Next Article in Journal / Special Issue
The Mechanism of Phosphoryl Transfer Reaction and the Role of Active Site Residues on the Basis of Ribokinase-Like Kinases
Previous Article in Journal
Molecular Dynamics Simulations of the O-glycosylated 21-residue MUC1 Peptides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electronic Density Approaches to the Energetics of Noncovalent Interactions

Department of Chemistry University of New Orleans, New Orleans, LA 70148, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2004, 5(4), 130-140; https://doi.org/10.3390/i5040130
Submission received: 15 December 2003 / Accepted: 11 March 2004 / Published: 1 April 2004
(This article belongs to the Special Issue Proceedings of the Workshop on Modeling Interaction in Biomolecules)

Abstract

:
We present an overview of procedures that have been developed to compute several energetic quantities associated with noncovalent interactions. These formulations involve numerical integration over appropriate electronic densities. Our focus is upon the electrostatic interaction between two unperturbed molecules, the effect of the polarization of each charge distribution by the other, and the total energy of interaction. The expression for the latter is based upon the Hellmann-Feynman theorem. Applications to a number of systems are discussed; among them are dimers of uracil and interacting pairs of molecules in the crystal lattice of the energetic compound RDX.

Introduction

Noncovalent interactions are ubiquitous: between enzymes and substrates, in hydrogen bonding, physical adsorption, solvation, condensation processes, etc. Calculating the energy associated with such an interaction is, in principle, straightforward; if a system M is formed from N components Mi, then the stabilization energy of M is,
Δ E stab = E M i = 1 N E M i
where EM and E M i are the respective equilibrium, ground-state energies. Eq. (1) is exact. However it suffers from the fact that ΔEstab is typically several orders of magnitude smaller than EM and the E M i ; thus, barring fortuitous cancellation, any errors in these quantities will be greatly magnified in ΔEstab. This problem can of course be minimized by computing EM and the E M i at high levels of accuracy, but this is likely to be prohibitively expensive in terms of processing resources for many systems of practical interest.
Another difficulty with eq. (1) is the so-called basis set superposition error (BSSE). This refers to the size-imbalance between the basis sets used for M and the Mi, which are smaller. The result is an artificial stabilization of M [1,2,3]. The effect diminishes with the use of larger basis sets, but this solution again involves increased computational cost. Some time ago, Boys and Bernardi suggested addressing BSSE by introducing “ghost” orbitals in treating the Mi [4]; this is known as the counterpoise procedure. It is now widely used, although there has been considerable controversy in the past concerning its effectiveness [1,3].
The use of eq. (1) in connection with noncovalent interactions is sometimes denoted the ab initio or supermolecular approach. For more extensive discussions, see Chalasinski and Szczesniak [5] and Rappe and Bernstein [6].
The problem of achieving sufficient accuracy with eq. (1) has made perturbation theory an attractive alternative [7,8,9]. This directly produces the interaction energy Eint between the components Mi, without the need to take differences between large numbers. However the Mi are normally assumed to retain their ground-state geometries; no account is taken of any changes in these that may accompany the interaction [5,7]. In contrast, ΔEstab, eq. (1), is obtained using the optimized structures of M and the Mi. Thus ΔEstab and Eint differ by the energy involved in any geometry changes that occur. This may be quite small, however, as shall be shown later.
In the perturbation theory formulation of Eint, it is expressed as the sum of a series of terms. This is frequently considered to be an advantage, since these can be assigned physical interpretations. For example, Eint is often viewed as composed of four elements:
1)
the electrostatic interaction between the unperturbed Mi;
2)
their mutual polarization of each other’s charge distributions;
3)
dispersion effects, involving intermolecular electronic correlation; and
4)
exchange/repulsion, reflecting the overlapping of electronic distributions [8,10].
Eint is thus written as,
Eint = Ees + Epol + Edisp + Eex-rep
Higher levels of theory yield more elaborate representations of Eint [5,7,9].
The various contributions to Eint, as in eq. (2), can all be formulated in terms of the perturbation operator and the wave functions of the Mi [7,8,9]. However simplified expressions are frequently used [8,10,11,12], for instance in molecular dynamics simulations [8,11]. In the latter, a point-charge approximation is commonly employed for the electrostatic interaction, Ees, and a Lennard-Jones or Buckingham-type potential for Edisp + Eex-rep. Epol is not taken into account, although it could be done, for example by periodically changing the magnitudes of the point charges in the course of the simulation [13].
In the remainder of this paper, we shall focus upon Ees, Epol and Eint. For convenience in notation, we shall treat the case of two components A and B forming a noncovalently-bound complex AB. Extension to more components, on a pair-by-pair basis, is straightforward.

