. Moreover, the favorable direction of a chemical reaction can also be understood by comparing the HOMO-LUMO energy gaps of the reactants and the products.
A statistical mechanical proof of the MHP was given by Parr and Chattaraj [21] in 1991. They showed that the MHP holds good under two conditions; (i) constant electronic chemical potential (μ) and (ii) constant external potential ( v ( r ) ). The validity of MHP has been checked for a wide range of physico-chemical processes and found to be valid in many situations like internal rotations [22,23,24,25,26,27,28], molecular vibrations [29,30,31,32], isomer stability [33], chemical reactions [34,35,36,37,38,39,40,41,42,43,44,45], Woodward-Hoffmann rules [46,47], aromaticity [48,49,50], stability of magic clusters [51], atomic shell structure [52,53], time-dependent situations [54,55,56], electronic excitations [57], chaotic ionizations [58], etc. There are also certain examples in literature where the MHP is not properly obeyed [59,60,61,62,63,64,65,66].
In an article, Poater et al. [67] studied 34 reactions given in the BH76 set, of which 28 are exothermic and the other 6 reactions are thermoneutral. They found that the products have greater hardness than the reactants in only 46% of the reactions, and this is reduced to only 18% of the total reactions upon the inclusion of another criterion, lower hardness in the transition state than that of the reactant. The reason may be that very hard atoms (like H, N, O, F, etc.) or molecules (like H2, N2, HF, HCN, CH4, etc.) were present on the reactant side. Recently, Pan et al. studied 101 chemical reactions to check the validity of the MHP in which most of them were inorganic reactions [68]. The study showed that the null hypothesis associated with the chemical reactions obeying the MHP is rejected at the 5% level of significance.
In this article, we are interested in checking whether the MHP is valid in some selected chemical reactions. For this purpose we have investigated 50 exothermic chemical reactions, the majority of which are organic, exhibiting a special steroelectronic effect called the anomeric effect [69,70,71], to analyze the validity of the MHP therein. Other than this we have studied some disproportionation reactions [72] in the light of the MHP. The effect of level of theory on the validity of the MHP in the studied set is also explored. For the reactions involving more than one reactant and/or product, we have computed the geometric mean (ηgeo) of the η values. Further, the combined hardness (ηcom) of the product or reactant side is also calculated as the difference between minimum ionization potential (Imin) and maximum electron affinity (Amax) values, respectively. The performance of these two approximations is compared in validating the MHP. Statistical testing of the null hypothesis is also performed in the validity of the MHP in the studied set of reactions.

2. Computational Details Section

Geometries of all of the studied molecules are modeled in the Gaussview 5.0.8 [73] graphical interface. The geometries are optimized at the DFT level of theory using two different functionals, the Becke three-parameter hybrid functional combined with Lee–Yang–Parr correlation, B3LYP [74,75] and long-range corrected LC-BLYP [75,76,77,78,79,80]. The 6-311++G(2df,3pd) [81,82] basis set is used for the calculations, except for iodine and mercury. For iodine and mercury, we have used the def2-QZVP [83] basis set along with effective core potential to take care of the relativistic effects. Frequency calculations with the optimized geometries are also carried out at the same level of theory. These frequency calculations are necessary to ensure that the optimized structures are at the minima of the respective potential energy surfaces. All of the above calculations are performed with the Gaussian 09 program package [84]. For these computations, we have used a fine grid, a pruned (75,302) grid with 75 radial shells per atom and 302 angular points per shell, which is the default in the Gaussian 09 program. Furthermore, the default values for convergence criteria of the self-consistent field and geometry are also used.
The I and A values using the Koopmans’ theorem and the η values using Equation (3) are calculated for each molecule appearing on the reactant and product sides. The combined hardness, ηcom [30,85,86], is computed as:
ηcom = (IminAmax)
where Imin and Amax are the minimum of the ionization potential and the maximum of the electron affinity values of the molecules appearing on the reactant or product side.

3. Results and Discussion

The computed values of I and A at the B3LYP/6-311++G(2df,3pd)/def2-QZVP and LC-BLYP/6-311++G(2df,3pd)/def2-QZVP levels for each molecule involved in the considered set of reactions are tabulated in Table S1 (in Supplementary Materials). It is found that I and A values differ considerably at these two levels. I values at the LC-BLYP level are found to be higher by 2.74–4.24 eV than those at the B3LYP level, whereas A values at the former level are lower by 1.06–3.95 eV than those at the latter level. Interestingly, while the A values are negative in all species studied here at the LC-BLYP level, they are positive at the B3LYP level. Therefore, not only do these two levels provide significantly different I and A values quantitatively, but in the case of A they also produce a qualitatively different trend. Such huge changes in I and A values led us to check whether they provide similar results in obeying the MHP, despite considerable changes in η values. Note that LC-BLYP functional was reported to be superior to B3LYP in predicting such properties. The computed η values of the reactants and products are presented in Table 1 and Table 2 obtained from the B3LYP and LC-BLYP levels, respectively. All of the reactions are exothermic in nature. Therefore, according to the MHP, the exothermic reactions are expected to have greater hardness on the product side as compared to the reactant side, considering the entropy effects to be negligible.

3.1. Results at the B3LYP Level

Among the 50 reactions considered here, 41 reactions and thus 82.0% of the total reactions obey the MHP when the geometric means of the η values are considered, while it is violated in nine reactions (see Table 1). Among these nine reactions, in two reactions (reaction number 29 and 45) the ηgeo values in the reactant and product sides are almost same. For these nine reactions which do not follow the MHP, two major observations can be made: (1) In these reactions, the hardest species like CH4, CF4, CH2Cl2, and CH3F lie on the reactant side, and (2) species with very small η values are found on the product side, which makes the geometric mean of the η on the product side lower than that of the reactant side. However, the comparison of ηcom between the reactants and products reduces the success of the MHP sharply. It is found that only 23 chemical reactions leading to just 46% of the total 50 reactions obey the MHP. It may be noted that in many cases the Imin and Amax are associated with the same reactant and/or product, which leads to the wrong trend. Therefore, for these cases we discourage the use of ηcom to analyze the validity of the MHP. We have also tested the validity of the criterion that the hardest species lies on the product side. It is noted that only six reactions among the reactions in this set violate this criterion (88% of the total reactions obey). Interestingly, four of nine reactions, which previously disobeyed the MHP based on their ηgeo values, follow this criterion. In one case (reaction number 34), although ηgeo of products is larger than that of the reactants, the hardest species, (CH3)2CO, lies on the reactant side.

