Home Physical Sciences Quantum chemistry of molecules in solution. A brief historical perspective
Article Publicly Available

Quantum chemistry of molecules in solution. A brief historical perspective

  • Manuel F. Ruiz-López ORCID logo EMAIL logo
Published/Copyright: May 23, 2025

Abstract

Calculating the wave function and the properties of a molecule in solution is a common procedure today that can be performed with a variety of methods. However, the study of solvated systems did not begin with the establishment of the Schrodinger equation in 1926 and the development of the first methods in computational chemistry. In fact, it took until 1973 to see the first self-consistent field equations that included the electrostatic potential of the solvent. It took another two decades for computational methods and solvent models to become consolidated, with an explosion of applications in the 1990s. This article briefly describes that fascinating story.

Introduction

The United Nations has proclaimed 2025 the International Year of Quantum Science and Technology. It celebrates the centenary of the birth of quantum mechanics 1 and the publication of the seminal papers by Werner Heisenberg 2 in Zeitschrift für Physik in 1925, and Erwin Schrödinger 3 , 4 in Annalen der Physik in the first months of 1926.

In the years that followed, the development of methods for solving Schrödinger’s equation for electronic motions in atoms and molecules was incredibly rapid, thanks to the contributions of eminent scientists such as Dirac, Heitler, London, Hund, Slater, Hartree, Fock, Mulliken, Hückel, Pauling, and many others. New methods soon emerged that would establish the physical foundations of modern theoretical chemistry and would have a profound impact on chemical science.

A few years later, in the 1950s and 1960s, the availability of computers led to a revolution in quantum chemistry (as in many other fields of science and technology), opening the way to the field of computational chemistry. 5 Of course, the limited capacity of computers meant that incipient electronic structure methods could only be applied to systems containing a small number of electrons. As a result, the vast majority of studies initially focused on describing isolated atoms or molecules containing a few atoms. The study of large molecules or complex systems such as liquids or solutions was not contemplated. It was in the 1970s, some 50 years after the advent of quantum mechanics, that theoretical chemists developed the first polarizable continuum solvation models and began to apply them to the study of the electronic structure of molecules in solution. It was not until the 1980s and especially the 1990s that realistic microscopic models and first-principles simulations began to take shape. Figure 1 depicts the main categories of solvation models.

Fig. 1: 
Scheme presenting the main types of solvation models discussed in the manuscript: supermolecule model (1), continuum model (2), and microscopic models combining quantum chemical methods and statistical mechanics simulations. In this case, the whole solute-solvent system can be described quantum mechanically (3), or using hybrid quantum/classical approaches (4).
Fig. 1:

Scheme presenting the main types of solvation models discussed in the manuscript: supermolecule model (1), continuum model (2), and microscopic models combining quantum chemical methods and statistical mechanics simulations. In this case, the whole solute-solvent system can be described quantum mechanically (3), or using hybrid quantum/classical approaches (4).

This article describes the main stages of this exciting journey in a fairly compact, formula-free way. It is not an in-depth account of solvation models. There are many exhaustive reviews in the literature, although generally, each review has focused on a particular class of models. For the interested readers, review references will be provided below. The aim here is to bring together in one place a summary of the main steps that led to the development by theoretical chemists of various models for the study of molecules in solution using quantum chemical methods. Thus, I have taken little, or no account of approaches developed after the late 1990s, giving priority to the birth of methods and the context in which they were developed. Solvation models underwent many interesting developments after 2000, of course, but I regard them as primarily inspired by the models described below. Furthermore, I will not discuss the development of statistical mechanics simulations using Monte Carlo or molecular dynamics techniques, nor that of effective intermolecular interaction potentials. Only statistical simulations incorporating a quantum-mechanical description of the solute, or solute-solvent system will be considered. The available solvent models can be classified according to various criteria, but here I have chosen to present them chronologically, even if in some cases this has obvious limitations, as different models or versions of models were developed in parallel. I have sought to document the facts reported by systematically providing the original bibliographical references, although inevitably parts of the text below are also the reflect of some personal memories.