Evaluation of Ees and Epol from electronic densities

Methodology

The energy Ees of the electrostatic interaction between the unperturbed components A and B is given rigorously in terms of their electronic densities ρ A ° ( r ) and ρ B ° ( r )
Ijms 05 00130 i001
In eq. (3), ZM,A and ZN,B are the charges on nuclei M and N of components A and B; RM and RN are their locations.
Since the charge distributions of A and B do not remain unperturbed as the interaction proceeds, but rather have a polarizing effect upon each other, it is necessary to include the associated energy, Epol, in Eint. Several ways of determining Epol have been proposed [7,8,12,14]. We have recently introduced another approach [15], which involves expressing the electronic densities of A and B after interaction as ρ A * ( r ) = ρ A ° ( r ) + Δ ρ A ( r ) and ρ B * ( r ) = ρ B ° ( r ) + Δ ρ B ( r ) ; Δ ρ A ( r ) and Δ ρ B ( r ) are the changes due to polarization. Replacing ρ A ° ( r ) and ρ B ° ( r ) in eq. (3) by ρ A * ( r ) and ρ B * ( r ) produces Ees + Epol; subtracting eq. (3) from both sides leaves,
Ijms 05 00130 i002a Ijms 05 00130 i002b
We represent Δ ρ A ( r ) and Δ ρ B ( r ) by partitioning them into overlapping and nonoverlapping portions [15], which are then treated separately. They are obtained from the electronic density of the complex, ρ AB ( r ) .
Eqs. (3) and (4) give Ees and Epol in terms of integrals over electronic densities. We evaluate these numerically [15], using a technique modeled after that of Gavezzotti [16]. The electronic charge distributions in the integrands are divided, by means of three-dimensional grids [17], into large numbers of uniform volume units, “e-voxels.” Each of these has a negative charge with magnitude equal to its volume times the electonic density at its center. We discard those e-voxels that are outside of the assigned boundary surface of the charge distribution, which is defined to be a specific value of its electronic density, ρmin. To facilitate the computations, blocks of n x n x n e-voxels are “condensed” into “super-e-voxels,” having charges equal to the sums of those of their constituents. The integrals in eqs. (3) and (4) are then evaluated by calculating the electrostatic interactions of the super e-voxels of each molecule with those of the other, or with its nuclei. More detailed discussions of this integration procedure can be found in Gavezzotti [16] and in Ma and Politzer [15].