3.2. Results at the LC-BLYP Level

Despite large alteration in η values at the LC-BLYP level compared to those at B3LYP level, the number of reactions obeying the MHP does not change drastically. In fact, the total number of reactions obeying the MHP increases slightly at this level. 42 of the total 50 reactions are found to have ηgeo values higher on the product side than that on the reactant side (see Table 2). Thus, 84% of the total reactions considered here are found to obey the MHP. Importantly, five of the nine reactions for which the MHP fails at the B3LYP level also give similar results at the LC-BLYP level. Similar to the previous level, here the use of the ηcom value yields discouraging results where only 21 (i.e., 42%) reactions have ηcom values higher on the product side than those on the reactant side. On the other hand, only eight reactions have the hardest species on the reactant side. Note that five of these eight reactions also contradict the MHP based on the ηgeo values.

3.3. Test of the Null Hypothesis

Among the 50 reactions considered in Table 1, 41 reactions are found to obey the MHP at B3LYP/6-311++G(2df,3pd)/def2-QZVP level of theory. Thus, 82.0% reactions obey the MHP. Now we have tested the validity of the 50% null hypothesis for these set of reactions in Table 1. Let P0 be the population proportion of the chemical reactions mentioned above, and let us check the null hypothesis
H0: P0 = 0.5
against the alternative hypothesis,
H1: P0 ≠ 0.5
We have fixed the level of significance at 1% to check the null hypothesis. Thus, here the level of confidence is 99%. Note that the sample proportion (P) is binomial with mean P0 and variance P0(1 − P0)/n. Here, the sample size is 50, and the distribution of the test statistic calculated by the following expression,
Z = ( P P 0 ) P 0 ( 1 P 0 ) n
which may be approximated by the standard normal distribution in case the null hypothesis turns out to be true.
The null hypothesis will be rejected at the 1% level of significance if the absolute value of the calculated Z is greater than Z0.005(=2.578). The calculated value for Z is found to be 4.525, which is greater than Z0.005. Hence, the 50% null hypothesis is rejected at the 1% level of significance and it may be concluded that the proportion of reactions cannot be considered to be equal to 0.5.
Now let us combine the present set of organic type reactions with the previously reported inorganic-based reactions [68] to have a large set of total 151 reactions studied at the B3LYP/6-311++G(2df,3pd)/def2-QZVP level. Among these 151 reactions considered, 103 reactions have a geometric mean of the hardness of the products higher than the geometric mean of hardness of the reactants. Thus, 68.2% of reactions obey the MHP. Now we have tested the validity of 50% null hypothesis for the present sample size of 151 reactions. The Z-value is found to be 4.473, which is greater than Z0.005(2.578). Hence, the 50% null hypothesis is rejected at the 1% level of significance and it may be concluded that the proportion of reactions cannot be considered to be equal to 0.5.
Thus, the level of the significance remains more or less same in the case of the 50 reactions as well as the larger set of the reactions with a sample size of 151. Moreover, it is known from ref. [68] that the calculated hardness values differ significantly by changing the approximate formulas used, the quality of the basis set and the level of theory used. In the present cases, both the levels provide qualitatively similar trends regarding the validity of the MHP.

4. Conclusions

We have studied 50 exothermic reactions, most of which are organic in nature, exhibiting anomeric effect to investigate whether the maximum hardness principle (MHP) is valid. The geometric mean values of hardness and combined hardness of the reactants and the products are used. The calculations are carried out at the B3LYP/6-311++G(2df,3pd)/def2-QZVP and LC-BLYP/6-311++G(2df,3pd)/def2-QZVP (def2-QZVP for iodine and Hg) theory level to evaluate the effect of level of theory on the validation of the MHP in a given set of reactions. The results from the B3LYP/6-311++G(2df,3pd)/def2-QZVP level show that 82% of the studied reactions obey the MHP as they have a higher geometric mean of the hardness values on the product side compared to that on the reactant side. However, when the combined hardness is considered, only 46% of the chemical reactions obey the MHP at the same level of theory. The results at the LC-BLYP/6-311++G(2df,3pd)/def2-QZVP level are even marginally better. The geometric mean consideration shows that 84% of the chemical reactions obey the MHP, while the combined hardness values show that only 42% of the chemical reactions follow the MHP. At both levels, the number of reactions with the hardest species on the product side is reasonably high (44 at B3LYP and 42 at LC-BLYP out of 50 reactions). In the case where a total of 151 reactions are considered, the 50% null hypothesis is rejected at the 1% level of significance. Therefore, the validity of the MHP in so many reactions is not fortuitous.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/21/11/1477/s1. The calculated ionization potential (I, eV) and electron affinity (A, eV) at B3LYP/6-311++G(2df,3pd)/def2-QZVP and LC-BLYP/6-311++G(2df,3pd)/def2-QZVP levels for the molecules are provided in the supplementary materials.

Acknowledgments

P. K. Chattaraj thanks Luis R. Domingo and Alessandro Ponti for kindly inviting him to contribute an article in this Special Issue of “Molecules” on “Density Functional Theory and Reactivity Indices: Applications in Organic Chemical Reactivity”. He would also like to thank DST, New Delhi, for the J. C. Bose National Fellowship. R. Saha thanks UGC, New Delhi for his Senior Research Fellowship.

