Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 13 June 2022
Sec. Theoretical and Computational Chemistry
Volume 10 - 2022 | https://doi.org/10.3389/fchem.2022.906674

Molecular Interactions From the Density Functional Theory for Chemical Reactivity: The Interaction Energy Between Two-Reagents

www.frontiersin.orgRamón Alain Miranda-Quintana1* www.frontiersin.orgFarnaz Heidar-Zadeh2 www.frontiersin.orgStijn Fias3 www.frontiersin.orgAllison E. A. Chapman3 www.frontiersin.orgShubin Liu4 www.frontiersin.orgChristophe Morell5 www.frontiersin.orgTatiana Gómez6* www.frontiersin.orgCarlos Cárdenas7,8* www.frontiersin.orgPaul W. Ayers3*
  • 1Department of Chemistry and Quantum Theory Project, University of Florida, Gainesville, FL, United States
  • 2Department of Chemistry, Queen’s University, Kingston, ON, Canada
  • 3Department of Chemistry and Chemical Biology, McMaster University, Hamilton, ON, Canada
  • 4Research Computing Center, University of North Carolina, Chapel Hill, NC, United States
  • 5Université de Lyon, Université Claude Bernard Lyon 1, Institut des Sciences Analytiques-UMR CNRS 5280, Villeurbanne, France
  • 6Theoretical and Computational Chemistry Center, Institute of Applied Chemical Sciences, Faculty of Engineering, Universidad Autonoma de Chile, Santiago, Chile
  • 7Departamento de Fisica, Facultad de Ciencias, Universidad de Chile, Santiago, Chile
  • 8Centro para el desarrollo de la Nanociencias y Nanotecnologia, CEDENNA, Santiago, Chile

Reactivity descriptors indicate where a reagent is most reactive and how it is most likely to react. However, a reaction will only occur when the reagent encounters a suitable reaction partner. Determining whether a pair of reagents is well-matched requires developing reactivity rules that depend on both reagents. This can be achieved using the expression for the minimum-interaction-energy obtained from the density functional reactivity theory. Different terms in this expression will be dominant in different circumstances; depending on which terms control the reactivity, different reactivity indicators will be preferred.

GRAPHICAL ABSTRACT
www.frontiersin.org

GRAPHICAL ABSTRACT

1 Introduction

When a reagent approaches a reactant molecule, the reactant is perturbed. If the perturbing reagent stabilizes the molecule (lowering its energy) or destabilizes the molecule comparatively little (raising its energy by a relatively small amount), then a reaction may occur. For example, regioselectivity can be predicted by finding the location where the reagent imparts the greatest additional stability to the reaction substrate. By comparing results for a series of reagents and a given molecule, or for a series of molecules and a single reagent, absolute chemical reactivity trends can be understood, and even predicted.

This basic viewpoint on chemical reactivity is the fundamental basis of the density functional theory for chemical reactivity (DFT-CR), often called chemical density functional theory or conceptual DFT (Parr and Yang, 1989; Chermette, 1999; Geerlings et al., 2003; Ayers et al., 2005a; Chattaraj et al., 2006; Gazquez, 2008; Liu, 2009a; Johnson et al., 2012b; De Proft et al., 2014; Fuentealba and Cardenas, 2015; Miranda-Quintana, 2018). The fundamental mathematics of DFT-CR is functional perturbation theory (Ayers et al., 2005a) using whichever representation of the chemical system (or “ensemble”) is deemed to be most convenient (Nalewajski and Parr, 1982; DeProft et al., 1997; Liu and Parr, 1997; Perez et al., 2008; Cardenas et al., 2009a). In this paper we will only consider the most common, “closed-system picture” (canonical ensemble). However, the same approach could be used for any of its Legendre (Johnson et al., 2012a) transforms (Ayers and Fuentealba, 2009).

In the closed-system picture, the state of a molecule is specified by its number of electrons, N, and the external potential that binds those electrons, v(r). When a reagent approaches, the number of electrons changes (from electron transfer between reagents) and the external potential changes (because the electrons in the molecule are attracted to the nuclei and repelled by the electrons of the approaching reagent). The change in the energy of a molecule with non-degenerate ground state (Cardenas et al., 2011a; Bultinck et al., 2013, 2014; Bultinck et al., 2015) is then,

ΔU[Δv(r);ΔN]=ΔVnn+(μΔN+12η(ΔN)2+)+(ΔNf(r)Δv(r)dr+12(ΔN)2Δf(r)Δv(r)dr+)+(ρ(r)Δv(r)dr+12χ(r,r)Δv(r)Δv(r)drdr+)(1)

Here ΔVnn is the change in nuclear-nuclear repulsion energy, ΔN is the number of electrons transferred from the reagent to the molecule under scrutiny and Δv(r) is the change in the molecular external potential due to the presence of the reagent. These quantities depend on the particular identity of the reagent. The coefficients in this expansion—the chemical potential μ, (Parr et al., 1978) the chemical hardness η, (Parr and Pearson, 1983; Pearson, 1997; Ayers, 2007b) the Fukui function f(r), (Parr and Yang, 1984; Yang et al., 1984; Ayers and Levy, 2000; Ayers P. W. et al., 2009) the dual descriptor Δf(r), (Fuentealba and Parr, 1991; Morell et al., 2005, 2006; Ayers et al., 2007; Cardenas et al., 2009b; Geerlings et al., 2012) the electron density ρ(r), and the linear-response (or polarizability) kernel χ(r,r’) (Berkowitz and Parr, 1988; Senet, 1996; Ayers, 2001; Sablon et al., 2010; Geerlings et al., 2014; Franco-Perez et al., 2015a)—are often taken to depend only on which molecule we are studying, and not on what reagent is being considered. The coefficients in the expansion determine how much (or how little) the reactant of interest is stabilized by a given reagent. Since energetically favorable interactions are associated with high reactivity, the coefficients in this expansion are reactivity indicators for the molecule being studied (Ayers et al., 2005a).

The assumption that the fine details of the reagent do not control the qualitative aspects of the reactivity is implicitly invoked when the coefficients of the Taylor expansion are used as reactivity indicators. This is often an adequate assumption. The mere idea that we can define scales of acidity and basicity (proclaiming, for example, that hydrochloric acid is a stronger acid than acetic acid) presumes that the properties of reactants (in this case, the acidity) do not depend very strongly on the identity of the attacking reagents (the base) or the solvent in which the reaction occurs. Similarly, when we say that nitrobenzene is a deactivating meta-director for electrophilic aromatic substitution we imply that the regioselectivity properties of nitrobenzene are largely independent of the choice of electrophile. Extending these ideas: most molecules are reactive at just one (and certainly no more than a few) molecular site(s); these reactivity preferences recur across a wide range of reagents, solvents, and experimental conditions. In cases like this, just a few coarse details of the reagent (e.g., the reagent’s electronegativity and hardness, which determine whether ΔN is small or large) and the reagent’s charge will suffice to describe the full range of possible reactions. In these cases, applying DFT-CR is as simple, and as complicated, as finding an appropriate coefficient (or linear combination of coefficients) from Eq. 1. The “perturbative perspective” on chemical reactivity says by choosing simple “model reagents” (often just a point charge with some capacity for accepting/donating electrons), one can decipher the reactivity preferences of the molecule of interest. There are many examples where this approach has been successfully employed to define new reactivity indicators (Parr et al., 1999; Ayers and Parr, 2000; Ayers and Parr, 2001; Ayers et al., 2005b; Anderson et al., 2007a; Anderson et al., 2007b; Cedillo et al., 2007; Gazquez et al., 2007; Cardenas et al., 2011b; Cardenas, 2011; Alain Miranda-Quintana R. and Ayers P. W., 2016), (Bochicchio, 2015; Alain Miranda-Quintana et al., 2016; Miranda-Quintana R. A., 2016, Miranda-Quintana, R. A. 2017; Gonzalez et al., 2018). This simple perspective has been successfully applied to the study of redox reactions, the study of solvation processes, and to the reactivity of aromatic species (Miranda-Quintana and Smiatek, 2020; Miranda-Quintana et al., 2022).

Chemistry is not always this simple. Certainly some reactions happen (or at least happen more readily than expected) because the reactant molecules are well-matched. Classic cases include the hard/soft acid/base principle and the Woodward-Hoffmann rules (e.g., ethene and butadiene are chemically similar, yet ethene is susceptible to cycloaddition with butadiene but not with itself). In the case of the hard/soft acid/base principle, treating both reagents as simplified “model reagents” suffices (Ayers, 2005; Ayers et al., 2006; Anderson et al., 2007a; Ayers, 2007b; Cardenas and Ayers, 2013; Miranda-Quintana, 2017a). The Woodward-Hoffmann rules can be treated using the DFT-CR framework, (De Proft et al., 2006; Ayers et al., 2007; De Proft et al., 2008; Sablon et al., 2009; Geerlings et al., 2012) but the fundamental reason that the dual descriptor and the initial hardness response suffice is still somewhat mysterious (Cardenas et al., 2009b). A more complete model for the interaction energy is clearly needed in this case. In the next section we will define just such a model. Section 3 discusses the key results. In follow-up papers we will derive an expression for the initial hardness response and reveal a new link between this quantity and the dual descriptor.

2 An Interaction Energy Model From Density Functional Theory Perturbation Theory

Our analysis is based on Eq. 1, including only the terms that are shown explicitly. Higher-order derivatives with respect to the number of electrons are usually small. (Fuentealba and Parr, 1991; Geerlings and De Proft, 2008; Morell et al., 2013; Heidar-Zadeh et al., 2016b; Munoz and Cardenas, 2017), and although they have attracted some recent interest, we will not take them into account here (Hoffmann et al., 2020; Miranda-Quintana R. A. et al., 2021; Miranda-Quintana R. A. et al., 2021). One may argue that since there are only one-body and two-body terms in the Hamiltonian, the energy and its functional derivatives with respect to v(r) should be nearly quadratic in the number of electrons (Ayers and Parr, 2008). Higher order derivatives with respect to the external potential correspond to hyperpolarizabilities; (Fuentealba and Parr, 1991; Senet, 1996; Cardenas et al., 2009a) these effects are difficult to compute accurately and are hopefully negligible. We assume that all the derivatives exist. Treating them exactly requires working with the system at finite temperature: (Franco-Perez et al., 2015a; Franco-Perez et al., 2015b; Malek and Balawender, 2015; Franco-Perez et al., 2016; Franco-Perez et al., 2017a; b;Franco-Perez et al., 2017c; Franco-Perez et al., 2017d; Franco-Perez et al., 2017e; Miranda-Quintana, 2017c; Miranda-Quintana RA. et al., 2017; Miranda-Quintana R. A. et al., 2017; Polanco-Ramirez et al., 2017; Franco-Pérez et al., 2018; Robles et al., 2018; Gázquez et al., 2019): since there are derivative discontinuities at the zero-temperature limit (Perdew et al., 1982; Yang et al., 2000; Ayers, 2008; Yang et al., 2016). One may argue that, for interacting molecular fragments, the “effective temperature” is nonzero; cf. refs. (Ayers, 2007a; Alain Miranda-Quintana R. and Ayers P. W., 2016).

The perturbation expansion, (1), gives us information about the energy of A in the presence of B (UA+ΔUA[ΔvA(r);ΔNA]) and the energy of B in the presence of A (UB+ΔUB[ΔvB(r);ΔNA]). We need to add these two interaction-energy expression and make corrections for 1) the double-counting of interactions and 2) nonadditive energy contributions like exchange and correlation between A and B.

The electron density of the interacting reagents, ρAB(r), is divided into electron densities for the subsystems,

ρAB(r)=ρA(r)+ρB(r).(2)

It is convenient to also define the nuclear charge density of the reactants, i.e.,

zA0(r)=αAZαδ(rRα)zB0(r)=βBZβδ(rRβ)(3)

The total charge density of reactant A is then zA0(r)ρA(r). There is a similar expression for reactant B. The external potentials of the reactants are given by expressions like

vA0(r)=zA0(r)|rr|dr(4)

The interaction energy can then be written in terms of the fundamental density functional for the energy, (Hohenberg and Kohn, 1964)

ΔUAB=ΔVnn+(EvA0+vB0[ρAB]EvA0[ρA0]EvB0[ρB0])(5)

Here ρA0(r) and ρB0(r) denote the electron densities of the separated reactants and

ΔVnn=zA0(r)zB0(r)|rr|drdr(6)

denotes the change in nuclear-nuclear repulsion energy when the reactants are brought together. The contribution from rearrangements of the nuclei within the fragments is neglected for simplicity; this assumes that in the initial approach between the reagents, the internal molecular geometries of the individual reagents are preserved. This approximation was suggested, for example, in ref. (Ayers and Parr, 2001). This approximation can be relaxed if a term corresponding to the “preparation” energy of the reagents is added (Morokuma, 1971; Kitaura and Morokuma, 1976; Ziegler and Rauk, 1977, 1979; Wu et al., 2009; Azar and Head-Gordon, 2012).