Applications

We used the methodology that has been outlined to determine Ees for several molecular dimers [15]: (H2O)2, (CH3OH)2, (CH2Cl2)2, (CH3CN)2, (CH3COCH3)2, (CH3SOCH3)2. This was done primarily at several Hartree-Fock levels. One of our objectives was to ascertain the number of e-voxels, the condensation number n, and the ρmin that would be most effective. We found that Ees converges for approximately 106 e-voxels and, for the smaller systems, when ρmin ≤ 105 au (electrons/bohr3). It was also our experience that at least 2000 super e-voxels are needed; thus, for 1.0 x 106 e-voxels, n should be no larger than 7. Our Ees agree well with those obtained by the Morokuma-Kitaura scheme for partitioning interaction energies [18,19].
Epol was computed for the water dimer [15], for which we could compare the results with those from the Morokuma-Kitaura and also the reduced variational space self-consistent-field methods [20]. Hartree-Fock electronic densities were used, corresponding to ten different basis set combinations. The three sets of Epol were in good accord when ρmin was taken to be 0.01 au. It is not surprising that the optimum ρmin is not the same for Epol as for Ees; the extent of overlap between the components, which is determined by ρmin, influences Epol more directly than Ees. Epol was observed to have a relatively low sensitivity to basis set, less than that of Ees [15]. Among the ten (H2O)2 calculations, the largest difference in Epol was 0.41 kcal/mole, between HF/6-31G(d,p)//6-31G(d,p) and HF/cc-pVDZ//6-31G(d,p).
We have also extended these studies to some larger systems, the first of which was the dimer of uracil (1). Since this involves bigger molecules than any of the other dimers for which we computed Ees, we tested whether ρmin ≤ 105 au is still sufficient for Ees to converge [15]. Two dimer structures were investigated, which differ in the relative orientations of the uracil molecules: face-to-face and face-to-back [21]. We found that ρmin ≤ 106 au is now required; this can be seen in Table I. The fact that the magnitude of Ees for the face-to-face dimer is more than double that for the face-to-back was attributed to the former having more intermolecular N-H---O hydrogen bonds [15].
Ijms 05 00130 i003
Finally, we investigated the intermolecular interactions in the crystal lattice of RDX (2, hexahydro-1,3,5-trinitro-s-triazine) [22], which is of considerable interest as an important energetic compound. RDX is frequently the subject of molecular dynamics simulations [11,23], which generally obtain Ees by a point-charge approximation. One of our objectives was to examine how this compares with our Ees from eq. (3). We considered (a) the interaction within an interlocked pair of molecules, and (b) that between two molecules in neighboring interlocked pairs, at the Hartree-Fock and B3PW91 levels, with 6-311+G** basis sets. The electrostatic interaction energies Ees were calculated by two point-charge models, involving (a) Mulliken and (b) CHelpG atomic charges [17]; the latter are derived from
Table 1. Electrostatic interaction energies Ees, in kcal/mole, for two uracil dimers. Uracil molecular geometry taken from MP2/TZ2P(f,d)++ optimized dimer structures.a Number of e-voxels is 1.0 x 106, stepsize is 0.0860 A, and condensation number n = 5.
Table 1. Electrostatic interaction energies Ees, in kcal/mole, for two uracil dimers. Uracil molecular geometry taken from MP2/TZ2P(f,d)++ optimized dimer structures.a Number of e-voxels is 1.0 x 106, stepsize is 0.0860 A, and condensation number n = 5.
DimerComputationalρmin, au
level
1.0 x 10−31.0 x 10−41.0 x 10−51.0 x 10−61.0 x 10−7
Face-to-faceHF/aug-cc-pVDZ −6.69−10.06 −11.79−12.07−12.07
HF/aug-cc-pVTZ−6.56−10.30−12.24−12.42−12.42
HF/aug-cc-pVQZ−6.34−10.08−12.10−12.26−12.26
Face-to-backHF/aug-cc-pVDZ−4.46−4.28−4.95−5.16−5.16
HF/aug-cc-pVTZ−4.19−5.69−5.01−5.11−5.11
HF/aug-cc-pVQZ−3.98−4.02−5.01−5.10−5.10
aRef. 21.
electrostatic potentials. We also determined Ees from the electronic densities by means of eq. (3), using 1.4 x 106 e-voxels and ρmin = 1.0 x 106 au. The Mulliken charges produced very poor Ees, positive for both pairs of interacting molecules. The CHelpG were negative but significantly smaller in magnitude than the Ees from eq. (3) (e.g. −8 vs. −3 kcal/mole for the interlocked pair). We conclude that these point-charge models do not satisfactorily reproduce Ees. Epol was also computed for the two RDX pairs, with eq. (4); it was in the neighborhood of −1 kcal/mole in each case.
Ijms 05 00130 i004