Simple models: the supermolecule approach

The first attempts to describe the influence of a solvent on the molecular properties of a solute using quantum chemical methods focused on the effect of hydrogen-bonds. This type of intermolecular interaction has received a great deal of attention since the beginning of the twentieth century as a possible cause of certain observed irregularities. A nice example of this was the work by Moore and Winmill 6 on the dissociation constants of amines, in particular of ammonia and methylamines (I will come back to this issue below). The first detailed description was probably done in 1920 by Latimer and Rodebush, 7 who were studying the structure and properties of water, and proposed a Lewis-type formalism for the interactions between two water molecules in which a hydrogen nucleus is “held between two octets”. Nevertheless, it was the seminal book by Pauling which popularized the concept and provided the first definition of hydrogen bonds. A brief history of the hydrogen-bond can be found in the paper of Smith, 8 and also in papers by Bratoz, 9 Kollman, 10 and Szalewicz, 11 for instance.

Theoretical studies of hydrogen bonded systems using electronic theories were carried out by several authors starting in the 1950’s. The works by Pimentel 12 in 1951 and Nukasawa 13 in 1953 were probably the first attempts done in this direction, and the works by Lennard-Jones, 14 Pople, 15 and Coulson 16 contributed also to improve the understanding of the hydrogen bond interaction. In the 1960s and early 1970s, a large number of works using SCF theories began to be published. Some reference works of this period are for instance the papers by Clementi (1967), 17 which studied the NH3 + HCl reaction (cited by Kollman 10 as the first ab initio study of an H-bonded system), Morokuma (1968), 18 who studied the H2O dimer, Del Bene (1972), 19 who described the dimers of H2O2 and H2O, and Alagona et al. (1973), 20 devoted to formamide-water complexes. The vast literature covering the mentioned years was compiled in various books and articles, which can be consulted for more details on the progress of hydrogen bonding theory during this period. Let us cite in particular the 1960 book “The Hydrogen Bond” by Pimentel and McClellan, 21 the 1967 review paper by Bratoz, 9 and the review by Kollman in 1972. 10

Thus, gradually, at the end of the 1950s, the idea of performing electronic calculations on molecular edifices comprising a solute molecule interacting with several surrounding molecules through hydrogen bonds emerged as a general approach to considering the effects of the environment on the structure and other properties of molecules in solution. This led to the establishment of the “supermolecule method” which was popularized and strongly developed in the 1960s and 1970s by Alberte and Bernard Pullman. The Pullmans had been working on the applications of quantum theories to biology 22 , 23 from the late 1950s and 1960s, and indeed they are recognized as the founders of the field of quantum biochemistry. Since biochemical processes take place mainly in aqueous media, accounting for biomolecule-water interactions is essential to understanding the mechanisms involved, and the main type of interaction to be considered is of course hydrogen bonding. Thus, the supermolecule (or supermolecular) approach was initially implemented by the Pullmans to elucidate the structure of water around the main chemical constituents of living organisms, amino acids, peptides, DNA base pairs, etc., and to clarify how hydration affects their properties. 24 , 25 The various features of the supermolecule method, possible calculation strategies and numerous case studies were detailed in a review published by A. Pullman in 1976, 26 where she gave the following definition of the method: “It has become customary in the recent past to call supermolecule or supersystem, an entity made of two (or more) molecules in interaction, when they are treated as a single system by the methods of quantum chemistry, computing the global wave function of the whole set of electrons of the entire system in the field of all the nuclei fixed in the appropriate mutual relative arrangement”.