The electronic energy functional is then decomposed,

ΔUAB=ΔVnn+(Ts[ρAB]+Exc[ρAB]+ρAB(r)(vA0(r)+vB0(r))dr+12ρAB(r)ρAB(r)|rr|drdrEvA0[ρA0]EvB0[ρB0]).(7)

Here

F[ρ]=Ts[ρ]+J[ρ]+Exc[ρ](8)

is the Hohenberg-Kohn functional, (Hohenberg and Kohn, 1964) Ts[ρ] is the Kohn-Sham kinetic energy functional, (Kohn and Sham, 1965) and Exc[ρ] is the exchange-correlation energy functional.

Ev[ρ]=F[ρ]+ρ(r)v(r)dr(9)

is the variational energy functional in DFT (Hohenberg and Kohn, 1964; Kohn and Sham, 1965)

Equation 7 can be simplified by breaking each contribution to the energy into its subsystem contributions:

ΔUAB=ΔVnn+(Ts[ρA]+Ts[ρB]+{Ts[ρAB]Ts[ρA]Ts[ρB]}+Exc[ρA]+Exc[ρB]+{Exc[ρAB]Exc[ρA]Exc[ρB]}ρA(r)vA0(r)dr+ρB(r)vB0(r)dr+(ρA(r)vB0(r)+ρB(r)vA0(r))dr+12ρA(r)ρA(r)|rr|drdr+12ρB(r)ρB(r)|rr|drdr+ρA(r)ρB(r)|rr|drdrEvA0[ρA0]EvB0[ρB0])(10)

Using the definition of the electronic energy functional and Eq. 4, this simplifies to

ΔUAB=(Ts[ρAB]Ts[ρA]Ts[ρB]+Exc[ρAB]Exc[ρA]Exc[ρB]+(zA0(r)ρA(r))(zB0(r)ρB(r))|rr|drdrEvA0[ρA]EvA0[ρA0]+EvB0[ρB]EvB0[ρB0])(11)

The terms in the first two lines are identified as the non-additive kinetic energy and the non-additive exchange-correlation energy,

Tsnon-add[ρA,ρB]Ts[ρAB]Ts[ρA]Ts[ρB](12)
Excnon-add[ρA,ρB]Exc[ρAB]Exc[ρA]Exc[ρB](13)

The next term in Eq. 11 is the “non-additive Coulomb energy” this models the electrostatic interactions between the systems. The final terms in Eq. 11 include the polarization and electron-transfer energies. When the reactants approach each other, the electron densities of the reactants are polarized and the number of electrons in the reactants changes; the terms on the last line correct for these effects. In this approach, dispersion is treated as a facet of polarization.

The electrostatic interactions are readily evaluated in any quantum chemistry program, so the only “difficult” terms are the electron-transfer terms on the last line of Eq. 11. These could be approximated using a quadratic expansion in the electron density, i.e., (Liu and Parr, 1997; Ayers and Parr, 2000)

EvA0[ρA]EvA0[ρA0](ρA(r)ρA0(r))μA0dr+12(ρA(r)ρA0(r))η[ρA0;r,r](ρA(r)ρA0(r))drdr=(ΔNA)μA0+12ΔρA(r)ΔρB(r)η[ρA0;r,r]drdr(14)

The second-order derivative is the hardness kernel, (Berkowitz and Parr, 1988)

η[ρ,r,r]=δF[ρ]δρ(r)δρ(r)(15)

The chemical potential and the hardness kernel of the isolated reagent could be evaluated exactly from the output of a Kohn-Sham program or using one of the many approximate forms that have been proposed in the literature (Liu et al., 1997; De Proft et al., 1998; Ayers, 2001; Chattaraj and Maiti, 2001; Torrent-Sucarrat et al., 2005; Torrent-Sucarrat et al., 2007; Gómez et al., 2021). Inserting (14) into (11) gives

ΔUAB(Tsnon-add[ρA0+ΔρA,ρB0+ΔρB]+Excnon-add[ρA0+ΔρA,ρB0+ΔρB]+ΔVnn+(ΔρA(r)ΦB(r)+ΔρB(r)ΦA(r))drΔρA(r)ΔρB(r)|rr|drdr+(μA0μB0)ΔNA+12[ΔρA(r)ΔρA(r)η[ρA0;r,r]+ΔρB(r)ΔρB(r)η[ρB0;r,r]]drdr)(16)

In order to obtain a simple and intuitive form in Eq. 16, we have defined the electrostatic potential of the reactants in the usual way, e.g.,

ΦA(r)=zA0(r)ρA0(r)ΔρA(r)|rr|dr.(17)

with

ΔρA(r)=ρA(r)ρA0(r)(18)

The actual interaction energy is then approximated by minimizing the expression in Eq. 16 with respect to all changes in density that preserve the total number of electrons and which do not cause the electron density of either fragment to become negative. i.e.,

ΔUABminΔρA(r);ΔρB(r)ΔUAB[ΔρA,ΔρB](19)

subject to

ρA0(r)+ΔρA(r)0ρB0(r)+ΔρB(r)0(20)
ΔNAΔρA(r)dr=ΔρB(r)dr(21)

Equation 19 extends the variational principle for the Fukui function to include effects other than electron transfer (Chattaraj et al., 1995; Ayers and Parr, 2000).

For computational purposes, the preceding approach is probably the most useful. For conceptual purposes, however, an alternative perspective that makes references to the most commonly used reactivity indicators may be preferred. Notice that many of the commonly employed reactivity indicators appear in the expression for the change in electron density. e.g.,

ΔρA(r)=(ΔNAfA(r)+(ΔNA)22ΔfA(r)+)+(χA(r,r)ΔvA(r)dr+ΔNA(χA(r,r)NA)v(r)ΔvA(r)dr+)+ΔNAfA(r)+(ΔNA)22ΔfA(r)+ΔρA(pol)[ΔvA(r);r](22)

Note that we have defined the change in the electron density of A due to polarization as ΔρA(pol)[ΔvA(r);r].

Equation 22 is directly useful only if we know ΔvA(r). However, we do not know how the external potential of A changes due to the presence of the other reagent. (There is an exact formulation to compute ΔvA(r), but it is computationally demanding (Ayers, 2000; Ayers and Parr, 2001; Kiewisch et al., 2008; Roncero et al., 2008; Fux et al., 2010)) Among the various approximations that have been proposed, the simplest is to identify ΔvA(r) with the electrostatic potential of molecule B (Ayers and Parr, 2001). (This is certainly acceptable when the reactants are reasonably far apart, so that the electrons in reactant A and reactant B can be considered distinguishable particles.) This simple approximate form may also be rationalized from Eq. 16, where the electrostatic potential of B serves as an “effective external potential” for the electrons in A.

Using

ΔvA(r)=ΦB(r)ΔvB(r)=ΦA(r)(23)

and Eq. 22, we can expand the electrostatic terms in Eq. 11, giving,

(zA0(r)ρA(r))(zB0(r)ρB(r))|rr|drdr(zA0(r)ρA0(r))(zB0(r)ρB0(r))|rr|drdrΔNA(fA(r)ΦB0(r)+fB(r)ΦA0(r))dr(ΔNA)22(ΔfA(r)ΦB0(r)+ΔfB(r)ΦA0(r))dr(ΔρA(pol)[ΦB;r]ΦB0(r)+ΔρB(pol)[ΦA;r]ΦA0(r))+ΔρA(pol)[ΦB;r]ΔρB(pol)[ΦA;r]|rr|drdrΔNAΔρA(pol)[ΦB;r]fB(r)ΔρB(pol)[ΦA;r]fA(r)|rr|drdr+(ΔNA)22ΔρA(pol)[ΦB;r]ΔfB(r)+ΔρB(pol)[ΦA;r]ΔfA(r)|rr|drdr(ΔNA)2fA(r)fB(r)|rr|drdr+(ΔNA)3fA(r)ΔfB(r)fB(r)ΔfA(r)|rr|drdr+(ΔNA)44ΔfA(r)ΔfB(r)|rr|drdr(24)

In addition to the contributions to the interaction energy from electrostatic interactions of the reactants and the electron transfer between them, there are contributions due to mutual polarization of the reactants. The leading order term of this type is

JΔρ(pol)=(ΔρA(pol)[ΦB;ΔNA,r]ΦB0(r)+ΔρB(pol)[ΦA;ΔNA,r]ΦA0(r))(25)

This term is negative: the electrostatic potential of reactant B polarizes the electron density of A in a way that increases the attraction between the reagents.

Some readers may question whether JΔρ(pol) should be multiplied by a factor of 1/2. After all, the change in energy due to polarization is

ΔEpol12Δv(r)Δv(r)χ(r,r)drdr12Δρ(pol)[Δv;r]Δv(r)dr.(26)

When the two reactants approach each other, they polarize each other. This polarization increases the attraction of the reactants for each other. However, polarization also mixes excited-state wavefunctions with the ground-state wavefunction; this increases the electronic energy of the reactants(Ayers, 2007a). The total change in energy, cf. Eq. 11, arises because the electronic energy of the reactants, EvA0[ρA] and EvB0[ρB], increases by ΔEpol while the interaction energy between the reactants decreases by twice that amount. This explains why there is no factor of 1/2 in Eq. 24 and draws our attention to the final contributions to the fundamental expression for the interaction energy, Eq. 11. When the electron density of the reactants changes (due to polarization or electron transfer), the energy of the reactants also changes. One has

EvA0[ρA]EvA0[ρA0]=(ΔNAμA+(ΔNA)22ηA+)+12ΔρA[ΦB;ΔNA,r]ΦB(r)dr+(27)

The terms in the first line represent the change in system energy due to electron transfer; the terms in the second line shows how the energy of the system increases as the density is polarized away from the ground-state density.

Combining Eqs 11, 22, 24, 27 gives an alternative method for computing the interaction energy between reagents.

A. Determine the non-additive kinetic energy and the non-additive exchange-correlation energy. Sometimes simple Thomas-Fermi-Dirac-like functionals will suffice here. One has

Tsnon-add[ρA0,ρB0]Ts[ρA0+ρB0]Ts[ρA0]Ts[ρB0](28)
Excnon-add[ρA0,ρB0]Exc[ρA0+ρB0]Exc[ρA0]Exc[ρB0].(29)

B. Compute the electrostatic potential of the isolated reactants. E.g.,

ΦA0(r)=zA0(r)ρA0(r)|rr|dr.(30)

Use the electrostatic potential to estimate how the density of each reactant is polarized by the electrostatic potential of the other reactant,

ΔρA(pol)[ΦB0;r]=(δρA(r)δv(r))NA0ΦB0(r)dr.(31)

Also compute the polarization energy,

EA(pol)[ΦB0]=12ΔρA(pol)[ΦB0;r]ΦB0(r)dr.(32)

C. Determine the extent of electron transfer, ΔNA, by minimizing the expression for the interaction energy with respect to ΔNA.

ΔUAB[ΔNA]=(Tsnon-add[ρA0,ρB0]+Excnon-add[ρA0,ρB0])+((zA0(r)ρA0(r))(zB0(r)ρB0(r))|rr|drdr+2(EA(pol)[ΦB0]+EB(pol)[ΦA0])+ΔρA(pol)[ΦB0,r]ΔρB(pol)[ΦA0,r]|rr|drdrΔNA(fA(r)ΦB0(r)fB(r)ΦA0(r))dr+ΔNA(ΔρB(pol)[ΦA0,r]fA(r)ΔρA(pol)[ΦB0,r]fB(r))|rr|drdr(ΔNA)22(ΔfA(r)ΦB0(r)+ΔfB(r)ΦA0(r))dr+(ΔNA)22(ΔρA(pol)[ΦB0,r]ΔfB(r)+ΔρB(pol)[ΦA0,r]ΔfA(r))|rr|drdr(ΔNA)2fA(r)fB(r)|rr|drdr+(ΔNA)3fA(r)ΔfB(r)fB(r)ΔfA(r)|rr|drdr+(ΔNA)44ΔfA(r)ΔfB(r)|rr|drdr)+(ΔNA(μA0μB0)+(ΔNA)22(ηA0+ηB0)EA(pol)[ΦB0]EB(pol)[ΦA0])(33)

D. The electron density of the reagents is now changed due to electron transfer and mutual polarization. Use the updated electron densities,

ρA1(r)=ρA0(r)+ΔNA0fA0(r)+(ΔNA0)22ΔfA0(r)+ΔρA(pol)[ΦB0;r](34)

to compute an updated electrostatic potential ΦA1(r). This can then be used to update the density change due to polarization and the polarization energy (using Eqs 31, 32; step B). The extent of electron transfer is now computed by substituting the updated quantities into (Eq. 33) and minimizing the energy with respect to ΔNA (step C). The electron density of the reactants can then be updated

ρA2(r)=ρA1(r)+ΔNA1fA0(r)+(ΔNA1)22ΔfA0(r)+ΔρA(pol)[ΦB1;r](35)

and steps B and C can be iterated until value of the interaction energy has converged to the desired level of precision. (Notice how the supra-indices in Eqs 34, 35 distinguish between the magnitudes that are updated, from those that are kept constant.)