Author Contributions

P.K.C. has designed the work. R.S. has performed the calculations and has written the first draft of the manuscript. S.P. has helped in preparing the manuscript. All the authors have reviewed this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Proft, F.; Geerlings, P. Conceptual and computational DFT in the study of aromaticity. Chem. Rev. 2001, 101, 1451–1464. [Google Scholar] [CrossRef]
  2. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Applications of the conceptual density functional theory indices to organic chemistry reactivity. Molecules 2016, 21, 748. [Google Scholar] [CrossRef]
  3. Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press: Ithaca, NY, USA, 1960. [Google Scholar]
  4. Sen, K.D.; Jorgenson, C.K. Electronegativity. In Structure and Bonding; Springer: Berlin, Germany, 1987. [Google Scholar]
  5. Chattaraj, P.K. Electronegativity and hardness: A density functional treatment. J. Ind. Chem. Soc. 1992, 69, 173–183. [Google Scholar]
  6. Parr, R.G.; Donnelly, R.A.; Levy, M.; Palke, W.E. Electronegativity: The density functional viewpoint. J. Chem. Phys. 1978, 68, 3801–3807. [Google Scholar] [CrossRef]
  7. Pearson, R.G. Hard and soft acids and bases. J. Am. Chem. Soc. 1963, 85, 3533–3539. [Google Scholar] [CrossRef]
  8. Parr, R.G.; Pearson, R.G. Absolute hardness: Companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. [Google Scholar] [CrossRef]
  9. Sen, K.D. Chemical Hardness. In Structure and Bonding; Springer: Berlin, Germany, 1993. [Google Scholar]
  10. Pearson, R.G. Chemical Hardness. In Applications from Molecules to Solids; Wiley-VCH: Weinheim, Germany, 1997. [Google Scholar]
  11. Parr, R.G.; Szentpaly, L.V.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  12. Chattaraj, P.K.; Sarkar, U.; Roy, D.R. Electrophilicity index. Chem. Rev. 2006, 106, 2065–2091. [Google Scholar] [CrossRef]
  13. Chattaraj, P.K.; Roy, D.R. Update 1 of: Electrophilicity index. Chem. Rev. 2007, 107, PR46–PR75. [Google Scholar] [CrossRef]
  14. Chattaraj, P.K.; Giri, S.; Duley, S. Update 2 of: Electrophilicity index. Chem. Rev. 2011, 111, PR43–PR75. [Google Scholar] [CrossRef]
  15. Pearson, R.G. Hard and Soft Acids and Bases; Dowden, Hutchinson & Ross: Stroudsberg, PA, USA, 1973. [Google Scholar]
  16. Pearson, R.G. Hard and soft acids and bases-the evolution of a chemical concept. Coord. Chem. Rev. 1990, 100, 403–425. [Google Scholar] [CrossRef]
  17. Chattaraj, P.K.; Lee, H.; Parr, R.G. HSAB principle. J. Am. Chem. Soc. 1991, 113, 1855–1856. [Google Scholar] [CrossRef]
  18. Chattaraj, P.K.; Schleyer, P.V.R. An ab initio study resulting in a greater understanding of the HSAB principle. J. Am. Chem. Soc. 1994, 116, 1067–1071. [Google Scholar] [CrossRef]
  19. Pearson, R.G. Recent advances in the concept of hard and soft acids and bases. J. Chem. Educ. 1987, 64, 561–567. [Google Scholar] [CrossRef]
  20. Koopmans, T. Über diezuordnung von wellenfunktionen und eigenwertenzu den einzelnen elektronen eines atoms. Physica 1933, 1, 104–113. [Google Scholar] [CrossRef]
  21. Parr, R.G.; Chattaraj, P.K. Principle of maximum hardness. J. Am. Chem. Soc. 1991, 113, 1854–1855. [Google Scholar] [CrossRef]
  22. Chattaraj, P.K.; Nath, S.; Sannigrahi, A.B. Hardness, chemical potential, and valency profiles of molecules under internal rotations. J. Phys. Chem. 1994, 98, 9143–9145. [Google Scholar] [CrossRef]
  23. Cárdenas-Jirón, G.I.; Toro-Labbé, A. Hardness profile and activation hardness for rotational isomerization processes. 2. The maximum hardness principle. J. Phys. Chem. 1995, 99, 12730–12738. [Google Scholar] [CrossRef]
  24. Cárdenas-Jirón, G.I.; Letelier, J.R.; Toro-Labbé, A. The internal rotation of hydrogen thioperoxide: Energy, chemical potential, and hardness profiles. J. Phys. Chem. A 1998, 102, 7864–7871. [Google Scholar] [CrossRef]
  25. Gutiérrez-Oliva, S.; Letelier, J.R.; Toro-Labbé, A. Energy, chemical potential and hardness profiles for the rotational isomerization of HOOH, HSOH and HSSH. Mol. Phys. 1999, 96, 61–70. [Google Scholar] [CrossRef]
  26. Chattaraj, P.K.; Gutiérrez-Oliva, S.; Jaque, P.; Toro-Labbé, A. Towards understanding the molecular internal rotations and vibrations and chemical reactions through the profiles of reactivity and selectivity indices: An ab initio SCF and DFT study. Mol. Phys. 2003, 101, 2841–2853. [Google Scholar] [CrossRef]