The model has been widely used until today in many different versions, sometimes being referred as the “discrete” or “explicit” solvation model, in opposition to the so-called “continuum” or “implicit” solvation models described below. Note that the supermolecule technique has been used to study different aspects of intermolecular interactions, not necessarily related to solvation phenomena, though of course this is the focus here. However, in practice, the supermolecule model is useful mainly when there are specific (sufficiently strong) interactions between the solute and the solvent and is not recommended for the study of solutes in “inert” solvents such as hydrocarbons. Another drawback of the model is the small number of solvent molecules that can be handled in practice, as the computational time in quantum chemistry calculations increases very rapidly with the size of the system. In addition, with an increasing number of molecules, a statistical treatment may be mandatory, since many geometrical arrangements with similar energy may occur. Therefore, usually only solvent molecules that interact directly with the solute via hydrogen bonds or other specific interactions are taken into account, thus neglecting the potentially important effect of long-range solute-solvent electrostatic interactions.

Despite the above limitations, the model can be useful to obtain some main trends, for example, on the structure of the first solvation shell or on the solvent effect on the solute properties. The case of the irregular basicity of the ammonia and methylamines molecules in the experiments by Moore and Winmill 6 mentioned above is a simple and nice example illustrating this principle. Thus, by the end of the 1970s, some theoretical studies had already been carried out on the proton affinities of amines in the gas phase, 27 , 28 , 29 , 30 , 31 , 32 as well as on the solvation energies of ammonium ions (specially by the Pullmans). 33 , 34 In this context, using the supermolecule approach and basic semiempirical quantum chemical calculations, we were able to elucidate the origin of the basicity irregularity in aqueous solution experimentally observed, 35 , 36 which results of an opposite trend between the gas phase proton affinity and the amine/ammonium solvation energy difference term. The results provided only qualitative trends and pointed out the necessity to perform higher level ab initio calculations that take into account long-range solvation effects to obtain more accurate data.

It was also in the 1970s that the first dielectric continuum models began to emerge. The birthplace of the developments of this method was the University of Nancy, in France, and the context in which they were performed is outlined below.

The birth of dielectric continuum models

The continuum solvation model has completely revolutionized our understanding of the effects of solvation on chemical processes. During the last decades, this model has become a routine technique and is implemented in most of the available quantum chemical programs (Gaussian, 37 GAMESS, 38 Q-Chem, 39 MOLCAS, 40 etc.). There are probably thousands of publications using it every year. The principles of the model are extremely simple, as is its use, what has certainly contributed to its rapid and broad popularization.

The continuum model is based on the dielectric theory of liquids, which goes back to the works by Mossotti, 41 Clausius, 42 Debye, 43 Born, 44 Kirkwood, 45 and Onsager, 46 among others. The book of Böttcher 47 “The Theory of Electric Polarization” is a classical reference in the field that offers an excellent treatment of the subject. In the classical electrostatics model, the solvent is represented as a structureless polarizable continuum medium characterized by its macroscopic dielectric properties, and the solute is placed in a cavity created in the dielectric continuum. As a result, there is a mutual polarization: the dielectric is polarized due to the charge distribution (electrons and nuclei) of the solute, and the latter is polarized in turn by the electrostatic potential (the so-called reaction field) created by the polarized dielectric continuum. The book of Böttcher 47 summarizes the electrostatic equations for a number of simple cases. For instance, the solute can be represented by a non-polarizable or a polarizable point dipole, or by a series of multipole moments, and the cavity shape can be spherical, spheroidal or ellipsoidal.

In Nancy, Jean Barriol, who had held the first chair of Theoretical Chemistry in France since 1948 (for a Barriol’s biography, see Ref. 48]), was interested on the dielectric behavior and physical properties of liquid mixtures, making key contributions in the field. 49 , 50 , 51 Although Barriol was a theoretician, he placed primary importance on experimental work, and indeed, students in the laboratory always started their training by doing experiments. 48 One of his students, Jean-Louis Rivail moved to Coulson’s lab in Oxford for a post-doctoral stay in 1967 after completing his doctorate in Nancy. Rivail wanted to complete his training in quantum chemistry and his work in Oxford dealt with computational methods for predicting molecular electrostatic properties and H-bonding in σ-electron systems. Rivail came back to Barriol’s lab the same year thanks to a Maître de Conférence position and introduced a new line of research based on a theoretical approach to intermolecular interactions involving the use of quantum chemical methods. He supervised the PhD work of Daniel Rinaldi, who had joined the lab from INSA Lyon interested by his reading of Barriol’s book on quantum mechanics.