Step A could also be included in the iteration cycle, but if that is done then one needs to use the linear response kernel for the perturbed molecular fragment when evaluating Eq. 31; this may not be practical.

In this form, we are always expanding the Taylor series given in Eq. 33 around a fixed point that corresponds to the isolated reagents. We can improve this procedure by centering the expansion at a point that already contains the initial interaction between the molecules or, even more generally, the influence of the molecular environement (e.g., solvent) (Alain Miranda-Quintana R. and Ayers P. W., 2016, Alain Miranda-Quintana R. and Ayers, P. W. 2016; Alain Miranda-Quintana et al., 2016). For example, we can correct the energies using:

EM=EM0+ρM0(r)Δv(r)dr(36)

where the sub-index M indicates the number of electrons, and the supra-index 0 identifies the isolated-system properties. We can also update the density by considering the effect of the perturbation on the Kohn-Sham orbitals:

|φk=pkφp0|Δv(r)|φk0εkεp|φp0(37)

In these cases we take the perturbation, Δv(r), according to Eqs 23, 30. The modified energies and densities can then be used to estimate the reactivity descriptors present in Eq. 33.

An even more realistic solution will be to change the point around which we will expand the Taylor expansion after each iteration step. To do so, we will have to recalculate the perturbed energies and orbitals (e.g., densities) after each iteration, using the electrostatic potentials associated to Eq. 35. But, additionally, we must correct the expressions used to calculate the chemical potentials and Fukui functions to accomodate fractionally-charged systems:

μM+ΔN=μM+ηMΔN(38)
f(r)M+ΔN=f(r)M+Δf(r)MΔN(39)

Extending the preceding mathematical framework to include reagents with degenerate ground states is somewhat difficult, but is not impossible (Cardenas et al., 2011a; Bultinck et al., 2013, 2014; Bultinck et al., 2015). Extending this procedure to include open-shell reagents and spin-specific reaction pathways is quite straightforward: (Galvan et al., 1988; Galvan and Vargas, 1992; Ghanty and Ghosh, 1994; Ghosh, 1994; Ghanty and Ghosh, 1996; Vargas and Galvan, 1996; Vargas et al., 1998; Melin et al., 2003; Miranda-Quintana R. A. and Ayers P. W., 2016): none of the formulas change if the matrix-vector notation for spin-resolved reactivity indicator functions is invoked (Garza et al., 2006; Perez et al., 2008).

3 Discussion

3.1 Connections to Other Theoretical Methods

This method of computing the interaction energy from its fragment decomposition is fundamentally related to the method of density-functional embedding (Cortona, 1991; Vaidehi et al., 1992; Wesolowski and Warshel, 1993; Govind et al., 1999; Wesolowski, 2004; Wesolowski and Leszczynski, 2006; Elliott et al., 2010; Fux et al., 2010; Goodpaster et al., 2010; Goodpaster et al., 2011; Laricchia et al., 2011). That method, unlike this one, performs iterative ab initio calculations on the fragments rather than using the perturbation expansion.

This method is also fundamentally related to electronegativity equalization approaches to molecular mechanics (Mortier et al., 1985; Mortier et al., 1986; Yang and Mortier, 1986; Mortier, 1987; Rappe and Goddard, 1991; Bultinck et al., 2002a; Bultinck et al., 2002b; Verstraelen et al., 2009; Verstraelen et al., 2011). In those methods one replaces the non-additive kinetic and exchange-correlation energies with an empirically parameterized “repulsive wall.” One accelerates the method by approximating the Coulomb-like integrals by using atom-centered charges and (sometimes) dipoles, with the charges and dipoles varying in response to external fields (polarization) and electron transfer to/from the molecule; this amounts to using a coarse-grained version of Eq. 16. The result is a relatively fast and accurate method for computing molecular interaction energies. In this latter context it bears mention that both of the fundamental expressions for the interaction energy, Eqs 16, 33, can be adapted to cases where more than two molecules need to be considered (e.g., two reactants and a catalyst, several reactant molecules and the surrounding solvent). Assuming that three-body forces are negligible, the interaction energy in these more complicated situations is merely the sum over all the pairwise interaction energies.

This approach is fundamentally related to the electron-density-based energy decomposition analysis (DFT-EDA) (Wu et al., 2009). In DFT-EDA, two fragments are brought together with frozen density (corresponding to the first line in Eqs 16, 33) and then the exact (nonperturbative) electron-transfer energy and polarization-energies are defined and computed using constrained search computations. Like the approach in this paper, DFT-EDA is intended to elucidate chemistry, not to provide numbers. This approach, however, is mainly intended to predict chemistry in the perturbative regime, while DFT-EDA is generally employed for understanding known chemistry using an exact (nonperturbative) expression for the energy.

Finally, this approach is partly justified by the nearsightedness principle in density-functional theory (Li et al., 1993; Kohn, 1996; Baer and Head-Gordon, 1997; Prodan and Kohn, 2005; Prodan, 2006). It is safe to assume that when two reactants are far from each other, the “long-range” effects are mainly electron transfer and electrostatics, and that large-scale induced electron rearrangements are minimal (Fias et al., 2017). This means, in particular, that the estimate of the polarization density in step B, which implicitly assumes that the response of one subsystem to the other is small, and that the (total system) linear response kernel, χAB(r,r), and higher-order responses is negligible (except perhaps for features described by electron transfer and the Fukui function/dual descriptor) when r and r are located on different subsystems (Cardenas et al., 2009b; Fias et al., 2017). That is, in step (B) we assume that the response kernel is nearsighted enough for the approximation χAB(r,r)χA(r,r)+χB(r,r) to be valid. A similar nearsightedness argument applies to the hardness kernel in the alternative expression for the interaction energy in Eq. 16.

3.2 Connections to Previous Results in Density Functional Theory for Chemical Reactivity

The two expressions for the interaction energy, Eqs 16, 33, are the key result of this paper. In certain limits, these expressions recover known results from the DFT-CR literature. For example, suppose that the reactivity trends in fragment A do not depend on the detailed properties of fragment B. Then we can neglect the terms in Eq. 16 that depend on fragment B, obtaining,

ΔUAB12ΔρA(r)ΔρA(r)η[ρA0;r,r]drdr(40)

Minimizing this expression with respect to all ΔρA(r) with ΔNA electrons is equivalent to the variational principle for the Fukui function (Chattaraj et al., 1995; Ayers and Parr, 2000). This establishes the Fukui function as the key nonspecific reactivity indicator for electron-transfer reactivity. Recall that the lowest-curvature normal modes on a potential energy surface correspond to the most facile molecular rearrangements. From Eq. 40 we can approximate the elements of the Hessian matrix as,

UABRiRj=12ρA(r)RiρA(r)Rjη[ρA0;r,r]drdr.(41)

This suggests that eigenvectors of the hardness kernel with small eigenvalues indicate favorable reactive modes. Or, to phrase this in the more conventional way: the eigenvectors of the softness kernel with the highest eigenvalues indicate favorable molecular rearrangements (Nalewajski and Koninski, 1987; Cohen et al., 1995; Cardenas et al., 2006). These simple guiding principles based on Eq. 16 obviously fail when semiquantitative results are needed or when detailed “matching” of reactants plays a pivotal role in the chemical reaction.

Equation 33 may seem very intimidating, but its interpretation is relatively simple. There are strong links to the previous two-reagent expressions in the literature, especially from the work of Parr and Pearson (for electron transfer), (Parr and Pearson, 1983) Berkowitz (for Frontier-orbital interactions), (Berkowitz, 1987) Anderson et al. (for the balance between electrostatics and electron-transfer terms), (Anderson et al., 2007a) and Morell et al. (for electrophile/nucleophile matching) (Morell et al., 2005; Ayers et al., 2007; Morell et al., 2007; Morell et al., 2008; Cardenas et al., 2009b; Labet et al., 2009). For example, if one assumes that all regioselective interactions may be neglected, then the entire expression (33) reduces to the conventional Parr-Pearson quadratic expansion for the electron-transfer energy, (Parr and Pearson, 1983; Ayers et al., 2006)

ΔUAB[ΔNA]ΔNA(μA0μB0)+(ΔNA)22(ηA0+ηB0).(42)

If one furthermore chooses one of the reagents to be a perfect electron donor, the change in energy is equal to the electrophilicity (Parr et al., 1999; Chattaraj et al., 2003; Chattaraj et al., 2006; Liu, 2009b).

If electron transfer is dominant and one wishes to consider regioselectivity, then the leading-order term is the Coulomb attraction between the Fukui functions,

Jf(ΔNA)2fA(r)fB(r)|rr|drdr(43)

This term plays a decisive role in the theory of Frontier-molecular orbital interactions put forth by Berkowitz (Berkowitz, 1987) and it justifies the use of the Fukui potential as a fundamental reactivity indicator (Cardenas et al., 2011b; Cardenas, 2011; Osorio et al., 2011; Yañez et al., 2021). When only the potential is varied, but not the number of electrons, one obtains potential-based reactivity indicators (Ayers and Parr, 2001; Cedillo et al., 2007; Ayers P. W. et al., 2009; Liu et al., 2009; Muñoz et al., 2020).

3.3 Term-By-Term Interpretation of Eq. 33

The Taylor series expansion in Eq. 1 converges subject to reasonable assumptions (Ayers et al., 2005a). Equation 33 is therefore approximate only in two ways: 1) higher order terms in the Taylor series are neglected and 2) only electrostatic effects are included in the model for the external potential change. Both assumptions are accurate for distant, weakly-interacting, reagents with nondegenerate ground states; they will be useful in all cases where the initial approach of the reagents guides the reactivity preferences (e.g., when the transition state is early). Both assumptions can be relaxed. Many higher-order reactivity indicators are known, (Fuentealba and Parr, 1991; Senet, 1996; Geerlings and De Proft, 2008; Cardenas et al., 2009a; Heidar-Zadeh et al., 2016b) though they are difficult to compute and inconvenient to work with. The exact “effective external potential” that a molecule feels in the presence of a reagent can be computed, and various levels of approximation are known (Ayers and Parr, 2001). Some of these approximations are not that much more complicated than the simple electrostatic approximation. (For example, including an “exchange-correlation charge density” is relatively straightforward (Liu et al., 1999; Ayers and Levy, 2001; Ayers and Parr, 2001; Andrade and Aspuru-Guzik, 2011)) Even with the approximations implicit in Eq. 33, however, the main chemical effects that drive reactions are included.

3.3.1 Steric Repulsion and Electron Pairing

The non-additive kinetic and exchange-correlation energy terms,

Tsnon-add[ρA0,ρB0]+Excnon-add[ρA0,ρB0](44)

capture the effects of electon pairing (from combining the electron densities of the fragments) and, when closed shells are pushed together, capture the “Pauli” and “kinetic” portions of the repulsive wall in potential energy curves. The electrostatic repulsion between the fragments also contributes to steric hindrance; this contribution is included in the second line.

We note that this measure of the steric repulsion is different, both physically and mathematically, from the quantification of the steric effect proposed by one of us, (Liu, 2007) and then studied and extended by us and others (Liu, 2007; Nagy, 2007; Liu and Govind, 2008; Torrent-Sucarrat et al., 2009; Ugur et al., 2009; Ess et al., 2010; Liu et al., 2010; Tsirelson et al., 2010; Esquivel et al., 2011; Huang et al., 2011; Fang et al., 2014). Nonetheless, the two measures have some mathematical similarities (in both cases the dominant term is associated with the kinetic energy) and are expected to give similar insights. Note also that Eq. 44 also includes some long-range contributions associated with exchange and correlation and short-ranged contributions from electron pairing that are not true “steric” effects.

3.3.2 Electrostatic Interactions

The electrostatic attractions/repulsions between the isolated fragments are accounted for in the term

(zA0(r)ρA0(r))(zB0(r)ρB0(r))|rr|drdr.(45)

Notice that, Eqs 44, 45 together can already describe weak interactions; they are the key elements in the Gordon-Kim theory, for example (Gordon and Kim, 1972).

3.3.3 Polarization and Dispersion

The term

2(EA(pol)[ΦB0]+EB(pol)[ΦA0])(46)

accounts for the electrostatic stabilization that occurs because the fragments are polarized. (The very last line in Eq. 33 corrects for the “promotion energy” that is required to deform the electron density to make these favorable electrostatic interactions.) The second-order change in electrostatics due to the mutual polarization of the fragments is captured in the term

ΔρA(pol)[ΦB0,r]ΔρB(pol)[ΦA0,r]|rr|drdr(47)

If the interaction-energy expression is iteratively converged (step D in Section 2), this term includes the “induced multipole-induced multipole” interactions that are implicated in dispersion.