  27. Uchimaru, T.; Chandra, A.K.; Kawahara, S.; Matsumura, K.; Tsuzuki, S.; Mikami, M. Internal bond rotation in substituted methyl radicals, H2B−CH2, H3C−CH2, H2N−CH2, and HO−CH2: Hardness profiles. J. Phys. Chem. A 2001, 105, 1343–1353. [Google Scholar] [CrossRef]
  28. Gutiérrez-Oliva, S.; Toro-Labbé, A. The torsional problem of oxalyl chloride: A challenge for theoretical methods. Chem. Phys. Lett. 2004, 383, 435–440. [Google Scholar] [CrossRef]
  29. Chattaraj, P.K.; Fuentealba, P.; Jaque, P.; Toro-Labbé, A. Validity of the minimum polarizability principle in molecular vibrations and internal rotations: An ab initio SCF study. J. Phys. Chem. A 1999, 103, 9307–9312. [Google Scholar] [CrossRef]
  30. Pearson, R.G.; Palke, W.E. Support for a principle of maximum hardness. J. Phys. Chem. 1992, 96, 3283–3285. [Google Scholar] [CrossRef]
  31. Makov, G. Chemical hardness in density functional theory. J. Phys. Chem. 1995, 99, 9337–9339. [Google Scholar] [CrossRef]
  32. Pal, S.; Naval, N.; Roy, R. Principle of maximum hardness: An accurate ab initio study. J. Phys. Chem. 1993, 97, 4404–4406. [Google Scholar] [CrossRef]
  33. Chattaraj, P.K.; Nath, S.; Sannigrahi, A.B. Ab initio SCF study of maximum hardness and maximum molecular valency principles. Chem. Phys. Lett. 1993, 212, 223–230. [Google Scholar] [CrossRef]
  34. Jaque, P.; Toro-Labbé, A. Theoretical study of the double proton transfer in the CHX−XH···CHX−XH (X = O, S) complexes. J. Phys. Chem. A 2000, 104, 995–1003. [Google Scholar] [CrossRef]
  35. Datta, D. “Hardness Profile” of a reaction path. J. Phys. Chem. 1992, 96, 2409–2410. [Google Scholar] [CrossRef]
  36. Kar, T.; Scheiner, S. Hardness profiles of some 1,2-hydrogen shift reactions. J. Phys. Chem. 1995, 99, 8121–8124. [Google Scholar] [CrossRef]
  37. Chattaraj, P.K.; Cedillo, A.; Parr, R.G.; Arnett, E.M. Appraisal of chemical bond making, bond breaking, and electron transfer in solution in the light of the principle of maximum hardness. J. Org. Chem. 1995, 60, 4707–4714. [Google Scholar] [CrossRef]
  38. Toro-Labbé, A. Characterization of chemical reactions from the profiles of energy, chemical potential, and hardness. J. Phys. Chem. A 1999, 103, 4398–4403. [Google Scholar] [CrossRef]
  39. Hohm, U. Is there a minimum polarizability principle in chemical reactions? J. Phys. Chem. A 2000, 104, 8418–8423. [Google Scholar] [CrossRef]
  40. Jaque, P.; Toro-Labbé, A. Characterization of copper clusters through the use of density functional theory reactivity descriptors. J. Chem. Phys. 2002, 117, 3208–3218. [Google Scholar] [CrossRef]
  41. Pérez, P.; Toro-Labbé, A. Characterization of keto-enol tautomerism of acetyl derivatives from the analysis of energy, chemical potential, and hardness. J. Phys. Chem. A 2000, 104, 1557–1562. [Google Scholar] [CrossRef]
  42. Chattaraj, P.K.; Pérez, P.; Zevallos, J.; Toro-Labbé, A. Ab initio SCF and DFT studies on solvent effects on intramolecular rearrangement reactions. J. Phys. Chem. A 2001, 105, 4272–4283. [Google Scholar] [CrossRef]
  43. Domingo, L.R.; Pérez, P. Global and Local Reactivity Indices for Electrophilic/Nucleophilic Free Radicals. Org. Biomol. Chem. 2013, 11, 4350–4358. [Google Scholar] [CrossRef]
  44. Domingo, L.R. Applications of Reactivity Indices based on Density Functional Theory to the Study of Organic Reactions. The Case of the Diels-Alder Reaction. Lett. Org. Chem. 2011, 8, 81. [Google Scholar] [CrossRef]
  45. Aurell, M.J.; Domingo, L.R.; Perez, P.; Contreras, R.A. Theoretical study on the regioselectivity of 1,3-dipolar cycloadditions using DFT-based reactivity indexes. Tetrahedron 2004, 60, 11503–11509. [Google Scholar] [CrossRef]
  46. Chattaraj, P.K.; Fuentealba, P.; Gómez, B.; Contreras, R. Woodward−hoffmann rule in the light of the principles of maximum hardness and minimum polarizability: DFT and Ab initio SCF studies. J. Am. Chem. Soc. 2000, 122, 348–351. [Google Scholar] [CrossRef]