The PhD project of Rinaldi was devoted to the development of the computational code GEOMO 52 , 53 for quantum chemical calculations using the most popular semi-empirical methods of the time (CNDO, 54 INDO, 55 MINDO 56 ), which in addition included, for the first time, an algorithm to optimize molecular geometries. 57 Obviously, while working in Barriol’s laboratory, the idea of including the role of the environment in quantum chemical calculations of molecules in solution, using a dielectric continuum model to represent the solvent effect, came naturally to Rivail and Rinaldi. The aim 58 was motivated by NMR 59 and microwave 60 experiments done in the laboratory showing that simple molecules such as allylamine had a very different structure in solution and in the gas phase, confirming that solvents effects on molecular properties may be very significant. 61 They incorporated the solute-solvent electrostatic interaction into the solute Hamiltonian to obtain the solute wave function in the presence of the solvent. The mutual solute-solvent polarization was accounted for through the self-consistent-field (SCF) equations of the Hartree–Fock method, after adding to the standard equations the interaction of the solute with the solvent reaction field. Nowadays, these equations are usually called the self-consistent reaction-field (SCRF) equations. The reaction field depends on the shape of cavity, the solvent dielectric constant, and the solute charge distribution. There are analytical expressions for the potential when the cavity has regular shapes (e.g. spherical, spheroidal or ellipsoidal), which take the form of a multipolar expansion centered in a given point (or set of points), and logically, the former implementations of the model assumed such hypothesis. In the very first application of the continuum model published in 1973, 62 Rivail and Rinaldi studied the properties (geometries, dipole moments, polarizabilities) of the water molecule and of different isomers of the water dimer using a spheroidal cavity, whose ellipticity was defined from the polarizability tensor. The interaction was limited to the first term of the expansion, i.e. the dipole moment, and the calculations were carried out with the CNDO/2 method. It must be emphasized that in the SCF equations of this first paper, electrostatic and dispersion energies were included, 62 although the influence of the dispersion term was found to play a minor role and will be omitted in future versions of the model.

Soon after this pioneer work, in 1976, the same authors published an extension of the model implementing now the full multipole moment expansion of the solute-solvent interaction, assuming for simplicity a spherical cavity shape. 63 The calculations were still done at the CNDO/2 semiempirical level and dealt with the study of different conformational equilibria in solution.

The two papers by Rivail and Rinaldi above marked the birth of continuum models in quantum chemistry, although it is worth mentioning that similar initiatives were undertaken during the same period by Yomosa, 64 Johnson, 65 Claverie, 66 , 67 Tapia, 68 , 69 and others. 70 , 71 , 72 Coulson himself had been interested in the effect of a dielectric continuum on electronic energy levels as early as 1958, 73 but the systems considered were too simple (e.g. the hydrogen atom). 74 , 75 Nevertheless, it was the approach developed in Nancy that brought definitive progress in the field and paved the way for the methods we use today. If the 1970s saw the birth of the model, the 1980s saw it grow and develop spectacularly, reaching the somewhat definitive form in various versions and implementations in the 1990s.