3.3.4 Electron Transfer; Fukui Function

The Coulomb interaction between the Fukui functions (Parr and Yang, 1984; Yang et al., 1984; Ayers and Levy, 2000; Ayers P. W. et al., 2009; Osorio et al., 2011; Yañez et al., 2021) is the leading-order purely electron-transfer term,

(ΔNA)2fA(r)fB(r)|rr|drdr(48)

The importance of this term had already been noted by Berkowitz (Berkowitz, 1987). Electron transfer will be favorable when the Coulomb repulsion between the donor Fukui function and the acceptor Fukui function is large; this is associated with strong overlap of the Frontier orbitals. Alternatively, Eq. 48 can be viewed as the attraction between the donor orbital and the acceptor orbital, or between the electron and the hole (Ayers and Levy, 2000). Similar terms are found to play a role in Frontier orbital theories, especially when expansions around the separated atom limit are considered (De Proft et al., 2006; Cardenas et al., 2009b).

Electron transfer also changes the electrostatic interactions between reagents at both the isolated reagent,

ΔNA(fA(r)ΦB0(r)fB(r)ΦA0(r))dr(49)

and polarized reagent

+(ΔNA)22(ΔρA(pol)[ΦB0,r]fB(r)+ΔρB(pol)[ΦA0,r]fA(r))|rr|drdr(50)

levels. Terms similar to Eqs 48, 49 can be important even when electron transfer is small because the asymptotic behavior of the linear response function is governed by the Fukui function (Cardenas et al., 2009b; Fias et al., 2017).

3.3.5 Electron Transfer; Dual Descriptor

The dual descriptor (Morell et al., 2005, 2006; Ayers et al., 2007; Morell et al., 2007; Geerlings et al., 2012) tends to play a role when the reagents are electrostatically flat and the Fukui functions do not establish a strong reactivity preference. The leading-order purely dual-descriptor term is

(ΔNA)44ΔfA(r)ΔfB(r)|rr|drdr.(51)

This expression has appeared previously in the literature, but without derivation (Ayers et al., 2007). Notice that, unlike the interaction between the Fukui functions (Eq. 48), it is favorable to have a large electrostatic attraction between the dual descriptors of the fragments; this corresponds to aligning the nucleophilic portions of one fragment with the electrophilic portions of a different fragment, and vice versa (Morell et al., 2005, 2006; Ayers et al., 2007; Morell et al., 2007; Geerlings et al., 2012).

The change in electron density due to second-order electron transfer also causes a shift in the electrostatic interactions with the isolated (compare Eq. 49)

(ΔNA)22(ΔfA(r)ΦB0(r)+ΔfB(r)ΦA0(r))dr(52)

and polarized (compare Eq. 50)

+(ΔNA)22(ΔρA(pol)[ΦB0,r]ΔfB(r)+ΔρB(pol)[ΦA0,r]ΔfA(r))|rr|drdr(53)

reagents. There is also a coupling to the first-order electron transfer (Fukui function) term,

+(ΔNA)3fA(r)ΔfB(r)fB(r)ΔfA(r)|rr|drdr(54)

If any of these terms are large, it indicates that the electrostatic potential, polarization, or first-order electron-transfer (Fukui function) terms are likely to be dominant. The dual descriptor is also important when describing the asymptotic behavior of the quadratic and cubic density response functions; it has been argued that the dual descriptor’s chemical relevance may be mostly attributed to those nonlinear polarization terms (Cardenas et al., 2009b).

3.3.6 Energy Change From Electron Transfer; Chemical Potential and Hardness

The remaining terms in Eq. 33 represent the change in energy of the isolated reagents. First of all, the isolated reagents change in energy because the number of electrons in the reagents changes. This leads to the term

ΔNA(μA0μB0)+(ΔNA)22(ηA0+ηB0)(55)

The electronic chemical potential is minus one times the electronegativity, (Parr et al., 1978) and measures the intrinsic strength of a Lewis acid or base (Ayers et al., 2006). The chemical hardness is the key to describing the HSAB principle (Parr and Pearson, 1983). The electron-transfer energy expression in Eq. 55 suffices, all by itself, to derive the HSAB principle (Chattaraj et al., 1991; Ayers, 2005; Chattaraj and Ayers, 2005; Ayers and Cardenas, 2013; Cardenas and Ayers, 2013).

3.3.7 Promotion Energy

Polarizing the reagents so that they can form more favorable electrostatic interactions requires mixing excited-state character into the ground-state fragment densities. This raises the electronic energy of the fragments by,

EA(pol)[ΦB0]EB(pol)[ΦA0](56)

though this term is doubly-cancelled out (cf. Eq. 46) by the favorable electrostatic contacts that are created (Ayers, 2007a).

3.3.8 Condensed Reactivity Indicators

It is often convenient to replace the pointwise interactions between reactants in the interaction energy model with interactions between the atoms in the reactants. This gives an atom-condensed version of the reaction energy expression (Yang and Mortier, 1986) which strongly resembles the structure of electronegativity equalization molecular mechanics force fields. Letting wa(r) denote the atomic weight function associated with the αth atom in reactant A, and let Za and Ra denote its nuclear charge and position (consistent with the notation in Eq. 3). The expressions for the atom-condensed charges, condensed Fukui function, (Fuentealba et al., 2016) condensed dual descriptor, and condensed polarization density are, respectively

qα=Zαwα(r)ρA(r)dr(57)
fα=wα(r)fA(r)dr(58)
Δfα=wα(r)ΔfA(r)dr(59)
Δnα=βBwα(r)(δρA(r)δv(r))NA0qβ|rRβ|drdr.(60)

Here we have elected to use the fragment-of-molecular response method, though the response-of-molecular-fragment method could be used instead (Fuentealba et al., 2000; Ayers et al., 2002; Bultinck et al., 2007; Padmanabhan et al., 2007; Zielinski et al., 2012; Miranda-Quintana R. A., 2016). We could also include dipole (and higher-order multipole) components of the descriptors in Eq. 57, but this would complicate the formula for the polarization energy

EA(pol)=12αAβBΔnαqβ|RαRβ|(61)

and the interaction energy

ΔUAB[ΔNA]=(Tsnon-add[ρA0,ρB0]+Excnon-add[ρA0,ρB0])+(αAβBqαqβ|RαRβ|+2(EA(pol)+EB(pol))+αAβBΔnαΔnβ|RαRβ|ΔNAαAβBfαqβqαfβ|RαRβ|+ΔNAαAβBfαΔnβΔnαfβ|RαRβ|(ΔNA)2αAβBfαfβ|RαRβ|(ΔNA)22αAβBΔfαqβ+qαΔfβ|RαRβ|+(ΔNA)22αAβBΔfαΔnβ+ΔnαΔfβ|RαRβ|+(ΔNA)3αAβBfαΔfβΔfαfβ|RαRβ|+(ΔNA)44αAβBΔfαΔfβ|RαRβ|)+(ΔNA(μA0μB0)+(ΔNA)22(ηA0+ηB0)EA(pol)EB(pol))(62)

This equation is more convenient as a “working formula” than Eq. 33 because it removes the need to perform six-dimensional integrals. We note, however, that once one chooses a suitable energy model, (Parr and Bartolotti, 1982; Parr and Pearson, 1983; Noorizadeh and Shakerzadeh, 2008; Fuentealba and Cardenas, 2013; Heidar-Zadeh et al., 2016a; Miranda-Quintana R. A. and Ayers P. W., 2016) all the integrations in Eq. 33 can be performed analytically by combining the basis-set expansion of one-electron density matrices and the linear response function with the appropriate one- and two-electron integrals.

4 Summary

In the density functional theory approach to chemical reactivity (DFT-CR), it is traditional to focus on the reactivity indicators of a single reagent. This is not always adequate, as interactions between reagents can be decisive when, for example, partner-specific interactions are critically important for determining reactivity preferences. In this work, we have derived mathematical formulas for the interaction energy between reagents in terms of chemical reactivity indicators (Eqs. 16, 33) and shown how these expressions are related to existing computational approaches (density-functional embedding, density-functional energy decomposition analysis) and previous reactivity rules (Section 3.2). We also showed that all of the commonly accepted chemical interactions that influence chemical reactivity are contained in the key Eq. 33 (Section 3.3). We believe that this work provides not only a basis for future work in DFT-CR, but that we can rationalize existing DFT-CR success stories by noting that whenever one term in Eq. 33 is decisive, the conventional one-reagent-one-reactivity-indicator approach is likely to be successful. In further contributions in this series we will explore how the two-reagent picture can be used to provide further analytical arguments in favor of several reactivity principles.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Author Contributions

PA, SL, CC and RAM-Q contributed to conception and design of the study. PA wrote the first draft of the manuscript. CC, RAM-Q, TG wrote sections of the manuscript. All authors contributed to manuscript revision, read, and approved the submitted version.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

PA thanks the Canada Research Chairs, NSERC, and Compute Canada for funding. CC acknowledges financial support from Fondecyt through grant No 1220366 and also by Centers of Excellence With Basal-Conicyt Financing, Grant FB0807. RAM-Q thanks support from the University of Florida in the form of a startup grant.

References

Alain Miranda-Quintana, R., and Ayers, P. W. (2016a). Charge Transfer and Chemical Potential in 1,3-dipolar Cycloadditions. Theor. Chem. Accounts 135. doi:10.1007/s00214-016-1924-7

CrossRef Full Text | Google Scholar

Alain Miranda-Quintana, R., and Ayers, P. W. (2016b). Fractional Electron Number, Temperature, and Perturbations in Chemical Reactions. Phys. Chem. Chem. Phys. 18, 15070–15080. doi:10.1039/c6cp00939e

PubMed Abstract | CrossRef Full Text | Google Scholar

Alain Miranda-Quintana, R., Martínez González, M., and Ayers, P. W. (2016). Electronegativity and Redox Reactions. Phys. Chem. Chem. Phys. 18, 22235–22243. doi:10.1039/c6cp03213c

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, J. S. M., Melin, J., and Ayers, P. W. (2007a). Conceptual Density-Functional Theory for General Chemical Reactions, Including Those that Are Neither Charge- Nor Frontier-Orbital-Controlled. 1. Theory and Derivation of a General-Purpose Reactivity Indicator. J. Chem. Theory Comput. 3, 358–374. doi:10.1021/ct600164j

PubMed Abstract | CrossRef Full Text | Google Scholar

Anderson, J. S. M., Melin, J., and Ayers, P. W. (2007b). Conceptual Density-Functional Theory for General Chemical Reactions, Including Those that Are Neither Charge- Nor Frontier-Orbital-Controlled. 2. Application to Molecules where Frontier Molecular Orbital Theory Fails. J. Chem. Theory Comput. 3, 375–389. doi:10.1021/ct6001658

PubMed Abstract | CrossRef Full Text | Google Scholar

Andrade, X., and Aspuru-Guzik, A. (2011). Prediction of the Derivative Discontinuity in Density Functional Theory from an Electrostatic Description of the Exchange and Correlation Potential. Phys. Rev. Lett. 107, 183002. doi:10.1103/physrevlett.107.183002

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W. (2005). An Elementary Derivation of the Hard/soft-Acid/base Principle. J. Chem. Phys. 122, 141102. doi:10.1063/1.1897374

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., Anderson, J. S. M., and Bartolotti, L. J. (2005a). Perturbative Perspectives on the Chemical Reaction Prediction Problem. Int. J. Quantum Chem. 101, 520–534. doi:10.1002/qua.20307

CrossRef Full Text | Google Scholar

Ayers, P. W., Anderson, J. S. M., Rodriguez, J. I., and Jawed, Z. (2005b). Indices for Predicting the Quality of Leaving Groups. Phys. Chem. Chem. Phys. 7, 1918–1925. doi:10.1039/b500996k

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W. (2000). Atoms in Molecules, an Axiomatic Approach. I. Maximum Transferability. J. Chem. Phys. 113, 10886–10898. doi:10.1063/1.1327268

CrossRef Full Text | Google Scholar

Ayers, P. W., and Cárdenas, C. (2013). Communication: A Case where the Hard/soft Acid/base Principle Holds Regardless of Acid/base Strength. J. Chem. Phys. 138, 181106. doi:10.1063/1.4805083

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., and Fuentealba, P. (2009). Density-functional Theory with Additional Basic Variables: Extended Legendre Transform. Phys. Rev. A 80, 032510. doi:10.1103/physreva.80.032510

CrossRef Full Text | Google Scholar

Ayers, P. W., and Levy, M. (2000). Perspective on "Density Functional Approach to the Frontier-Electron Theory of Chemical Reactivity". Theor. Chem. Accounts 103, 353–360. doi:10.1007/978-3-662-10421-7_59

CrossRef Full Text | Google Scholar

Ayers, P. W., and Levy, M. (2001). Sum Rules for Exchange and Correlation Potentials. J. Chem. Phys. 115, 4438–4443. doi:10.1063/1.1379333

CrossRef Full Text | Google Scholar

Ayers, P. W., Liu, S., and Li, T. (2009a). Chargephilicity and Chargephobicity: Two New Reactivity Indicators for External Potential Changes from Density Functional Reactivity Theory. Chem. Phys. Lett. 480, 318–321. doi:10.1016/j.cplett.2009.08.067