  47. De Proft, F.; Chattaraj, P.K.; Ayers, P.W.; Torrent-Sucarrat, M.; Elango, M.; Subramanian, V.; Giri, S.; Geerlings, P. Initial hardness response and hardness profiles in the study of Woodward–Hoffmann rules for electrocyclizations. J. Chem. Theor. Comput. 2008, 4, 595–602. [Google Scholar] [CrossRef]
  48. Zhou, Z.; Parr, R.G. New measures of aromaticity: Absolute hardness and relative hardness. J. Am. Chem. Soc. 1989, 111, 7371–7379. [Google Scholar] [CrossRef]
  49. Chattaraj, P.K.; Roy, D.R.; Elango, M.; Subramanian, V. Chemical reactivity descriptor based aromaticity indices applied to Al42− and Al44− systems. J. Mol. Struct. (Theochem.) 2006, 759, 109–110. [Google Scholar] [CrossRef]
  50. Chattaraj, P.K.; Roy, D.R.; Elango, M.; Subramanian, V. Stability and reactivity of all-metal aromatic and antiaromatic systems in light of the principles of maximum hardness and minimum polarizability. J. Phys. Chem. A 2005, 109, 9590–9597. [Google Scholar] [CrossRef]
  51. Harbola, M.K. Magic numbers for metallic clusters and the principle of maximum hardness. Proc. Natl. Acad. Sci. USA 1992, 89, 1036–1039. [Google Scholar] [CrossRef]
  52. Parr, R.G.; Zhou, Z. Absolute hardness: Unifying concept for identifying shells and subshells in nuclei, atoms, molecules, and metallic clusters. Acc. Chem. Res. 1993, 26, 256–258. [Google Scholar] [CrossRef]
  53. Chattaraj, P.K.; Maiti, B. Electronic structure principles and atomic shell structure. J. Chem. Educ. 2001, 78, 811–813. [Google Scholar] [CrossRef]
  54. Chattaraj, P.K.; Sengupta, S. Popular electronic structure principles in a dynamical context. J. Phys. Chem. 1996, 100, 16126–16130. [Google Scholar] [CrossRef]
  55. Deb, B.M.; Chattaraj, P.K.; Mishra, S. Time-dependent quantum-fluid density-functional study of high-energy proton-helium collisions. Phys. Rev. A 1991, 43, 1248–1257. [Google Scholar] [CrossRef]
  56. Chattaraj, P.K.; Maiti, B. HSAB Principle Applied to the time evolution of chemical reactions. J. Am. Chem. Soc. 2003, 125, 2705–2710. [Google Scholar] [CrossRef]
  57. Chattaraj, P.K.; Poddar, A. A density functional treatment of chemical reactivity and the associated electronic structure principles in the excited electronic states. J. Phys. Chem. A 1998, 102, 9944–9948. [Google Scholar] [CrossRef]
  58. Mineva, T.; Sicilia, E.; Russo, N. Density-functional approach to hardness evaluation and its use in the study of the maximum hardness principle. J. Am. Chem. Soc. 1998, 120, 9053–9058. [Google Scholar] [CrossRef]
  59. Chattaraj, P.K.; Sengupta, S. Chemical hardness as a possible diagnostic of the chaotic dynamics of rydberg atoms in an external field. J. Phys. Chem. A 1999, 103, 6122–6126. [Google Scholar] [CrossRef]
  60. Jayakumar, N.; Kolanadaivel, P. Studies of isomer stability using the maximum hardness principle (MHP). Int. J. Quantum Chem. 2000, 76, 648–655. [Google Scholar] [CrossRef]
  61. Anandan, K.; Kolandaivel, P.; Kumaresan, R. Quantum chemical studies on molecular structural conformations and hydrated forms of salicylamide and O-hydroxybenzoyl cyanide. Int. J. Quantum Chem. 2005, 104, 286–298. [Google Scholar] [CrossRef]
  62. Gomez, B.; Fuentealba, P.; Contreras, R. The maximum hardness and minimum polarizability principles as the basis for the study of reaction profiles. Theor. Chem. Acc. 2003, 110, 421–427. [Google Scholar] [CrossRef]
  63. Selvarengan, P.; Kolandaivel, P. Potential energy surface study on glycine, alanine and their zwitterionic forms. J. Mol. Struct. 2004, 671, 77–86. [Google Scholar] [CrossRef]
  64. Noorizadeh, S. The maximum hardness and minimum polarizability principles in accordance with the Bent rule. J. Mol. Struct. 2005, 713, 27–32. [Google Scholar] [CrossRef]
  65. Zhang, Y.-L.; Yang, Z.-Z. Tautomerism and the maximum hardness principle. Int. J. Quantum Chem. 2006, 106, 1723–1735. [Google Scholar] [CrossRef]
  66. Noorizadeh, S. Is there a minimum electrophilicity principle in chemical reactions? Chin. J. Chem. 2007, 25, 1439–1444. [Google Scholar] [CrossRef]
  67. Poater, J.; Swart, M.; Solà, M. An assessment of the validity of the maximum hardness principle in chemical reactions. J. Mex. Chem. Soc. 2012, 56, 311–315. [Google Scholar]
  68. Pan, S.; Sola, M.; Chattaraj, P.K. On the validity of the maximum hardness principle and the minimum electrophilicity principle during chemical reactions. J. Phys. Chem. A 2013, 117, 1843–1852. [Google Scholar] [CrossRef]