Evaluation of Eint from electronic densities

Methodology

We shall not approach Eint in terms of eq. (2), but rather from the standpoint of the Hellmann-Feynman theorem [24,25,26]. In its general form, this states that, for a system described by,
Ijms 05 00130 i005
it follows that,
Ijms 05 00130 i006
In eq. (6), λ is any parameter appearing in the Hamiltonian H ^ . Thus, the theorem can be expressed in various ways, depending upon the choice of λ. For example, letting λ = Z, the nuclear charge of an N-electron atom with electronic density ρ(r), produces [27,28],
Ijms 05 00130 i007
in which V0 represents the electrostatic potential at the nucleus due to the electrons. This leads to exact relationships between E and V0 [27,29,30], including,
Ijms 05 00130 i008
Ijms 05 00130 i009
Ijms 05 00130 i010
Analogous equations can be derived for molecules [29,30,31].
For our present purpose, we take λ to be RM, the position of nucleus M in the system of interest. Since E/ R M gives the force FM exerted upon M by the electrons and other nuclei, then eq. (6) becomes,
Ijms 05 00130 i011
where ρ(r) is the electronic density of the system. Eq. (11) shows that the force upon any nucleus is given by classical electrostatics.
The binding energy of a nucleus can in principle be determined by using eq. (11) to calculate the work done in bringing it from infinity to its equilibrium position in the force field of the remainder of the system [24,32,33,34]. Extending this approach, we have recently formulated the stabilization energy of a noncovalently-bound complex AB as the work done upon the nuclei and electrons of component A as it is brought from infinity to its position in AB in the force fields of the nuclei and electrons of B [35]. A complicating factor, which was pointed out by Bader in a different context [34], is that the electronic densities of A and B change somewhat as they approach. We neglect this, and use the geometries and electronic densities of A and B as they are in AB. In view of this, the quantity that we obtain is the interaction energy of A and B as they are in the complex, and shall be designated E int * . It is given by [35],
Ijms 05 00130 i012
In deriving eq. (12), it is assumed that the components A and B retain their identities in AB.
Conceptually, E int * differs from Eint, eq. (2), in that the latter is obtained using the ground-state geometries of isolated A and B. Since these often remain essentially the same during the formation of AB, it can be anticipated that E int * , Eint and ΔEstab will frequently be quite similar, if evaluated at comparable levels of accuracy. Indeed, the total energies required to transform the component molecules in (H2O)2 and (HF)2 from their ground states to their forms in the dimers were found to be 0.09 kcal/mole [36] and 0.03 kcal/mole [37], respectively. For the face-to-face dimer of uracil [21], a larger molecule, this energy is 0.79 kcal/mole at the MP2/6-31+G* level [35]. It can be significantly greater, however, for ion-molecule interactions, e.g. F(H2O) [36].
Eq. (12) shows, in the spirit of Feynman [25], that the total interaction energy is classically electrostatic. Eq. (12) is in fact formally identical to eq. (3), differing only in that it involves the electronic densities of A and B as they are in the complex rather than in the free states. Thus the quantities Epol, Edisp and Eex-rep in eq. (2) are simply compensating for Ees not being in terms of the appropriate electronic densities. If it were, then Ees alone would suffice in eq. (2).
A practical concern with regard to eq. (12) is partitioning the total electronic density ρAB(r) into ρA(r) and ρB(r). To do this, we first establish a boundary surface ρmin for the complex; this may differ from those used to obtain Ees and Epol. At each point r within this surface, we determine the ratios of its distance from each nucleus divided by the van der Waals radius of that atom. The point r and the corresponding ρAB(r) are then assigned to the atom (and therefore the component A or B) for which this ratio has the lowest value. We perform the integrations in eq. (12) by means of the numerical technique described in an earlier section of this paper.

Applications

The number of e-voxels that we use in computing E int * depends upon the size of the molecules that are involved, but it continues to be of the order of 106. Thus, it was 1.0 x 106 for (H2O)2 [35], but 3 x 106 for the pair interactions in the crystal lattice of RDX (2) [22]. With regard to ρmin, we found that E int * converges for ρmin ≤ 10−4 au [35].
Our initial calculations of E int * were for four molecular dimers for which reasonable computational/experimental estimates of the stabilization energies ΔEstab are available in the literature, to which our results could be compared. The systems studied included (H2O)2, (HF)2, (H3COH)2 and (HCOOH)2 [35]; the calculations were carried out with the Hartree-Fock, MP2, B3LYP and B3PW91 procedures, and three different correlation-consistent basis sets.
For (H2O)2 and (HF)2, there was overall very good agreement between E int * and ΔEstab. For (H3COH)2 and (HCOOH)2, however, the MP2, B3LYP and B3PW91 E int * underestimated by roughly 2 to 3 kcal/mole the magnitudes of ΔEstab, which are reported as −4.6 to −5.9 kcal/mole for (H3COH)2 and −13.2 kcal/mole for (HCOOH)2 [38]. Several factors may contribute to this (besides the approximations in our procedure), one being a degree of uncertainty in the literature values. It is also likely that our computed dimer structures differ somewhat from the experimental ones upon which these ΔEstab are based. Finally, the calculated electronic densities are of course not exact. The Hartree-Fock E int * were invariably more negative than the others, but the spread was less than 1 kcal/mole, except for (HCOOH)2, for which it was about 3 kcal/mole. For a given computational method, the three basis sets usually gave quite similar results, particularly the two larger ones, cc-pVTZ and cc-pVQZ.
We also determined E int * for the two pairs of molecules in the RDX crystal lattice for which, earlier in this paper, we discussed Ees and Epol. Our predicted E int * for the interlocked pair was about −8 kcal/mole, and −2 to −3 kcal/mole for the interaction between interlocked pairs [22]. These values are very similar to what we obtained for the corresponding Ees, approximately −8 and −3 kcal/mole. A similar situation was found by Bukowski et al [39] in a symmetry-adapted perturbation theory analysis of dimers of dimethylnitramine, (H3C)2N−NO2, a molecule with the same structural elements as RDX. They found Ees and Eint to differ by ≤ 1 kcal/mole for each of the three most stable dimer configurations.