Expansion of continuum models

Developments carried out in Nancy

During the 1980s, the development of the continuum model took place essentially in Nancy and Pisa. Let me first to follow up on the developments carried out in Nancy, in which I was strongly involved starting from 1980. I had moved from Madrid to Nancy after defending my Master Thesis at the Autonomous University of Madrid, which was supervised by Juan Bertrán and Antonio Oliva. Bertrán had spent some time in Nancy to explore the capabilities of the continuum model to study solvation phenomena and the result was a fruitful and long-term collaboration 76 between the two groups. When I joined Rivail’s group, one of the most important challenges I faced during my PhD project was to extend the model to ab initio computations. The first version was implemented in the Gaussian 76 program 77 and incorporated a full multipole moment development of the reaction field using general ellipsoidal molecular cavities. The work was presented at the 4th International Congress of Quantum Chemistry (Uppsala, 1982) 78 and was published in 1983. 79 My PhD project 80 did also focus on applications of the model to the calculation of molecular spectroscopic properties in different domains, IR, 80 NMR (spin-spin nuclear 81 and nuclear quadrupole 82 coupling constants), and UV–Vis. 83 , 84 In this latter case, a key achievement was the development of a perturbational theory 84 for the calculation of electronic excited states of solvated molecules in the Random Phase Approximation (RPA) and the implementation of the corresponding code interfacing our SCRF program and the RPA program of Flament. 85 The role of vibronic coupling on UV–Vis spectra was also considered. 86

After my PhD, and now as a permanent CNRS researcher in the same laboratory, I turned to the formidable and exciting challenge of studying chemical reactions in solution. The SCRF method had already been applied to study solvent effects on some chemical reactions 87 , 88 , 89 but only in a qualitative way. To move further, it was compulsory to be able to optimize the geometrical structure of solvated molecules and to compute vibrational frequencies in order to locate transition states and calculate free energies accurately. This became possible in 1991 thanks to the development of highly efficient analytical optimization algorithms by Rinaldi. 90 , 91 , 92 Indeed, the use of a multipole moment development to represent the reaction-field was a major advantage because it allows to obtain analytical expressions for the energy derivatives. The first application of the model with full geometry optimization was devoted to the structure of the water dimer in liquid water and the analysis of cooperative effects. 93 In 1993–1994, the studies of chemical reactions began to take shape. They focused on isomerization, 94 Diels-Alder, 95 and hydrolysis 96 processes, illustrating the key role solvation may have on their mechanism, chemical selectivity, kinetics and thermodynamics. Two other key extensions of the Nancy continuum model in the 1990s were the implementation into the deMon program, 97 allowing the use of density functional theory (DFT) methods, which were not yet widely employed by theoretical chemists, and the development of algorithms to deal with general cavity shapes 98 and distributed multipole moment expansions. 99

The influence of dispersion forces on the electronic structure of the solute was examined by Costa Cabral after adapting the SCRF equations. 100 , 101 The study showed negligible or slight modifications (compared to the electrostatic effect) in the case of polar solutes but non-negligible effects in the case of apolar ones. A parameterized non-electrostatic solvation term including the dispersion energy, as well as cavitation and other contributions, was introduced in the Nancy model a few years later. 102 Costa Cabral also carried out Monte Carlo simulations with a dipolar hard-sphere solvent model and solutes of various types (dipolar, quadrupolar, multipolar) to discuss some of the approximations done (cavity volume/shape) in the continuum model. 103

I have mentioned that from the beginning we had interfaced our SCRF package with the Gaussian program. In August 1990, I presented the model in an invited talk at a NATO Conference in Vimeiro, Portugal, which several members of the Gaussian development team had attended. The title of my talk was “The Self-Consistent Reaction Field model for molecular computations in solution”. 104 We had interesting discussions on the model, and the Vimeiro meeting arguably stimulated the development of the SCRF method in the official version of Gaussian 90, with the first publication appearing in 1991 105 (the initial model assumed the simplest Onsager approximation, i. e. a dipole moment in a spherical cavity).