  69. Hati, S.; Datta, D. Anomeric effect and hardness. J. Org. Chem. 1992, 57, 6057–6060. [Google Scholar] [CrossRef]
  70. Pearson, R.G. Chemical hardness and bond dissociation energies. J. Am. Chem. Soc. 1988, 110, 7684–7690. [Google Scholar] [CrossRef]
  71. Reed, A.E.; Schleyer, P.V.R. The anomeric effect with central atoms other than carbon. 1. Strong interactions between nonbonded substituents in polyfluorinated first- and second-row hydrides. J. Am. Chem. Soc. 1987, 109, 7362–7373. [Google Scholar] [CrossRef]
  72. Benson, S.W. Electrostatics, the chemical bond and molecular stability. Angew. Chem. Int. Ed. Engl. 1978, 17, 812–819. [Google Scholar] [CrossRef]
  73. Dennington, R.; Keith, T.; Millam, J. GaussView, version 5; Semichem, Inc.: Shawnee Mission, KS, USA, 2009. [Google Scholar]
  74. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem.Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  75. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  76. Lee, T.J.; Taylor, P.R. A diagnostic for determining the quality of single-reference electron correlation methods. J. Quantum Chem. Quantum Chem. Symp. 1989, S23, 199–207. [Google Scholar] [CrossRef]
  77. Becke, A.D. Correlation energy of an inhomogeneous electron gas: A coordinate-space model. J. Chem. Phys. 1988, 88, 1053–1062. [Google Scholar] [CrossRef]
  78. Tawada, Y.; Tsuneda, T.; Yanagisawa, S.; Yanai, T.; Hirao, K. A long-range-corrected time-dependent density functional theory. J. Chem. Phys. 2004, 120, 8425–8433. [Google Scholar] [CrossRef]
  79. Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K. A long-range correction scheme for generalized-gradient-approximation exchange functional. J. Chem. Phys. 2001, 115, 3540–3544. [Google Scholar] [CrossRef]
  80. Singh, R.K.; Tsuneda, T. Reaction energetics on long-range corrected density functional theory: Diels–Alder reactions. J. Comp. Chem. 2013, 34, 379–386. [Google Scholar] [CrossRef]
  81. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  82. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11−18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  83. Weigend, F.; Baldes, A. Segmented contracted basis sets for one- and two-component Dirac? Fock effective core potentials. J. Chem. Phys. 2010, 133, 174102. [Google Scholar] [CrossRef]
  84. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  85. Zhou, Z.; Parr, R.G. Activation hardness: New index for describing the orientation of electrophilic aromatic substitution. J. Am. Chem. Soc. 1990, 112, 5720–5724. [Google Scholar] [CrossRef]
  86. Chattaraj, P.K. The maximum hardness principle: An overview. Proc. Indian Natl. Sci. Acad. 1996, 62A, 513–531. [Google Scholar]
  • Sample Availability: Not Available.
Table 1. Hardness (η, eV) values of the individual reactants and products, geometric mean (ηgeo, eV) of the hardness values and combined hardness values (ηcom, eV) for the reactant and product sides at the B3LYP/6-311++G(2df,3pd)/def2-QZVP (def2-QZVP for iodine and Hg) level for the chemical reactions. The enthalpy changes (∆H) for the reactions are in kcal/mol.
Table 1. Hardness (η, eV) values of the individual reactants and products, geometric mean (ηgeo, eV) of the hardness values and combined hardness values (ηcom, eV) for the reactant and product sides at the B3LYP/6-311++G(2df,3pd)/def2-QZVP (def2-QZVP for iodine and Hg) level for the chemical reactions. The enthalpy changes (∆H) for the reactions are in kcal/mol.
Reactionsηgeo(R)ηgeo(P)P > R?ηcom(R)ηcom(P)P > R?∆H #
1.2CH3FCH4+CH2F29.4910.12Yes9.499.64Yes−14 [69]
9.49 10.62 9.64
2.2CH3OHCH4+CH2(OH)27.419.02Yes7.417.66Yes−15 ± 5 [69]
7.41 10.62 7.66
3.CH3F+CH3OHCH4+FCH2OH8.399.55Yes7.418.58Yes−ve [69]
9.49 7.41 10.62 8.58
4.CH3NH2+CH3OHCH4+HOCH2NH26.868.28Yes6.316.46Yes−ve [69]
6.35 7.41 10.62 6.46
5.2CH3NH2CH4+CH2(NH2)26.358.00Yes6.356.03No−ve [69]
6.35 10.62 6.03
6.CH3NH2+CH3FCH4+FCH2NH27.778.93Yes6.357.51Yes−ve [69]
6.35 9.49 10.62 7.51
7.2SiH3FSiH4+SiH2F28.949.47Yes8.949.30Yes−8 [69]
8.94 9.54 9.41
8.2CF2Cl2CF4+CCl47.849.22Yes7.846.76No−16.3 [69]
7.84 12.59 6.76
9.3CH3F2CH4+CHF39.4910.78Yes9.4910.62Yes−31.4 [69]
9.49 10.62 11.11