Discussion and summary

The fact that Ees is a good approximation to E int * for two pairs of RDX molecules, i.e., the electrostatic interations between the separate components nearly match the total interaction energies, suggests that (barring fortuitous cancellation) ρ A ° ( r ) and ρ B ° ( r ) , eq. (3), are similar to ρA(r) and ρB(r), eq. (12). This appears to be the case for the dimethylnitramine dimers as well, in view of the findings of Bukowski et al [39] (vide supra). On the other hand, for the two uracil dimers mentioned earlier, our computed Ees differed in each instance by about 3 kcal/mole from the best estimate of ΔEstab [15], being more negative for the face-to-face dimer and more positive for the face-to-back.
The point-charge model was not successful in reproducing Ees for the two pairs of RDX molecules, for the charge definitions investigated. We tested the possibility that the model might be more effective if the charges were obtained for each pair after interaction, perhaps yielding reasonable approximations to E int * , but there was, in general, no improvement.
We believe that our results overall support the formulations in terms of electronic densities that have been given for Ees, Epol and E int * , and the validity of the numerical integration technique that is used to evaluate them. However there is certainly a need for continuing efforts to optimize the assignments of the parameters − i.e., number of e-voxels, ρmin and condensation number n − taking into account the energy quantity being sought and the sizes and shapes of the molecules. The effects of various computational methods (e.g., Hartree-Fock, MP2, density functional) and basis sets should also be further explored.

Acknowledgement

We greatly appreciate the support of the Office of Naval Research, Contract No. N00014-99-1-0393 and Project Officer Dr. Judah M. Goldwasser.