CrossRef Full Text | Google Scholar

Ayers, P. W., Morell, C., De Proft, F., and Geerlings, P. (2007). Understanding the Woodward-Hoffmann Rules by Using Changes in Electron Density. Chem. Eur. J. 13, 8240–8247. doi:10.1002/chem.200700365

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., Morrison, R. C., and Roy, R. K. (2002). Variational Principles for Describing Chemical Reactions: Condensed Reactivity Indices. J. Chem. Phys. 116, 8731–8744. doi:10.1063/1.1467338

CrossRef Full Text | Google Scholar

Ayers, P. W. (2007a). On the Electronegativity Nonlocality Paradox. Theor. Chem. Acc. 118, 371–381. doi:10.1007/s00214-007-0277-7

CrossRef Full Text | Google Scholar

Ayers, P. W., and Parr, R. G. (2008). Local Hardness Equalization: Exploiting the Ambiguity. J. Chem. Phys. 128, 184108. doi:10.1063/1.2918731

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., Parr, R. G., and Pearson, R. G. (2006). Elucidating the Hard/soft Acid/base Principle: A Perspective Based on Half-Reactions. J. Chem. Phys. 124, 194107. doi:10.1063/1.2196882

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., and Parr, R. G. (2000). Variational Principles for Describing Chemical Reactions: The Fukui Function and Chemical Hardness Revisited. J. Am. Chem. Soc. 122, 2010–2018. doi:10.1021/ja9924039

CrossRef Full Text | Google Scholar

Ayers, P. W., and Parr, R. G. (2001). Variational Principles for Describing Chemical Reactions. Reactivity Indices Based on the External Potential. J. Am. Chem. Soc. 123, 2007–2017. doi:10.1021/ja002966g

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W. (2001). Strategies for Computing Chemical Reactivity Indices. Theor. Chem. Acc. 106, 271–279. doi:10.1007/pl00012385

CrossRef Full Text | Google Scholar

Ayers, P. W. (2008). The Dependence on and Continuity of the Energy and Other Molecular Properties with Respect to the Number of Electrons. J. Math. Chem. 43, 285–303. doi:10.1007/s10910-006-9195-5

CrossRef Full Text | Google Scholar

Ayers, P. W. (2007b). The Physical Basis of the Hard/soft Acid/base Principle. Faraday Discuss. 135, 161–190. doi:10.1039/b606877d

PubMed Abstract | CrossRef Full Text | Google Scholar

Ayers, P. W., Yang, W. T., and Bartolotti, L. J. (2009b). “Fukui Function,” in Chemical Reactivity Theory: A Density Functional View. Editor P. K. Chattaraj (Boca Raton: CRC Press), 255–267. doi:10.1201/9781420065442.ch18

CrossRef Full Text | Google Scholar

Azar, R. J., and Head-Gordon, M. (2012). An Energy Decomposition Analysis for Intermolecular Interactions from an Absolutely Localized Molecular Orbital Reference at the Coupled-Cluster Singles and Doubles Level. J. Chem. Phys. 136, 024103. doi:10.1063/1.3674992

PubMed Abstract | CrossRef Full Text | Google Scholar

Baer, R., and Head-Gordon, M. (1997). Sparsity of the Density Matrix in Kohn-Sham Density Functional Theory and an Assessment of Linear System-Size Scaling Methods. Phys. Rev. Lett. 79, 3962–3965. doi:10.1103/physrevlett.79.3962

CrossRef Full Text | Google Scholar

Berkowitz, M. (1987). Density Functional Approach to Frontier Controlled Reactions. J. Am. Chem. Soc. 109, 4823–4825. doi:10.1021/ja00250a012

CrossRef Full Text | Google Scholar

Berkowitz, M., and Parr, R. G. (1988). Molecular Hardness and Softness, Local Hardness and Softness, Hardness and Softness Kernels, and Relations Among These Quantities. J. Chem. Phys. 88, 2554–2557. doi:10.1063/1.454034

CrossRef Full Text | Google Scholar

Bochicchio, R. C. (2015). On the Non-integer Number of Particles in Molecular System Domains: Treatment and Description. Theor. Chem. Acc. 134, 138. doi:10.1007/s00214-015-1743-2

CrossRef Full Text | Google Scholar

Bultinck, P., Cardenas, C., Fuentealba, P., Johnson, P. A., and Ayers, P. W. (2013). Atomic Charges and the Electrostatic Potential Are Ill-Defined in Degenerate Ground States. J. Chem. Theory Comput. 9, 4779–4788. doi:10.1021/ct4005454

PubMed Abstract | CrossRef Full Text | Google Scholar

Bultinck, P., Cardenas, C., Fuentealba, P., Johnson, P. A., and Ayers, P. W. (2014). How to Compute the Fukui Matrix and Function for Systems with (Quasi-)Degenerate States. J. Chem. Theory Comput. 10, 202–210. doi:10.1021/ct400874d

PubMed Abstract | CrossRef Full Text | Google Scholar

Bultinck, P., Fias, S., Van Alsenoy, C., Ayers, P. W., and Carbó-Dorca, R. (2007). Critical Thoughts on Computing Atom Condensed Fukui Functions. J. Chem. Phys. 127, 034102. doi:10.1063/1.2749518

PubMed Abstract | CrossRef Full Text | Google Scholar

Bultinck, P., Jayatilaka, D., and Cardenas, C. (2015). A Problematic Issue for Atoms in Molecules: Impact of (Quasi-)degenerate States on Quantum Theory Atoms in Molecules and Hirshfeld-I Properties. Comput. Theor. Chem. 1053, 106–111. doi:10.1016/j.comptc.2014.06.017

CrossRef Full Text | Google Scholar

Bultinck, P., Langenaeker, W., Lahorte, P., De Proft, F., Geerlings, P., Van Alsenoy, C., et al. (2002a). The Electronegativity Equalization Method II: Applicability of Different Atomic Charge Schemes. J. Phys. Chem. A 106, 7895–7901. doi:10.1021/jp020547v

CrossRef Full Text | Google Scholar

Bultinck, P., Langenaeker, W., Lahorte, P., De Proft, F., Geerlings, P., Waroquier, M., et al. (2002b). The Electronegativity Equalization Method I: Parametrization and Validation for Atomic Charge Calculations. J. Phys. Chem. A 106, 7887–7894. doi:10.1021/jp0205463

CrossRef Full Text | Google Scholar

Cárdenas, C., Ayers, P. W., and Cedillo, A. (2011a). Reactivity Indicators for Degenerate States in the Density-Functional Theoretic Chemical Reactivity Theory. J. Chem. Phys. 134, 174103–174113. doi:10.1063/1.3585610

CrossRef Full Text | Google Scholar

Cárdenas, C., and Ayers, P. W. (2013). How Reliable Is the Hard-Soft Acid-Base Principle? an Assessment from Numerical Simulations of Electron Transfer Energies. Phys. Chem. Chem. Phys. 15, 13959–13968. doi:10.1039/c3cp51134k

PubMed Abstract | CrossRef Full Text | Google Scholar

Cárdenas, C., Echegaray, E., Chakraborty, D., Anderson, J. S. M., and Ayers, P. W. (2009a). Relationships between the Third-Order Reactivity Indicators in Chemical Density-Functional Theory. J. Chem. Phys. 130, 244105. doi:10.1063/1.3151599

CrossRef Full Text | Google Scholar

Cárdenas, C., Lamsabhi, A. M., and Fuentealba, P. (2006). Nuclear Reactivity Indices in the Context of Spin Polarized Density Functional Theory. Chem. Phys. 322, 303–310. doi:10.1016/j.chemphys.2005.09.001

CrossRef Full Text | Google Scholar

Cárdenas, C., Rabi, N., Ayers, P. W., Morell, C., Jaramillo, P., and Fuentealba, P. (2009b). Chemical Reactivity Descriptors for Ambiphilic Reagents: Dual Descriptor, Local Hypersoftness, and Electrostatic Potential. J. Phys. Chem. A 113, 8660–8667. doi:10.1021/jp902792n

CrossRef Full Text | Google Scholar

Cárdenas, C. (2011). The Fukui Potential Is a Measure of the Chemical Hardness. Chem. Phys. Lett. 513, 127–129. doi:10.1016/j.cplett.2011.07.059

CrossRef Full Text | Google Scholar

Cárdenas, C., Tiznado, W., Ayers, P. W., and Fuentealba, P. (2011b). The Fukui Potential and the Capacity of Charge and the Global Hardness of Atoms. J. Phys. Chem. A 115, 2325–2331. doi:10.1021/jp109955q

CrossRef Full Text | Google Scholar

Cedillo, A., Contreras, R., Galván, M., Aizman, A., Andrés, J., and Safont, V. S. (2007). Nucleophilicity Index from Perturbed Electrostatic Potentials. J. Phys. Chem. A 111, 2442–2447. doi:10.1021/jp068459o

PubMed Abstract | CrossRef Full Text | Google Scholar

Chattaraj, P. K., and Ayers, P. W. (2005). The Maximum Hardness Principle Implies the Hard/soft Acid/base Rule. J. Chem. Phys. 123, 086101. doi:10.1063/1.2011395

PubMed Abstract | CrossRef Full Text | Google Scholar

Chattaraj, P. K., Cedillo, A., and Parr, R. G. (1995). Variational Method for Determining the Fukui Function and Chemical Hardness of an Electronic System. J. Chem. Phys. 103, 7645–7646. doi:10.1063/1.470284

CrossRef Full Text | Google Scholar

Chattaraj, P. K., Lee, H., and Parr, R. G. (1991). HSAB Principle. J. Am. Chem. Soc. 113, 1855–1856. doi:10.1021/ja00005a073

CrossRef Full Text | Google Scholar

Chattaraj, P. K., and Maiti, B. (2001). Reactivity Dynamics in Atom−Field Interactions: A Quantum Fluid Density Functional Study. J. Phys. Chem. A 105, 169–183. doi:10.1021/jp0019660

CrossRef Full Text | Google Scholar

Chattaraj, P. K., Maiti, B., and Sarkar, U. (2003). Philicity: A Unified Treatment of Chemical Reactivity and Selectivity. J. Phys. Chem. A 107, 4973–4975. doi:10.1021/jp034707u

CrossRef Full Text | Google Scholar

Chattaraj, P. K., Sarkar, U., and Roy, D. R. (2006). Electrophilicity Index. Chem. Rev. 106, 2065–2091. doi:10.1021/cr040109f

PubMed Abstract | CrossRef Full Text | Google Scholar

Chermette, H. (1999). Chemical Reactivity Indexes in Density Functional Theory. J. Comput. Chem. 20, 129–154. doi:10.1002/(sici)1096-987x(19990115)20:1<129:aid-jcc13>3.0.co;2-a

CrossRef Full Text | Google Scholar

Cohen, M. H., Ganduglia‐Pirovano, M. V., and Kudrnovský, J. (1995). Reactivity Kernels, the Normal Modes of Chemical Reactivity, and the Hardness and Softness Spectra. J. Chem. Phys. 103, 3543–3551. doi:10.1063/1.470238

CrossRef Full Text | Google Scholar

Cortona, P. (1991). Self-Consistently Determined Properties of Solids without Band-Structure Calculations. Phys. Rev. B 44, 8454–8458. doi:10.1103/physrevb.44.8454

PubMed Abstract | CrossRef Full Text | Google Scholar

De Proft, F., Ayers, P. W., Fias, S., and Geerlings, P. (2006). Woodward-hoffmann Rules in Density Functional Theory: Initial Hardness Response. J. Chem. Phys. 125, 214101. doi:10.1063/1.2387953

PubMed Abstract | CrossRef Full Text | Google Scholar

De Proft, F., Chattaraj, P. K., Ayers, P. W., Torrent-Sucarrat, M., Elango, M., Subramanian, V., et al. (2008). Initial Hardness Response and Hardness Profiles in the Study of Woodward-Hoffmann Rules for Electrocyclizations. J. Chem. Theory Comput. 4, 595–602. doi:10.1021/ct700289p

PubMed Abstract | CrossRef Full Text | Google Scholar

De Proft, F., Geerlings, P., Liu, S., and Parr, R. G. (1998). Variational Calculation of the Global Hardness and the Fukui Function via an Approximation of the Hardness Kernel. Pol. J. Chem. 72, 1737–1746.