10.4CHF3CH4+3CF411.1112.07Yes11.1110.62No−22.9 [69]
11.11 10.62 12.59
11.4CH3F3CH4+CF49.4911.08Yes9.4910.62Yes−63 [69]
9.49 10.62 12.59
12.4CH3Cl3CH4+CCl47.819.49Yes7.816.76No−6 [69]
7.81 10.62 6.76
13.4CH3OCH33CH4+C(OCH3)47.029.77Yes7.027.59Yes−52 [69]
7.02 10.62 7.59
14.4CF3Cl3CF4+CCl49.1110.78Yes9.116.76No−27.1 [69]
9.11 12.59 6.76
15.4CH3CH33CH4+C(CH3)49.2510.04Yes9.258.47No−13 [69]
9.25 10.62 8.47
16.4SiH3F3SiH4+SiF48.9410.04Yes8.948.85No−23 [69]
8.94 9.54 11.72
17.SiF3H+CF4SiF4+CF3H11.8311.41No11.1110.44No−37 [69]
11.11 12.59 11.72 11.11
18.C(OCH3)4+SiH4CH4+Si(OCH3)48.518.98Yes7.597.60Yes−144 [69]
7.59 9.54 10.62 7.59
19.2CH3OHH2O+(CH3)2O7.417.60Yes7.416.64No−6.0 [72]
7.41 8.23 7.02
20.2HOFH2O+F2O7.097.39Yes7.095.82No−5 ± 3 [72]
7.09 8.23 6.63
21.2HOClH2O+Cl2O5.995.90No5.994.22No−1 ± 1 [72]
5.99 8.23 4.22
22.2CH3SHH2S+(CH3)2S6.116.28Yes6.115.55No−2.8 [72]
6.11 6.78 5.82
23.2HSSHH2S+(HS)2S4.376.09Yes4.375.47Yes−4.4 [72]
4.37 6.78 5.47
24.2CH3NH2NH3+(CH3)2NH6.356.47Yes6.355.77No−4.5 [72]
6.35 7.05 5.94
25.2(CH3)2NH(CH3)3N+CH3NH25.946.01Yes5.945.65No−2.4 [72]
5.94 5.69 6.35
26.2CH3CH3CH4+(CH3)2CH29.259.67Yes9.258.80No−2.5 [72]
9.25 10.62 8.80
27.2CF3HCH2F2+CF411.1111.02No11.119.64No−0 ± 2 [72]
11.11 9.64 12.59
28.2CH3ClCH4+CH2Cl27.819.03Yes7.817.68No−1.5 ± 1.4 [72]
7.81 10.62 7.68
29.CH2Cl2+CCl42CHCl3 7.207.19No6.437.19Yes−2.7 ± 1 [72]
7.68 6.76 7.19
30.CH4+CH2I22CH3I 7.215.77No4.895.77Yes−4.4 [72]
10.62 4.89 5.77
31.2H(CH3)C=CH2H2C=CH2+(CH3)2C=CH26.946.98Yes6.945.56No−1.3 [72]
6.94 7.40 6.60
32.2HFC=CF2H2C=CF2+F2C=CF27.377.42Yes7.377.34No−4.5 ± 4 [72]
7.37 7.50 7.34
33.H2C=CCl2+Cl2C=CCl22HClC=CCl2 6.126.74Yes5.826.74Yes−2 ± 4 [72]
6.43 5.82 6.74
34.H2CO+(CH3)2CO2CH3CHO 6.106.25Yes5.316.25Yes−1.7 ± 1.5 [72]
5.94 6.26 6.25
35.COCl2+(CH3)2CO2Cl(CH3)CO 6.616.98Yes4.986.98Yes−13.5 [72]
6.98 6.26 6.98
36.(CH3)2Hg+Cl2Hg2MeHgCl 6.116.79Yes3.946.79Yes−12 ± 3 [72]
6.39 5.84 6.79
37.2COSCO2+CS27.247.32Yes7.245.38No0.0 [72]
7.24 9.95 5.38
38.2CH2COCO2+CH2=C=CH25.668.58Yes5.667.00Yes−27 ± 2 [72]
5.66 9.95 7.40
39.2CH2COCO2+CH45.6610.28Yes5.667.24Yes−60 ± 2 [72]
5.66 9.95 10.62
40.2CH2CSCS2+CH44.017.56Yes4.015.38Yes−30 ± 4 [72]
4.01 5.38 10.62
41.2COCl2CO2+CCl46.988.20Yes6.986.76No−12 [72]
6.98 9.95 6.76
42.2COF2CO2+CF49.2611.20Yes9.269.95Yes−14 ± 4 [72]
9.26 9.95 12.59
43.HC(OH)3HCOOH+H2O8.117.99No8.117.75No−7 [72]
8.11 7.75 8.23
44.SiH3Cl+CH3FSiH3F+CH3Cl8.788.36No8.637.56No−20 [70]
8.13 9.49 8.94 7.81
45.SiH3PH2+CH3NH2SiH3NH2+CH3PH26.626.60No6.176.46Yes−16 [70]
6.91 6.35 6.68 6.53
46.SiH3SiH3+CH3CH32SiH3CH3 8.548.67Yes7.898.67Yes−7 [70]
7.89 9.25 8.67
47.SiH3I+HOFSiH3F+HOI6.656.02No5.454.05No−80 [70]
6.25 7.09 8.94 4.05
48.2CH2F2CHF3+CH3F9.6410.27Yes9.649.49No−8.5 [71]
9.64 11.11 9.49
49.2NH2FNHF2+NH37.517.69Yes7.517.01No−10.0 [71]
7.51 8.39 7.05
50.2NHF2NF3+NH2F8.398.66Yes8.397.51No−6.0 [71]
8.39 9.99 7.51
# The reference numbers from where the enthalpy values are taken are given in the square brackets. Hardness values (in eV) are provided below each molecule.
Table 2. Hardness (η, eV) values of the individual reactants and products, geometric mean (ηgeo, eV) of the hardness values and combined hardness values (ηcom, eV) for the reactant and product sides at the LC-BLYP/6-311++G(2df,3pd)/def2-QZVP (def2-QZVP for iodine and Hg) level of theory for the chemical reactions. The enthalpy changes (∆H) for the reactions are in kcal/mol.
Table 2. Hardness (η, eV) values of the individual reactants and products, geometric mean (ηgeo, eV) of the hardness values and combined hardness values (ηcom, eV) for the reactant and product sides at the LC-BLYP/6-311++G(2df,3pd)/def2-QZVP (def2-QZVP for iodine and Hg) level of theory for the chemical reactions. The enthalpy changes (∆H) for the reactions are in kcal/mol.
Reactionsηgeo(R)ηgeo(P)P > R?ηcom(R)ηcom(P)P > R?∆H #
1.2CH3FCH4+CH2F214.1014.72Yes14.1014.21Yes−14 [69]
14.10 15.22 14.21
2.2CH3OHCH4+CH2(OH)212.0813.81Yes12.0812.52Yes−15 ± 5 [69]
12.08 15.22 12.52
3.CH3F+CH3OHCH4+FCH2OH13.0514.24Yes12.0813.32Yes−ve [69]
14.10 12.08 15.22 13.32
4.CH3NH2+CH3OHCH4+HOCH2NH211.4212.93Yes10.7710.98Yes−ve [69]
10.80 12.08 15.22 10.98
5.2CH3NH2CH4+CH2(NH2)210.8012.62Yes10.8010.47No−ve [69]
10.80 15.22 10.47
6.CH3NH2+CH3FCH4+FCH2NH212.3413.58Yes10.8012.11Yes−ve [69]