References

  1. Szalewicz, K.; Cole, S. J.; Kolos, W.; Bartlett, R. J. A Theoretical Study of Water Dimer Interaction. J. Chem. Phys. 1988, 89, 3662–3673. [Google Scholar] [CrossRef]
  2. Davidson, E. R.; Chakravorty, S. J. A Possible Definition of Basis Set Superposition Error. Chem. Phys. Lett. 1994, 217, 48–54. [Google Scholar] [CrossRef]
  3. van Duijneveldt, F. B.; van Duijneveldt-van de Rijdt, J. G. C. M.; van Lenthe, J. H. State of Art in Counterpoise Theory. Chem. Rev. 1994, 94, 1873–1885. [Google Scholar] [CrossRef]
  4. Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  5. Chalasinski, G.; Szczesniak, M. M. State of the Art and Challenges of the ab Initio Theory of Intermolecular Interactions. Chem. Rev. 2000, 100, 4227–4252. [Google Scholar] [CrossRef] [PubMed]
  6. Rappe, A. K.; Bernstein, E. R. Ab Initio Calculation of Nonbonded Interactions: Are We There Yet? J. Phys. Chem. A 2000, 104, 6117–6128. [Google Scholar] [CrossRef]
  7. Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of Van der Waals Complexes. Chem. Rev. 1994, 94, 1887–1930. [Google Scholar] [CrossRef]
  8. Engkvist, O.; Astrand, P.-O.; Karlstrom, G. Accurate Intermolecular Potentials Obtained from Molecular Wave Functions: Bridging the Gap between Quantum Chemistry and Molecular Simulations. Chem. Rev. 2000, 100, 4087–4108. [Google Scholar] [CrossRef] [PubMed]
  9. Langlet, J.; Caillet, J.; Berges, J.; Reinhardt, P. Comparison of Two Ways to Decompose Intermolecular Interactions for Hydrogen-Bonded Dimer Systems. J. Chem. Phys. 2003, 118, 6157–6166. [Google Scholar] [CrossRef]
  10. Nobeli, I.; Price, S. L. A Non-Empirical Intermolecular Potential for Oxalic Acid Crystal Structures. J. Phys. Chem. A 1999, 103, 6448–6457. [Google Scholar] [CrossRef]
  11. Politzer, P.; Boyd, S. Molecular Dynamics Simulations of Energetic Solids. Struct. Chem. 2002, 13, 105–113. [Google Scholar] [CrossRef]
  12. Gavezzotti, A. Calculation of Intermolecular Interaction Energies by Direct Numerical Integration over Electron Densities. 2. An Improved Polarization Model and the Evaluation of Dispersion and Repulsion Energies. J. Phys. Chem. B 2003, 107, 2344–2353. [Google Scholar] [CrossRef]
  13. Seminario, J. M.; Concha, M. C.; Politzer, P. A Density-Functional/Molecular Dynamics Study of the Structure of Liquid Nitromethane. J. Chem. Phys. 1995, 102, 8281–8282. [Google Scholar] [CrossRef]
  14. Chipot, C.; Dehez, F.; Angyan, J.; Millot, C.; Orozco, M.; Luque, F. J. Alternative Approaches for the Calculation of Induction Energies: Characterization, Effectiveness, and Pitfalls. J. Phys. Chem. A 2001, 105, 11505–11514. [Google Scholar] [CrossRef]
  15. Ma, Y.; Politzer, P. Calculation of Electrostatic and Polarization Energies from Electron Densities. J. Chem. Phys. 2004, 120, 3152–3157. [Google Scholar] [CrossRef]
  16. Gavezzotti, A. Calculation of Intermolecular Interaction Energies by Direct Numerical Integration over Electron Densities. I. Electrostatic and Polarization Energies in Molecular Crystals. J. Phys. Chem. B 2002, 106, 4145–4154. [Google Scholar] [CrossRef]
  17. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrezewski, V. G.; Montgomery, J. A.; Stratmann, R. E.; Burant, J. C.; Dappich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G.; Aayala, P. Y.; Cui, Q.; Morokuma, K.; Malick, D. K.; Rubuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian 98, Revision A.7; Gaussian, Inc.: Pittsburgh, PA, 1998. [Google Scholar]
  18. Morokuma, K. Molecular Orbital Studies of Hydrogen Bonds. III. C=O---H4O Hydrogen Bond in H2CO---H2O and H2CO---2H2O. J. Chem. Phys. 1971, 55, 1236–1244. [Google Scholar] [CrossRef]
  19. Morokuma, K.; Kitaura, K. Chemical Applications of Atomic and Molecular Electrostatic Potentials; Politzer, P., Truhlar, D. G., Eds.; Plenum: New York, 1981; Chapter 10; pp. 215–242. [Google Scholar]
  20. Stevens, W. J.; Fink, W. H. Frozen Fragment Reduced Variational Space Analysis of Hydrogen Bonding Interactions. Application to Water Dimer. Chem. Phys. Lett. 1987, 139, 15–22. [Google Scholar] [CrossRef]