It is interesting to note that in the period considered above, during which only a reduced number of theoretical groups in the world have started to use continuum models and to study solvent effects, the group of Nancy (and to some extent also the group of Pisa, see below) established very fruitful collaborations with several labs in Spanish Universities. I would like to mention a few explicitly because, in my opinion, they had a decisive impact on spreading the popularity of the continuum model at the time. Most of those collaborations involved the study of reaction mechanisms in solution. I have already mentioned the long-term collaboration with Bertrán and Oliva (first in Madrid, then in Barcelona), 81 , 87 , 88 , 93 , 102 , 106 , 107 but very fruitful collaborations were also carried out with the Universities of Seville (E. Sánchez-Marcos and R. R. Pappalardo; Pappalardo was involved on the program developments during his post-doctoral stay in Nancy), 89 , 92 , 94 , 106 , 108 , 109 and Oviedo (T. Sordo, R. López). 110 , 111 , 112 , 113 , 114 , 115 The collaboration with the organic chemistry group of the University of Zaragoza (J.A. Mayoral, J.I. García, L. Salvatella) was also key to check the reliability of the model. 95 , 116 , 117 , 118 , 119 , 120

The development of the continuum model was gradually abandoned in our group for a variety of reasons, the most relevant of which was probably the shift in the 1990s to more elaborate theoretical approaches, as the combined quantum/classical simulations (see below). Nevertheless, the model continued to be used by several groups for some time. Indeed, the use of multipole expansions to represent the reaction field had several advantages with respect to other models being developed in parallel. First, it offered a straightforward physical interpretation of the calculated effects. Second the computations were quite efficient and fast since the SCRF equations involved analytical expressions in terms of spherical harmonics. Of course, when the cavity shape has a non-uniform form, some numerical computations are needed, but they only involve coefficients (the reaction-field factors) that depend on the cavity surface and the solvent dielectric constant (not on the solute electronic density); therefore, they are not recalculated in the SCRF iterations, which saves computing time. Finally, energy derivatives are easily calculated, simplifying geometry optimization and vibrational frequency calculations. In the literature, it has sometimes been said that one of the shortcomings of the model is the poor convergence of the multipole expansion. While this is partly true, in practice, computational issues are generally linked to the amount of electron density external to the cavity. This is a general problem in quantum chemical continuum models 121 and has been treated in different ways by different groups. It can be addressed in the case of the Nancy model too by using corrected multipolar expansions corresponding to a generalized charge distribution inside the cavity that fits the “exact” solute electrostatic potential outside the cavity. A detailed review on the theory and applications of the Nancy continuum model can be found in Ref. 122].

The polarizable continuum model developed in Pisa

As mentioned above, the other cornerstone of the continuum model in the early years was the Pisa group. The version of the model developed by Jacopo Tomasi and collaborators in Pisa 123 has its origins on the work done in the group on molecular aggregates in the early seventies at the ab initio level. 124 Though the Pisa and Nancy continuum models share the same physical foundations, they differ by the mathematical approach used to obtain the reaction field, i. e. the electrostatic potential interacting with the solute. While in the Nancy model, expressions based on multipolar expansions were used to develop the potential of the polarized dielectric medium in a point or a set of points, in the Pisa model, the potential was obtained from a discrete set of virtual charges on the cavity surface, whose values are derived by numerically solving the Poisson-Laplace equations with the appropriate boundary conditions. The two approaches are equivalent in the classical electrostatics model, as described in the book by Böttcher, 47 but present different features in their implementation for quantum chemistry calculations.

The Polarizable Continuum model (PCM) first developed at Pisa is certainly the most broadly used continuum solvation model nowadays and it has been the subject of many methodological improvements along the years. Some of the most important milestones achieved on PCM until about the year 2000 are the following. Dispersion and repulsion contributions to the free energy were introduced at the end of the 1980s. 125 , 126 Adaptation to the study of non-equilibrium phenomena began in 1992. 127 , 128 In 1996 the model was implemented in the Gaussian 94 program, 129 which marked the beginning of the collaboration of the Pisa group with Gaussian Inc, and at the same time the start of the methodological work on PCM carried out by the Naples group. 130 In 1997, problems related to the effect of the solute charge lying out of the cavity were addressed using a classical electrostatics approach. 131 In 1997–1998, two new formalisms of the PCM model were proposed. The integral equation formalism (IEF) 132 , 133 , 134 appeared to be computationally faster and more stable than the standard one and could be used with anisotropic dielectrics. This formalism was shown, in addition, to provide an accurate treatment of the issues related to the solute electronic charge outside the cavity. 121 The conductor-like PCM model (CPCM) 135 was developed by the Barone group at Naples, and was based on the continuum metallic model of Klamt. 136 Finally, in 1998–1999, the development of analytical first 137 and second 138 derivatives of the energy with respect to nuclear coordinates opened new perspectives for applications in chemical reactivity and spectroscopy.