10.80 14.10 15.22 12.11
7.2SiH3FSiH4+SiH2F213.9914.19Yes13.9914.03Yes−8 [69]
13.99 14.03 14.34
8.2CF2Cl2CF4+CCl414.0115.33Yes14.0113.16No−16.3 [69]
14.01 17.85 13.16
9.3CH3F2CH4+CHF314.1015.41Yes14.1015.22Yes−31.4 [69]
14.10 15.22 15.78
10.4CHF3CH4+3CF415.7817.16Yes15.7815.22No−22.9 [69]
15.78 15.22 17.85
11.4CH3F3CH4+CF414.1015.84Yes14.1015.22Yes−63 [69]
14.10 15.22 17.85
12.4CH3Cl3CH4+CCl412.4114.68Yes12.4113.16Yes−6 [69]
12.41 15.22 13.16
13.4CH3OCH33CH4+C(OCH3)411.4814.68Yes11.4812.12Yes−52 [69]
11.48 15.22 12.12
14.4CF3Cl3CF4+CCl414.8916.54Yes14.8913.16No−27.1 [69]
14.89 17.85 13.16
15.4CH3CH33CH4+C(CH3)413.6514.60Yes13.6512.89No−13 [69]
13.65 15.22 12.89
16.4SiH3F3SiH4+SiF413.9914.79Yes13.9913.89No−23 [69]
13.99 14.03 17.31
17.SiF3H+CF4SiF4+CF3H16.9016.53No16.0015.77No−37 [69]
16.00 17.85 17.31 15.78
18.C(OCH3)4+SiH4CH4+Si(OCH3)413.0413.58Yes12.1212.11No−144 [69]
12.12 14.03 15.22 12.11
19.2CH3OHH2O+(CH3)2O12.0812.41Yes12.0811.40No−6.0 [69]
12.08 13.42 11.48
20.2HOFH2O+F2O13.5413.81Yes13.541.42No−5 ± 3 [69]
13.54 13.42 14.22
21.2HOClH2O+Cl2O11.7912.02Yes11.7910.76No−1 ± 1 [69]
11.79 13.42 10.76
22.2CH3SHH2S+(CH3)2S10.5810.68Yes10.589.95No−2.8 [69]
10.58 11.36 10.04
23.2HSSHH2S+(HS)2S11.2711.19No11.2711.02No−4.4 [69]
11.27 11.36 11.02
24.2CH3NH2NH3+(CH3)2NH10.8011.00Yes10.8010.26No−4.5 [69]
10.80 11.80 10.26
25.2(CH3)2NH(CH3)3N+CH3NH210.2610.37Yes10.269.96No−2.4 [69]
10.26 9.97 10.80
26.2CH3CH3CH4+(CH3)2CH213.6514.15Yes13.6513.14No−2.5 [69]
13.65 15.22 13.14
27.2CF3HCH2F2+CF415.7815.93Yes15.7814.21No−0 ± 2 [69]
15.78 14.21 17.85
28.2CH3ClCH4+CH2Cl212.4113.87Yes12.4112.64Yes−1.5 ± 1.4 [69]
12.41 15.22 12.64
29.CH2Cl2+CCl42CHCl3 12.9012.88No12.6412.88Yes−2.7 ± 1 [72]
12.64 13.16 12.88
30.CH4+CH2I22CH3I 12.5310.74No10.3110.74Yes−4.4 [72]
15.22 10.31 10.74
31.2H(CH3)C=CH2H2C=CH2+(CH3)2C=CH211.2411.42Yes11.2410.82No−1.3 [72]
11.24 12.06 10.82
32.2HFC=CF2H2C=CF2+F2C=CF211.8812.02Yes11.8811.81No−4.5 ± 4 [72]
11.88 11.94 12.11
33.H2C=CCl2+Cl2C=CCl22HClC=CCl2 11.4311.47Yes11.1211.47Yes−2 ± 4 [72]
11.49 11.36 11.47
34.H2CO+(CH3)2CO2CH3CHO 11.5311.51No11.1411.51Yes−1.7 ± 1.5 [72]
11.94 11.14 11.51
35.COCl2+(CH3)2CO2Cl(CH3)CO 12.1212.40Yes11.1412.40Yes−13.5 [72]
13.19 11.14 12.40
36.(CH3)2Hg+Cl2Hg2MeHgCl 11.3212.04Yes9.8912.04Yes−12 ± 3 [72]
10.79 11.88 12.04
37.2COSCO2+CS212.6312.55No12.6310.49No0.0 [72]
12.63 15.02 10.49
38.2CH2COCO2+CH2=C=CH211.0113.25Yes11.0111.69Yes−27 ± 2 [72]
11.01 15.02 11.69
39.2CH2COCO2+CH411.1015.12Yes11.0212.41Yes−60 ± 2 [72]
11.01 15.02 15.22
40.2CH2CSCS2+CH49.4912.64Yes9.4910.49Yes−30 ± 4 [72]
9.49 10.49 15.22
41.2COCl2CO2+CCl413.1914.06Yes13.1913.16No−12 [72]
13.19 15.02 13.16
42.2COF2CO2+CF415.0916.38Yes15.0915.02No−14 ± 4 [72]
15.09 15.02 17.85
43HC(OH)3HCOOH+H2O12.8313.15Yes12.8312.72No−7 [72]
12.83 12.89 13.42
44.SiH3Cl+CH3FSiH3F+CH3Cl13.6213.18No12.9612.41No−20 [70]
13.16 14.10 13.99 12.41
45.SiH3PH2+CH3NH2SiH3NH2+CH3PH210.9711.04Yes10.8010.74No−16 [70]
11.14 10.80 11.33 10.74
46.SiH3SiH3+CH3CH32SiH3CH3 13.0713.20Yes12.1413.20Yes−7 [70]
12.14 13.65 13.20
47.SiH3I+HOFSiH3F+HOI12.3811.72No11.209.82No−80 [70]
11.31 13.54 13.99 9.82
48.2CH2F2CHF3+CH3F14.2114.92Yes14.2114.10No−8.5 [71]
14.21 15.78 14.10
49.2NH2FNHF2+NH312.2412.42Yes12.2411.74No−10.0 [71]
12.24 13.07 11.80
50.2NHF2NF3+NH2F13.0714.09Yes13.0712.24No−6.0 [71]
13.07 16.22 12.24
# The reference numbers from where the enthalpy values are taken are given within the square brackets. Hardness values (in eV) are provided below the molecules.

Share and Cite

MDPI and ACS Style

Saha, R.; Pan, S.; Chattaraj, P.K. Statistical Significance of the Maximum Hardness Principle Applied to Some Selected Chemical Reactions. Molecules 2016, 21, 1477. https://doi.org/10.3390/molecules21111477

AMA Style

Saha R, Pan S, Chattaraj PK. Statistical Significance of the Maximum Hardness Principle Applied to Some Selected Chemical Reactions. Molecules. 2016; 21(11):1477. https://doi.org/10.3390/molecules21111477

Chicago/Turabian Style

Saha, Ranajit, Sudip Pan, and Pratim K. Chattaraj. 2016. "Statistical Significance of the Maximum Hardness Principle Applied to Some Selected Chemical Reactions" Molecules 21, no. 11: 1477. https://doi.org/10.3390/molecules21111477

Article Metrics

Back to TopTop