  21. Leininger, M. L.; Nielsen, I. M. B.; Colvin, M. E.; Janssen, C. L. Accurate Structures and Binding Energies for Stacked Uracil Dimers. J. Phys. Chem. A 2002, 106, 3850–3854. [Google Scholar] [CrossRef]
  22. Politzer, P.; Ma, Y. Noncovalent Intermolecular Energetics:  RDX Crystal. submitted to. Int. J. Quantum Chem.
  23. Sorescu, D. C.; Rice, B. M.; Thompson, D. L. Energetic Materials. Part I. Decomposition, Crystal and Molecular Properties; Politzer, P., Murray, J. S., Eds.; Elsevier: Amsterdam, 2003; pp. 125–184. [Google Scholar]
  24. Hellmann, H. Einfuhrung in die Quantenchemie; Franz Deuticke: Leipzig, 1937. [Google Scholar]
  25. Feynman, R. P. Forces in Molecules. Phys. Rev. 1939, 56, 340–343. [Google Scholar] [CrossRef]
  26. Berlin, T. Binding Regions in Diatomic Molecules. J. Chem. Phys. 1951, 19, 208–213. [Google Scholar] [CrossRef]
  27. Foldy, L. L. A Note on Atomic Binding Energies. Phys. Rev. 1951, 83, 397–399. [Google Scholar] [CrossRef]
  28. Gaspar, R. Many-Electron Problems. I. Energy Relations in the Theory of Neutral Atoms. Int. J. Quantum Chem. 1967, I, 139–145. [Google Scholar] [CrossRef]
  29. Politzer, P.; Parr, R. G. Some New Energy Formulas for Atoms and Molecules. J. Chem. Phys. 1974, 61, 4258–4262. [Google Scholar] [CrossRef]
  30. Politzer, P. Fundamental World of Quantum Chemistry. A Tribute to the Memory of Per-Olov Lowdin; Vol. I, Brandas, E. J., Kryachko, E. S., Eds.; Kluwer: Dordrecht, 2003; pp. 631–638. [Google Scholar]
  31. Wilson, E. B., Jr. Four-Dimensional Electron Density. J. Chem. Phys. 1962, 36, 2232–2233. [Google Scholar] [CrossRef]
  32. Hurley, A. C. The Electrostatic Calculation of Molecular Energies. I - III. Proc. Roy. Soc. (London) 1954, A226, 170-178, 179-192, 193-205. [Google Scholar] [CrossRef]
  33. Bader, R. F. W. The Use of the Hellmann-Feynman Theorem to Calculate Molecular Energies. Can. J. Chem. 1960, 38, 2117–2127. [Google Scholar] [CrossRef]
  34. Bader, R. F. W. The Force Concept in Chemistry; Deb, B. M., Ed.; Van Nostrand Reinhold: New York, 1981; Chapter 2; pp. 39–136. [Google Scholar]
  35. Ma, Y.; Politzer, P. Determination of Noncovalent Interaction Energies from Electronic Densities. J. Chem. Phys. 2004, 120, 8955–8959. [Google Scholar]
  36. Xantheas, S. S. On the Importance of the Fragment Relaxation Energy Terms in the Estimation of the Basis Set Superposition Error Correction to the Intermolecular Interaction Energy. J. Chem. Phys. 1996, 104, 8821–8824. [Google Scholar] [CrossRef]
  37. Peterson, K. A.; Dunning, T. H., Jr. Benchmark Calculations with Correlated Molecular Wave Functions. VII. Binding Energy and Structure of the HF Dimer. J. Chem. Phys. 1995, 102, 2032–2041. [Google Scholar] [CrossRef]
  38. Tsuzuki, S.; Uchimaru, T.; Matsumura, K.; Mikami, M.; Tanabe, K. Effects of Basis Set and Electron Correlation on the Calculated Interaction Energies of Hydrogen Bonding Complexes: MP2/cc-pV5Z Calculations of H2O--MeOH, H2O--Me2O, H2O--H2CO, MeOH--MeOH, and HCOOH--HCOOH Complexes. J. Chem. Phys. 1999, 110, 11906–11910. [Google Scholar] [CrossRef]
  39. Bukowski, R.; Szalewicz, K.; Chabalowski, C. F. Ab Initio Interaction Potentials for Simulations of Dimethylnitramine Solutions in Supercritical Carbon Dioxide with Cosolvents. J. Phys. Chem. A 1999, 103, 7322–7340. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Ma, Y.; Politzer, P. Electronic Density Approaches to the Energetics of Noncovalent Interactions. Int. J. Mol. Sci. 2004, 5, 130-140. https://doi.org/10.3390/i5040130

AMA Style

Ma Y, Politzer P. Electronic Density Approaches to the Energetics of Noncovalent Interactions. International Journal of Molecular Sciences. 2004; 5(4):130-140. https://doi.org/10.3390/i5040130

Chicago/Turabian Style

Ma, Yuguang, and Peter Politzer. 2004. "Electronic Density Approaches to the Energetics of Noncovalent Interactions" International Journal of Molecular Sciences 5, no. 4: 130-140. https://doi.org/10.3390/i5040130

Article Metrics

Back to TopTop