Google Scholar

De Proft, F., Liu, S., and Parr, R. G. (1997). Chemical Potential, Hardness, Hardness and Softness Kernel and Local Hardness in the Isomorphic Ensemble of Density Functional Theory. J. Chem. Phys. 107, 3000–3006. doi:10.1063/1.474657

CrossRef Full Text | Google Scholar

Elliott, P., Burke, K., Cohen, M. H., and Wasserman, A. (2010). Partition Density-Functional Theory. Phys. Rev. A 82, 024501. doi:10.1103/physreva.82.024501

CrossRef Full Text | Google Scholar

Esquivel, R. O., Liu, S., Angulo, J. C., Dehesa, J. S., Antolín, J., and Molina-Espíritu, M. (2011). Fisher Information and Steric Effect: Study of the Internal Rotation Barrier of Ethane. J. Phys. Chem. A 115, 4406–4415. doi:10.1021/jp1095272

PubMed Abstract | CrossRef Full Text | Google Scholar

Ess, D. H., Liu, S., and De Proft, F. (2010). Density Functional Steric Analysis of Linear and Branched Alkanes. J. Phys. Chem. A 114, 12952–12957. doi:10.1021/jp108577g

PubMed Abstract | CrossRef Full Text | Google Scholar

Fang, D., Piquemal, J.-P., Liu, S., and Cisneros, G. A. (2014). DFT-steric-based Energy Decomposition Analysis of Intermolecular Interactions. Theor. Chem. Accounts 133, 1484. doi:10.1007/s00214-014-1484-7

CrossRef Full Text | Google Scholar

Fias, S., Heidar-Zadeh, F., Geerlings, P., and Ayers, P. W. (2017). Chemical Transferability of Functional Groups Follows from the Nearsightedness of Electronic Matter. Proc. Natl. Acad. Sci. U.S.A. 114, 11633–11638. doi:10.1073/pnas.1615053114

PubMed Abstract | CrossRef Full Text | Google Scholar

Franco-Pérez, M., Ayers, P. W., Gázquez, J. L., and Vela, A. (2015a). Local and Linear Chemical Reactivity Response Functions at Finite Temperature in Density Functional Theory. J. Chem. Phys. 143, 244117. doi:10.1063/1.4938422

CrossRef Full Text | Google Scholar

Franco-Pérez, M., Ayers, P. W., Gázquez, J. L., and Vela, A. (2017b). Thermodynamic Responses of Electronic Systems. J. Chem. Phys. 147, 094105. doi:10.1063/1.4999761

CrossRef Full Text | Google Scholar

Franco-Pérez, M., Gázquez, J. L., Ayers, P. W., and Vela, A. (2015b). Revisiting the Definition of the Electronic Chemical Potential, Chemical Hardness, and Softness at Finite Temperatures. J. Chem. Phys. 143, 154103. doi:10.1063/1.4932539

CrossRef Full Text | Google Scholar

Franco-Pérez, M., Gázquez, J. L., Ayers, P. W., and Vela, A. (2017c). Thermodynamic Hardness and the Maximum Hardness Principle. J. Chem. Phys. 147, 074113. doi:10.1063/1.4998701

CrossRef Full Text | Google Scholar

Franco-Perez, M., Ayers, P. W., and Gazquez, J. L. (2016). Average Electronic Energy Is the Central Quantity in Conceptual Chemical Reactivity Theory. Theor. Chem. Accounts 135, 199. doi:10.1007/s00214-016-1961-2

CrossRef Full Text | Google Scholar

Franco-Pérez, M., Ayers, P. W., Gázquez, J. L., and Vela, A. (2017a). Local Chemical Potential, Local Hardness, and Dual Descriptors in Temperature Dependent Chemical Reactivity Theory. Phys. Chem. Chem. Phys. 19, 13687–13695. doi:10.1039/c7cp00692f

CrossRef Full Text | Google Scholar

Franco-PéRez, M., Gázquez, J. L., Ayers, P. W., and Vela, A. (2018). Thermodynamic Justification for the Parabolic Model for Reactivity Indicators with Respect to Electron Number and a Rigorous Definition for the Electrophilicity: the Essential Role Played by the Electronic Entropy. J. Chem. Theory Comput. 14, 597–606.

PubMed Abstract | Google Scholar

Franco-Pérez, M., Heidar-Zadeh, F., Ayers, P. W., Gázquez, J. L., and Vela, A. (2017d). Going beyond the Three-State Ensemble Model: the Electronic Chemical Potential and Fukui Function for the General Case. Phys. Chem. Chem. Phys. 19, 11588–11602. doi:10.1039/c7cp00224f

CrossRef Full Text | Google Scholar

Franco-Pérez, M., Polanco-Ramírez, C.-A., Ayers, P. W., Gázquez, J. L., and Vela, A. (2017e). New Fukui, Dual and Hyper-Dual Kernels as Bond Reactivity Descriptors. Phys. Chem. Chem. Phys. 19, 16095–16104. doi:10.1039/c7cp02613g

CrossRef Full Text | Google Scholar

Fuentealba, P., and Cardenas, C. (2015). “Density Functional Theory of Chemical Reactivity,” in Chemical Modelling: Volume 11. Editor J. O. J. Michael Springborg (London, UK: The Royal Society of Chemistry), 151–174.

Google Scholar

Fuentealba, P., and Cárdenas, C. (2013). On the Exponential Model for Energy with Respect to Number of Electrons. J. Mol. Model. 19, 2849–2853. doi:10.1007/s00894-012-1708-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Fuentealba, P., Cardenas, C., Pino-Rios, R., and Tiznado, W. (2016). “Topological Analysis of the Fukui Function,” in Applications of Topological Methods in Molecular Chemistry. Editors C. L. Remi Chauvin, B. Silvi, and E. Alikhani (Berlin, Germany: Springer International Publishing), 227–241. doi:10.1007/978-3-319-29022-5_8

CrossRef Full Text | Google Scholar

Fuentealba, P., and Parr, R. G. (1991). Higher‐order Derivatives in Density‐functional Theory, Especially the Hardness Derivative ∂η/∂N. J. Chem. Phys. 94, 5559–5564. doi:10.1063/1.460491

CrossRef Full Text | Google Scholar

Fuentealba, P., Pérez, P., and Contreras, R. (2000). On the Condensed Fukui Function. J. Chem. Phys. 113, 2544–2551. doi:10.1063/1.1305879

CrossRef Full Text | Google Scholar

Fux, S., Jacob, C. R., Neugebauer, J., Visscher, L., and Reiher, M. (2010). Accurate Frozen-Density Embedding Potentials as a First Step towards a Subsystem Description of Covalent Bonds. J. Chem. Phys. 132, 164101. doi:10.1063/1.3376251

PubMed Abstract | CrossRef Full Text | Google Scholar

Galvan, M., and Vargas, R. (1992). Spin-potential in Kohn-Sham Theory. J. Phys. Chem. 96, 1625–1630. doi:10.1021/j100183a026

CrossRef Full Text | Google Scholar

Galvan, M., Vela, A., and Gazquez, J. L. (1988). Chemical Reactivity in Spin-Polarized Density Functional Theory. J. Phys. Chem. 92, 6470–6474. doi:10.1021/j100333a056

CrossRef Full Text | Google Scholar

Garza, J., Vargas, R., Cedillo, A., Galván, M., and Chattaraj, P. K. (2006). Comparison between the Frozen Core and Finite Differences Approximations for the Generalized Spin-dependent Global and Local Reactivity Descriptors in Small Molecules. Theor. Chem. Acc. 115, 257–265. doi:10.1007/s00214-005-0002-3

CrossRef Full Text | Google Scholar

Gázquez, J. L., Cedillo, A., and Vela, A. (2007). Electrodonating and Electroaccepting Powers. J. Phys. Chem. A 111, 1966–1970. doi:10.1021/jp065459f

PubMed Abstract | CrossRef Full Text | Google Scholar

Gázquez, J. L., Franco‐Pérez, M., Ayers, P. W., and Vela, A. (2019). Temperature‐dependent Approach to Chemical Reactivity Concepts in Density Functional Theory. Int. J. Quantum Chem. 119, e25797. doi:10.1002/qua.25797

CrossRef Full Text | Google Scholar

Gazquez, J. L. (2008). Perspectives on the Density Functional Theory of Chemical Reactivity. J. Mexican Chem. Soc. 52, 3–10.

Google Scholar

Geerlings, P., Ayers, P. W., Toro-Labbé, A., Chattaraj, P. K., and De Proft, F. (2012). The Woodward-Hoffmann Rules Reinterpreted by Conceptual Density Functional Theory. Acc. Chem. Res. 45, 683–695. doi:10.1021/ar200192t

PubMed Abstract | CrossRef Full Text | Google Scholar

Geerlings, P., and De Proft, F. (2008). Conceptual DFT: the Chemical Relevance of Higher Response Functions. Phys. Chem. Chem. Phys. 10, 3028–3042. doi:10.1039/b717671f

PubMed Abstract | CrossRef Full Text | Google Scholar

Geerlings, P., De Proft, F., and Langenaeker, W. (2003). Conceptual Density Functional Theory. Chem. Rev. 103, 1793–1874. doi:10.1021/cr990029p

PubMed Abstract | CrossRef Full Text | Google Scholar

Geerlings, P., Fias, S., Boisdenghien, Z., and De Proft, F. (2014). Conceptual DFT: Chemistry from the Linear Response Function. Chem. Soc. Rev. 43, 14. doi:10.1039/c3cs60456j

PubMed Abstract | CrossRef Full Text | Google Scholar

Ghanty, T. K., and Ghosh, S. K. (1996). A Density Functional Approach to Hardness, Polarizability, and Valency of Molecules in Chemical Reactions. J. Phys. Chem. 100, 12295–12298. doi:10.1021/jp960276m

CrossRef Full Text | Google Scholar

Ghanty, T. K., and Ghosh, S. K. (1994). Spin-Polarized Generalization of the Concepts of Electronegativity and Hardness and the Description of Chemical Binding. J. Am. Chem. Soc. 116, 3943–3948. doi:10.1021/ja00088a033

CrossRef Full Text | Google Scholar

Ghosh, S. K. (1994). Electronegativity, Hardness, and a Semiempirical Density Functional Theory of Chemical Binding. Int. J. Quantum Chem. 49, 239–251. doi:10.1002/qua.560490314

CrossRef Full Text | Google Scholar

Gómez, T., Fuentealba, P., Robles‐Navarro, A., and Cárdenas, C. (2021). Links Among the Fukui Potential, the Alchemical Hardness and the Local Hardness of an Atom in a Molecule. J. Comput. Chem. 42, 1681–1688.

PubMed Abstract | Google Scholar

Gonzalez, M. M., Cardenas, C., Rodríguez, J. I., Liu, S., Heidar-Zadeh, F., Miranda-Quintana, R. A., et al. (2018). Quantitative Electrophilicity Measures. Acta Physico-Chimica Sinca 34, 662–674.

Google Scholar

Goodpaster, J. D., Ananth, N., Manby, F. R., and Miller, T. F. (2010). Exact Nonadditive Kinetic Potentials for Embedded Density Functional Theory. J. Chem. Phys. 133, 084103. doi:10.1063/1.3474575

PubMed Abstract | CrossRef Full Text | Google Scholar

Goodpaster, J. D., Barnes, T. A., and Miller, T. F. (2011). Embedded Density Functional Theory for Covalently Bonded and Strongly Interacting Subsystems. J. Chem. Phys. 134, 164108. doi:10.1063/1.3582913

PubMed Abstract | CrossRef Full Text | Google Scholar

Gordon, R. G., and Kim, Y. S. (1972). Theory for the Forces between Closed‐Shell Atoms and Molecules. J. Chem. Phys. 56, 3122–3133. doi:10.1063/1.1677649

CrossRef Full Text | Google Scholar

Govind, N., Wang, Y. A., and Carter, E. A. (1999). Electronic-structure Calculations by First-Principles Density-Based Embedding of Explicitly Correlated Systems. J. Chem. Phys. 110, 7677–7688. doi:10.1063/1.478679

CrossRef Full Text | Google Scholar

Heidar-Zadeh, F., Miranda-Quintana, R. A., Verstraelen, T., Bultinck, P., and Ayers, P. W. (2016a). When Is the Fukui Function Not Normalized? the Danger of Inconsistent Energy Interpolation Models in Density Functional Theory. J. Chem. Theory Comput. 12, 5777–5787. doi:10.1021/acs.jctc.6b00494

PubMed Abstract | CrossRef Full Text | Google Scholar

Heidar-Zadeh, F., Richer, M., Fias, S., Miranda-Quintana, R. A., Chan, M., Franco-Pérez, M., et al. (2016b). An Explicit Approach to Conceptual Density Functional Theory Descriptors of Arbitrary Order. Chem. Phys. Lett. 660, 307–312. doi:10.1016/j.cplett.2016.07.039

CrossRef Full Text | Google Scholar

Hoffmann, G., Tognetti, V., and Joubert, L. (2020). Electrophilicity Indices and Halogen Bonds: Some New Alternatives to the Molecular Electrostatic Potential. J. Phys. Chem. A 124, 2090–2101. doi:10.1021/acs.jpca.9b10233

PubMed Abstract | CrossRef Full Text | Google Scholar

Hohenberg, P., and Kohn, W. (1964). Inhomogeneous Electron Gas. Phys. Rev. 136, B864–B871. doi:10.1103/physrev.136.b864

CrossRef Full Text | Google Scholar

Huang, Y., Zhong, A.-G., Yang, Q., and Liu, S. (2011). Origin of Anomeric Effect: A Density Functional Steric Analysis. J. Chem. Phys. 134, 084103. doi:10.1063/1.3555760

PubMed Abstract | CrossRef Full Text | Google Scholar

Johnson, P. A., Bartolotti, L. J., Ayers, P. W., Fievez, T., and Geerlings, P. (2012a). “Charge Density and Chemical Reactions: a Unified View from Conceptual DFT,” in Modern Charge-Density Analysis (New York: Springer), 715–764.