As mentioned above for Nancy, in the early years of the continuum model, several collaborations with Spanish groups contributed to its quick development. In the case of PCM, one can stress the collaborations with the University of Badajoz (F. Olivares del Valle, M. Aguilar) on molecular spectroscopy 139 and non-equilibrium phenomena, 140 , 141 the University of Valencia (E. Silla, J.L. Pascual-Ahuir) on the development of the GEPOL algorithm to deal with molecular surfaces, 142 , 143 , 144 , 145 and the University of Barcelona (J. Luque, M. Orozco), who developed a home version of PCM (known as MST) with several methodological innovations. 146 , 147 , 148 , 149 , 150

For further details on the PCM approach, the reader may consult the many reviews published in the past by the authors, in particular those of Tomasi in 2011 151 and Mennucci in 2012. 152 Another very detailed historical perspective of the model has been written by Tomasi and is available at the webpage of the Dipartimento di Chimica e Chimica Industriale of the University of Pisa. 153 Other general reviews on PCM and continuum models will be listed below, at the end of Section “Expansion of Continuum Models”.

Other continuum models

Some other approaches using continuum models were proposed in the 1980s and 1990s, such as those reported by Karelson et al., 154 Mikkelsen et al., 155 Karlström et al., 156 van Duijnen et al., 157 or Chipman. 158 But those developed by Cramer and Truhlar 159 and by Klamt 136 have occupied a special place in subsequent literature, and deserves special mention here.

The COSMO (conductor-like screening model) model of Klamt, 136 like PCM, uses apparent surface charges to calculate the reaction-field, though it uses a different technique to solve the electrostatic problem. The method first treats the solvent as it was a hypothetical conductor medium. In practice, this means that the dielectric constant of the solvent is changed to infinity, which simplifies the electrostatic equations. Afterwards, an empirical correction is applied to the results to recover the dielectric behavior of the solvent under consideration. The recent extension of the model to COSMO-RS (COSMO for realistic solvation) 160 , 161 relays on a statistical thermodynamics approach. It can be used to predict various thermodynamic properties of solvents and solvent mixtures at different temperatures and has become a widely used method within the chemical engineering community.

The philosophy of the solvation models proposed by Cramer and Truhlar in Minnesota differed significantly from the previous ones. In the different versions of the SMx approach (Solvation Model), 159 , 162 the solvation free energies are obtained using a generalized Born (GB) approximation. In the GB approximation, the solute charge distribution is reduced to a series of partial atomic charges, and the interaction with the continuum takes the form of a generalized Born expression in which some parameters are introduced. The way of obtaining the atomic charges and the parameters depend on SMx version (see the original references). More recently, 163 the authors have developed a different implementation of the method called SMD, where the electrostatic contribution is computed in terms of the integral equation formalism. The D in the SMD acronym stands for density and stresses the fact that the full solute electron density is used, and not just partial atomic charges as in SMx models.