Google Scholar

Johnson, P. A., Bartolotti, L. J., Ayers, P. W., Fievez, T., and Geerlings, P. (2012b). “Charge Density and Chemical Reactivity: A Unified View from Conceptual DFT,” in Modern Charge Density Analysis. Editors C. Gatti, and P. Macchi (New York: Springer), 715–764.

Google Scholar

Kiewisch, K., Eickerling, G., Reiher, M., and Neugebauer, J. (2008). Topological Analysis of Electron Densities from Kohn-Sham and Subsystem Density Functional Theory. J. Chem. Phys. 128, 044114. doi:10.1063/1.2822966

PubMed Abstract | CrossRef Full Text | Google Scholar

Kitaura, K., and Morokuma, K. (1976). A New Energy Decomposition Scheme for Molecular Interactions within the Hartree-Fock Approximation. Int. J. Quantum Chem. 10, 325–340. doi:10.1002/qua.560100211

CrossRef Full Text | Google Scholar

Kohn, W. (1996). Density Functional and Density Matrix Method Scaling Linearly with the Number of Atoms. Phys. Rev. Lett. 76, 3168–3171. doi:10.1103/physrevlett.76.3168

PubMed Abstract | CrossRef Full Text | Google Scholar

Kohn, W., and Sham, L. J. (1965). Self-consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 140, A1133–A1138. doi:10.1103/physrev.140.a1133

CrossRef Full Text | Google Scholar

Labet, V., Morell, C., Cadet, J., Eriksson, L. A., and Grand, A. (2009). Hydrolytic Deamination of 5-Methylcytosine in Protic Medium-A Theoretical Study. J. Phys. Chem. A 113, 2524–2533. doi:10.1021/jp808902j

PubMed Abstract | CrossRef Full Text | Google Scholar

Laricchia, S., Fabiano, E., and Della Sala, F. (2011). Frozen Density Embedding Calculations with the Orbital-dependent Localized Hartree-Fock Kohn-Sham Potential. Chem. Phys. Lett. 518, 114–118. doi:10.1016/j.cplett.2011.10.055

CrossRef Full Text | Google Scholar

Li, X.-P., Nunes, R. W., and Vanderbilt, D. (1993). Density-matrix Electronic-Structure Method with Linear System-Size Scaling. Phys. Rev. B 47, 10891–10894. doi:10.1103/physrevb.47.10891

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Ayers, P. W., and Parr, R. G. (1999). Alternative Definition of Exchange-Correlation Charge in Density Functional Theory. J. Chem. Phys. 111, 6197–6203. doi:10.1063/1.479924

CrossRef Full Text | Google Scholar

Liu, S. B. (2009a). Conceptual Density Functional Theory and Some Recent Developments. Acta Physico-Chimica Sin. 25, 590–600.

Google Scholar

Liu, S. B. (2009b). “Electrophilicity,” in Chemical Reactivity Theory: A Density Functional View. Editor P. K. Chattaraj (Boca Raton: Taylor & Francis), 179. doi:10.1201/9781420065442.ch13

CrossRef Full Text | Google Scholar

Liu, S., De Proft, F., and Parr, R. G. (1997). Simplified Models for Hardness Kernel and Calculations of Global Hardness. J. Phys. Chem. A 101, 6991–6997. doi:10.1021/jp971263r

CrossRef Full Text | Google Scholar

Liu, S., and Govind, N. (2008). Toward Understanding the Nature of Internal Rotation Barriers with a New Energy Partition Scheme: Ethane and N-Butane. J. Phys. Chem. A 112, 6690–6699. doi:10.1021/jp800376a

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Hu, H., and Pedersen, L. G. (2010). Steric, Quantum, and Electrostatic Effects on SN2 Reaction Barriers in Gas Phase. J. Phys. Chem. A 114, 5913–5918. doi:10.1021/jp101329f

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Li, T., and Ayers, P. W. (2009). Potentialphilicity and Potentialphobicity: Reactivity Indicators for External Potential Changes from Density Functional Reactivity Theory. J. Chem. Phys. 131, 114106. doi:10.1063/1.3231687

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., and Parr, R. G. (1997). Second-Order Density-Functional Description of Molecules and Chemical Changes. J. Chem. Phys. 106, 5578–5586. doi:10.1063/1.473580

CrossRef Full Text | Google Scholar

Liu, S. (2007). Steric Effect: A Quantitative Description from Density Functional Theory. J. Chem. Phys. 126, 244103. doi:10.1063/1.2747247

PubMed Abstract | CrossRef Full Text | Google Scholar

Malek, A., and Balawender, R. (2015). Revisiting the Chemical Reactivity Indices as the State Function Derivatives. The Role of Classical Chemical Hardness. J. Chem. Phys. 142, 054104. doi:10.1063/1.4906555

PubMed Abstract | CrossRef Full Text | Google Scholar

Melin, J., Aparicio, F., Galván, M., Fuentealba, P., and Contreras, R. (2003). Chemical Reactivity in the {N, NS, V(r)} Space. J. Phys. Chem. A 107, 3831–3835. doi:10.1021/jp034195j

CrossRef Full Text | Google Scholar

Miranda‐Quintana, R. A., and Smiatek, J. (2020). Theoretical Insights into Specific Ion Effects and Strong‐Weak Acid‐Base Rules for Ions in Solution: Deriving the Law of Matching Solvent Affinities from First Principles. ChemPhysChem 21, 2605–2617.

Google Scholar

Miranda-Quintana, R. A., Chattaraj, P. K., and Ayers, P. W. (2017a). Finite Temperature Grand Canonical Ensemble Study of the Minimum Electrophilicity Principle. J. Chem. Phys. 147, 124103. doi:10.1063/1.4996443

PubMed Abstract | CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A., Ayers, P. W., and Heidar‐Zadeh, F. (2021b). Reactivity and Charge Transfer beyond the Parabolic Model: the “|Δμ| Big Is Good” Principle. ChemistrySelect 6, 96–100. doi:10.1002/slct.202004055

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A., Ayers, P. W., and Heidar-Zadeh, F. (2021a). Well-normalized Charge-Transfer Models: a More General Derivation of the Hard/soft-Acid/base Principle. Theor. Chem. Acc. 140, 140. doi:10.1007/s00214-021-02840-y

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A., and Ayers, P. W. (2016a). Interpolation of Property-Values between Electron Numbers Is Inconsistent with Ensemble Averaging. J. Chem. Phys. 144, 244112. doi:10.1063/1.4953557

PubMed Abstract | CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A., and Ayers, P. W. (2016b). Systematic Treatment of Spin-Reactivity Indicators in Conceptual Density Functional Theory. Theor. Chem. Accounts 135, 1–18. doi:10.1007/s00214-016-1995-5

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2016a). Comments on “On the Non-integer Number of Particles in Molecular System Domains: Treatment and Description”. Theor. Chem. Accounts 135, 1–3. doi:10.1007/s00214-016-1945-2

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2016b). Condensed-to-atoms Hardness Kernel from the Response of Molecular Fragment Approach. Chem. Phys. Lett. 658, 328–330. doi:10.1016/j.cplett.2016.06.068

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2018). Density Functional Theory for Chemical Reactivity. Toronto: Apple Academic Press.

Google Scholar

Miranda-Quintana, R. A., Deswal, N., and Roy, R. K. (2022). Hammett Constants from Density Functional Calculations: Charge Transfer and Perturbations. Theor. Chem. Accounts 141, 1–10. doi:10.1007/s00214-021-02863-5

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A., Kim, T. D., Cárdenas, C., and Ayers, P. W. (2017b). The HSAB Principle from a Finite-Temperature Grand-Canonical Perspective. Theor. Chem. Acc. 136, 135. doi:10.1007/s00214-017-2167-y

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2017a). Note: The Minimum Electrophilicity and the Hard/soft Acid/base Principles. J. Chem. Phys. 146, 046101. doi:10.1063/1.4974987

PubMed Abstract | CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2017b). Perturbed Reactivity Descriptors: the Chemical Hardness. Theor. Chem. Accounts 136, 1–8. doi:10.1007/s00214-017-2109-8

CrossRef Full Text | Google Scholar

Miranda-Quintana, R. A. (2017c). Thermodynamic Electrophilicity. J. Chem. Phys. 146, 214113. doi:10.1063/1.4984611

PubMed Abstract | CrossRef Full Text | Google Scholar

Morell, C., Ayers, P. W., Grand, A., Gutiérrez-Oliva, S., and Toro-Labbé, A. (2008). Rationalization of Diels-Alder Reactions through the Use of the Dual Reactivity Descriptor Δf(r). Phys. Chem. Chem. Phys. 10, 7239–7246. doi:10.1039/b810343g

PubMed Abstract | CrossRef Full Text | Google Scholar

Morell, C., Grand, A., Gutierrez-Oliva, S., Toro-Labbe, A., and Toro-Labbe, A. (2007). Theoretical Aspects of Chemical Reactivity. Amsterdam: Elsevier, 31–45.

Google Scholar

Morell, C., Grand, A., Toro-Labbé, A., and Chermette, H. (2013). Is Hyper-Hardness More Chemically Relevant Than Expected? J. Mol. Model. 19, 2893–2900. doi:10.1007/s00894-013-1778-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Morell, C., Grand, A., and Toro-Labbé, A. (2005). New Dual Descriptor for Chemical Reactivity. J. Phys. Chem. A 109, 205–212. doi:10.1021/jp046577a

PubMed Abstract | CrossRef Full Text | Google Scholar

Morell, C., Grand, A., and Toro-Labbé, A. (2006). Theoretical Support for Using the Δf(r) Descriptor. Chem. Phys. Lett. 425, 342–346. doi:10.1016/j.cplett.2006.05.003

CrossRef Full Text | Google Scholar

Morokuma, K. (1971). Molecular Orbital Studies of Hydrogen Bonds. III. C=O···H-O Hydrogen Bond in H2CO···H2O and H2CO···2H2O. J. Chem. Phys. 55, 1236–1244. doi:10.1063/1.1676210

CrossRef Full Text | Google Scholar

Mortier, W. J. (1987). Electronegativity Equalization and its Applications. Struct. Bond. 66, 125–143.

Google Scholar

Mortier, W. J., Ghosh, S. K., and Shankar, S. (1986). Electronegativity-equalization Method for the Calculation of Atomic Charges in Molecules. J. Am. Chem. Soc. 108, 4315–4320. doi:10.1021/ja00275a013

CrossRef Full Text | Google Scholar

Mortier, W. J., Van Genechten, K., and Gasteiger, J. (1985). Electronegativity Equalization: Application and Parametrization. J. Am. Chem. Soc. 107, 829–835. doi:10.1021/ja00290a017

CrossRef Full Text | Google Scholar

Muñoz, M., and Cárdenas, C. (2017). How Predictive Could Alchemical Derivatives Be? Phys. Chem. Chem. Phys. 19, 16003–16012. doi:10.1039/c7cp02755a

PubMed Abstract | CrossRef Full Text | Google Scholar

Muñoz, M., Robles-Navarro, A., Fuentealba, P., and Cárdenas, C. (2020). Predicting Deprotonation Sites Using Alchemical Derivatives. J. Phys. Chem. A 124, 3754–3760.

PubMed Abstract | Google Scholar

Nagy, Á. (2007). Fisher Information and Steric Effect. Chem. Phys. Lett. 449, 212–215. doi:10.1016/j.cplett.2007.10.026

CrossRef Full Text | Google Scholar

Nalewajski, R. F., and Koniński, M. (1987). Density Polarization and Chemical Reactivity. Z.Naturforsch., A Phys.Sci. 42, 451–462. doi:10.1515/zna-1987-0506

CrossRef Full Text | Google Scholar

Nalewajski, R. F., and Parr, R. G. (1982). Legendre Transforms and Maxwell Relations in Density Functional Theory. J. Chem. Phys. 77, 399–407. doi:10.1063/1.443620

CrossRef Full Text | Google Scholar

Noorizadeh, S., and Shakerzadeh, E. (2008). A New Scale of Electronegativity Based on Electrophilicity Index. J. Phys. Chem. A 112, 3486–3491. doi:10.1021/jp709877h

PubMed Abstract | CrossRef Full Text | Google Scholar

Osorio, E., Ferraro, M. B., Oña, O. B., Cardenas, C., Fuentealba, P., and Tiznado, W. (2011). Assembling Small Silicon Clusters Using Criteria of Maximum Matching of the Fukui Functions. J. Chem. Theory Comput. 7, 3995–4001. doi:10.1021/ct200643z

PubMed Abstract | CrossRef Full Text | Google Scholar

Padmanabhan, J., Parthasarathi, R., Elango, M., Subramanian, V., Krishnamoorthy, B. S., Gutierrez-Oliva, S., et al. (2007). Multiphilic Descriptor for Chemical Reactivity and Selectivity. J. Phys. Chem. A 111, 9130–9138. doi:10.1021/jp0718909

PubMed Abstract | CrossRef Full Text | Google Scholar

Parr, R. G., and Bartolotti, L. J. (1982). On the Geometric Mean Principle for Electronegativity Equalization. J. Am. Chem. Soc. 104, 3801–3803. doi:10.1021/ja00378a004

CrossRef Full Text | Google Scholar

Parr, R. G., Donnelly, R. A., Levy, M., and Palke, W. E. (1978). Electronegativity: the Density Functional Viewpoint. J. Chem. Phys. 68, 3801–3807. doi:10.1063/1.436185

CrossRef Full Text | Google Scholar

Parr, R. G., and Pearson, R. G. (1983). Absolute Hardness: Companion Parameter to Absolute Electronegativity. J. Am. Chem. Soc. 105, 7512–7516. doi:10.1021/ja00364a005

CrossRef Full Text | Google Scholar

Parr, R. G., Szentpály, L. v., and Liu, S. (1999). Electrophilicity Index. J. Am. Chem. Soc. 121, 1922–1924. doi:10.1021/ja983494x

CrossRef Full Text | Google Scholar

Parr, R. G., and Yang, W. (1984). Density Functional Approach to the Frontier-Electron Theory of Chemical Reactivity. J. Am. Chem. Soc. 106, 4049–4050. doi:10.1021/ja00326a036

CrossRef Full Text | Google Scholar

Parr, R. G., and Yang, W. (1989). Density-Functional Theory of Atoms and Molecules. New York: Oxford UP.

Google Scholar

Pearson, R. G. (1997). Chemical Hardness. Weinheim, Germany: Wiley VCH.

Google Scholar

Perdew, J. P., Parr, R. G., Levy, M., and Balduz, J. L. (1982). Density-functional Theory for Fractional Particle Number: Derivative Discontinuities of the Energy. Phys. Rev. Lett. 49, 1691–1694. doi:10.1103/physrevlett.49.1691

CrossRef Full Text | Google Scholar

Pérez, P., Chamorro, E., and Ayers, P. W. (2008). Universal Mathematical Identities in Density Functional Theory: Results from Three Different Spin-Resolved Representations. J. Chem. Phys. 128, 204108. doi:10.1063/1.2916714

PubMed Abstract | CrossRef Full Text | Google Scholar

Polanco-Ramírez, C. A., Franco-Pérez, M., Carmona-Espíndola, J., Gázquez, J. L., and Ayers, P. W. (2017). Revisiting the Definition of Local Hardness and Hardness Kernel. Phys. Chem. Chem. Phys. 19, 12355–12364. doi:10.1039/c7cp00691h

PubMed Abstract | CrossRef Full Text | Google Scholar

Prodan, E., and Kohn, W. (2005). Nearsightedness of Electronic Matter. Proc. Natl. Acad. Sci. U.S.A. 102, 11635–11638. doi:10.1073/pnas.0505436102

PubMed Abstract | CrossRef Full Text | Google Scholar

Prodan, E. (2006). Nearsightedness of Electronic Matter in One Dimension. Phys. Rev. B 73, 085108. doi:10.1103/physrevb.73.085108

CrossRef Full Text | Google Scholar

Proft, F. D., Ayers, P. W., and Geerlings, P. (2014). “The Conceptual Density Functional Theory Perspective of Bonding,” in The Chemical Bond: Fundamental Aspects of Chemical Bonding. Editors S. Shaik, and G. Frenking (Darmstadt: Wiley), 233–270. doi:10.1002/9783527664696.ch7

CrossRef Full Text | Google Scholar

Rappe, A. K., and Goddard, W. A. (1991). Charge Equilibration for Molecular Dynamics Simulations. J. Phys. Chem. 95, 3358–3363. doi:10.1021/j100161a070

CrossRef Full Text | Google Scholar

Robles, A., Franco-Pérez, M., Gázquez, J. L., Cárdenas, C., and Fuentealba, P. (2018). Local Electrophilicity. J. Mol. Model. 24, 245. doi:10.1007/s00894-018-3785-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Roncero, O., De Lara-Castells, M. P., Villarreal, P., Flores, F., Ortega, J., Paniagua, M., et al. (2008). An Inversion Technique for the Calculation of Embedding Potentials. J. Chem. Phys. 129, 184104. doi:10.1063/1.3007987

PubMed Abstract | CrossRef Full Text | Google Scholar

Sablon, N., De Proft, F., and Geerlings, P. (2009). Reformulating the Woodward-Hoffmann Rules in a Conceptual Density Functional Theory Context: the Case of Sigmatropic Reactions. Croat. Chem. Acta 82, 157–164.

Google Scholar

Sablon, N., De Proft, F., and Geerlings, P. (2010). The Linear Response Kernel: Inductive and Resonance Effects Quantified. J. Phys. Chem. Lett. 1, 1228–1234. doi:10.1021/jz1002132

CrossRef Full Text | Google Scholar

Senet, P. (1996). Nonlinear Electronic Responses, Fukui Functions and Hardnesses as Functionals of the Ground‐state Electronic Density. J. Chem. Phys. 105, 6471–6489. doi:10.1063/1.472498

CrossRef Full Text | Google Scholar

Torrent-Sucarrat, M., Liu, S., and De Proft, F. (2009). Steric Effect: Partitioning in Atomic and Functional Group Contributions. J. Phys. Chem. A 113, 3698–3702. doi:10.1021/jp8096583

PubMed Abstract | CrossRef Full Text | Google Scholar

Torrent-Sucarrat, M., Luis, J. M., Duran, M., and Solà, M. (2005). An Assessment of a Simple Hardness Kernel Approximation for the Calculation of the Global Hardness in a Series of Lewis Acids and Bases. J. Mol. Struct. THEOCHEM 727, 139–148. doi:10.1016/j.theochem.2005.02.018

CrossRef Full Text | Google Scholar

Torrent-Sucarrat, M., Salvador, P., Geerlings, P., and Solà, M. (2007). On the Quality of the Hardness Kernel and the Fukui Function to Evaluate the Global Hardness. J. Comput. Chem. 28, 574–583. doi:10.1002/jcc.20535

PubMed Abstract | CrossRef Full Text | Google Scholar

Tsirelson, V. G., Stash, A. I., and Liu, S. (2010). Quantifying Steric Effect with Experimental Electron Density. J. Chem. Phys. 133, 114110. doi:10.1063/1.3492377

PubMed Abstract | CrossRef Full Text | Google Scholar

Ugur, I., De Vleeschouwer, F., Tuzun, N., Aviyente, V., Geerlings, P., Liu, S. B., et al. (2009). Cyclopolymerization Reactions of Diallyl Monomers: Exploring Electronic and Steric Effects Using DFT Reactivity Indices. J. Phys. Chem. A 113, 8704–8711.

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaidehi, N., Wesolowski, T. A., and Warshel, A. (1992). Quantum‐mechanical Calculations of Solvation Free Energies. A Combinedabinitiopseudopotential Free‐energy Perturbation Approach. J. Chem. Phys. 97, 4264–4271. doi:10.1063/1.463928

CrossRef Full Text | Google Scholar

Vargas, R., and Galván, M. (1996). On the Stability of Half-Filled Shells. J. Phys. Chem. 100, 14651–14654. doi:10.1021/jp9603086

CrossRef Full Text | Google Scholar

Vargas, R., Galván, M., and Vela, A. (1998). Singlet−Triplet Gaps and Spin Potentials. J. Phys. Chem. A 102, 3134–3140. doi:10.1021/jp972984t

CrossRef Full Text | Google Scholar

Verstraelen, T., Bultinck, P., Van Speybroeck, V., Ayers, P. W., Van Neck, D., and Waroquier, M. (2011). The Significance of Parameters in Charge Equilibration Models. J. Chem. Theory Comput. 7, 1750–1764. doi:10.1021/ct200006e

PubMed Abstract | CrossRef Full Text | Google Scholar

Verstraelen, T., Van Speybroeck, V., and Waroquier, M. (2009). The Electronegativity Equalization Method and the Split Charge Equilibration Applied to Organic Systems: Parametrization, Validation, and Comparison. J. Chem. Phys. 131, 044127. doi:10.1063/1.3187034

PubMed Abstract | CrossRef Full Text | Google Scholar

Wesolowski, T. A., and Leszczynski, J. (2006). “One-electron Equations for Embedded Electron Density: Challenge for Theory and Practical Payoffs in Multi-Scale Modelling of Complex Polyatomic Molecules,” in Computational Chemistry: Reviews of Current Trends (Singapore: World Scientific), 1–82.

Google Scholar

Wesolowski, T. A. (2004). Quantum Chemistry 'without Orbitals' - an Old Idea and Recent Developments. Chimia 58, 311–315. doi:10.2533/000942904777677885

CrossRef Full Text | Google Scholar

Wesolowski, T. A., and Warshel, A. (1993). Frozen Density Functional Approach for Ab Initio Calculations of Solvated Molecules. J. Phys. Chem. 97, 8050–8053. doi:10.1021/j100132a040

CrossRef Full Text | Google Scholar

Wu, Q., Ayers, P. W., and Zhang, Y. (2009). Density-based Energy Decomposition Analysis for Intermolecular Interactions with Variationally Determined Intermediate State Energies. J. Chem. Phys. 131, 164112. doi:10.1063/1.3253797

PubMed Abstract | CrossRef Full Text | Google Scholar

Yañez, O., Báez-Grez, R., Inostroza, D., Pino-Rios, R., Rabanal-León, W. A., Contreras-García, J., et al. (2021). Kick–Fukui: A Fukui Function-Guided Method for Molecular Structure Prediction. J. Chem. Inf. Model. 61, 3955–3963.

PubMed Abstract | Google Scholar

Yang, W., and Mortier, W. J. (1986). The Use of Global and Local Molecular Parameters for the Analysis of the Gas-phase Basicity of Amines. J. Am. Chem. Soc. 108, 5708–5711. doi:10.1021/ja00279a008

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, W., Parr, R. G., and Pucci, R. (1984). Electron Density, Kohn-Sham Frontier Orbitals, and Fukui Functions. J. Chem. Phys. 81, 2862–2863. doi:10.1063/1.447964

CrossRef Full Text | Google Scholar

Yang, W., Zhang, Y., and Ayers, P. W. (2000). Degenerate Ground States and a Fractional Number of Electrons in Density and Reduced Density Matrix Functional Theory. Phys. Rev. Lett. 84, 5172–5175. doi:10.1103/physrevlett.84.5172

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, X. D., Patel, A. H. G., Miranda-Quintana, R. A., Heidar-Zadeh, F., González-Espinoza, C. E., and Ayers, P. W. (2016). Communication: Two Types of Flat-Planes Conditions in Density Functional Theory. J. Chem. Phys. 145, 031102. doi:10.1063/1.4958636

PubMed Abstract | CrossRef Full Text | Google Scholar

Ziegler, T., and Rauk, A. (1979). A Theoretical Study of the Ethylene-Metal Bond in Complexes between Copper(1+), Silver(1+), Gold(1+), Platinum(0) or Platinum(2+) and Ethylene, Based on the Hartree-Fock-Slater Transition-State Method. Inorg. Chem. 18, 1558–1565. doi:10.1021/ic50196a034

CrossRef Full Text | Google Scholar

Ziegler, T., and Rauk, A. (1977). On the Calculation of Bonding Energies by the Hartree Fock Slater Method. Theor. Chim. Acta 46, 1–10. doi:10.1007/bf02401406

CrossRef Full Text | Google Scholar

Zielinski, F., Tognetti, V., and Joubert, L. (2012). Condensed Descriptors for Reactivity: A Methodological Study. Chem. Phys. Lett. 527, 67–72. doi:10.1016/j.cplett.2012.01.011

CrossRef Full Text | Google Scholar

Keywords: chemical reactivity, density functional theory, molecular Interaction analysis, conceptual chemistry, response function

Citation: Miranda-Quintana RA, Heidar-Zadeh F, Fias S, Chapman AEA, Liu S, Morell C, Gómez T, Cárdenas C and Ayers PW (2022) Molecular Interactions From the Density Functional Theory for Chemical Reactivity: The Interaction Energy Between Two-Reagents. Front. Chem. 10:906674. doi: 10.3389/fchem.2022.906674

Received: 28 March 2022; Accepted: 19 April 2022;
Published: 13 June 2022.

Edited by:

Juan Frau, University of the Balearic Islands, Spain

Reviewed by:

Eduard Matito, Donostia International Physics Center (DIPC), Spain
Ruby Srivastava, Centre for Cellular & Molecular Biology (CCMB), India

Copyright © 2022 Miranda-Quintana, Heidar-Zadeh, Fias, Chapman, Liu, Morell, Gómez, Cárdenas and Ayers. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Ramón Alain Miranda-Quintana, quintana@chem.ufl.edu; Carlos Cárdenas, cardena@uchile.cl; Paul W. Ayers, payers@mcmaster.ca; Tatiana Gómez, tatiana.gomez@uautonoma.cl

Present Address: Stijn Fias, FoodPairing, Ghent, Belgium

Download