Continuum model final remarks

The simple continuum model is still widely used in the literature because it allows fast calculations of solvated species that are not much more expensive than the equivalent calculations for isolated systems. Over the past decades, different versions of the continuum model have provided invaluable insights into the effects of solute-solvent interactions on spectroscopy, chemical reactivity, and molecular properties in general. Each model has its advantages and drawbacks, and the choice is often a question of taste. Indeed, the electrostatic components of the solvation energy calculated through the self-consistent reaction field equations are similar when the different methods are used in a consistent way. This was shown for instance in Ref. 164] for the Nancy and PCM models, and in Ref. 165] for the Nancy, PCM (MST) and SM5 models. All models contemplate a way to obtain the non-electrostatic component, which includes the cavitation energy (the energy required to create an empty volume in the continuum), the solute-solvent dispersion-repulsion energy, and possibly other contributions not included in the preceding ones. Though there are differences between models, the non-electrostatic component is generally assigned a certain proportionality with surface area and parameterized to reproduce some reference data. The definition of the cavities (shape and volume) has also been a central question in this field, with assumptions going from regular surfaces to cavity surfaces based on the superposition of nuclear-centered spheres, solvent excluding or solvent accessible surfaces, isodensity or isopotential surfaces, etc. For the volume, early works 80 , 166 used the proportionality existing between liquid densities and the van der Waals molecular volume (i.e. the volume enclosed by atomic spheres with Bondi radii), in consistency with Onsager’s recipe. Later, optimized atomic radii were derived allowing to improve the agreement between experimental and theoretical solvation energies (see for instance 167 , 168 , 169 and references cited therein).

The continuum model for describing solvated molecules with quantum chemical methods continue to be developed after half a century of history. Together with J.-L.Rivail, 58 I recently published a short article to celebrate its 50th anniversary. Years before, Tomasi had written a remarkable historical perspective 170 on the occasion of the 30th anniversary of the model 62 and the congress I organized in Nancy for the official retirement of J.-L. Rivail (a satellite meeting of the 11th International Congress of Quantum Chemistry). To finish this section, let me cite some reviews which can be consulted by readers wishing to get further details on the mathematical formulation of the model and on its applications in different domains:

  1. Tomasi et al. (1994), 171 (2005), 169 (2007), 172 and (2011). 151

  2. Cramer & Truhlar (1995), 173 (1999), 174 (2002) 175 and (2008). 162

  3. Mennucci (2010), 176 (2012). 152

  4. Ruiz-López (2008). 122

Beyond the continuum: moving towards microscopic models

The approximations inherent to continuum models are clearly too crude and, in many cases, a realistic modeling of the system is not possible without considering the microscopic structure of the solvent. Naturally, handling systems containing a very large number of molecules requires recourse to statistical mechanics. The combination of quantum chemistry methods with statistical mechanics simulations was an ambition that had long been in the air to study solvation phenomena, but it was not until the 1980s, with the increase of computer performances, that it became a reality in various forms. Rivail 177 used the expression “real molecules in model liquids” to refer to the quantum chemical continuum model, by contrast with classical models using statistical simulations with effective interaction potentials, referred to as “real liquids of model molecules”. Following the same terminology, models using statistical simulations coupled with a quantum chemical description of the full solute-solvent system could be described as “real molecules in real liquids”. These types of approaches will be presented in the next section, but the road to them will be paved by the development of several methods coined in parallel to the continuum model, which are outlined now.

The limitations of the supermolecule and the continuum were partially solved through the combination of the two approaches, that is by carrying out quantum chemical calculations for a solute-solvent cluster embedded in a dielectric continuum. The cluster contains the solute plus one or more solvation shells, and hence the possible role of specific solute-solvent interactions is explicitly treated. The continuum, on the other hand, allows taking into account the long-ranged electrostatic contribution from solvent effects beyond the explicit solvent shell(s), which is often a fundamental term. This intuitive scheme is known as the discrete-continuum model. It was used from the very beginning of the continuum model. Indeed, in the first paper by Rinaldi and Rivail in 1973, 62 they already studied a supermolecular system, the water dimer. However, the first study using the discrete-continuum model in the form given by the definition above was reported by Claverie et al. in 1978. 34 The authors made a detailed comparison of the calculated solvated energies using the supermolecule (or discrete), the continuum, and the combined discrete-continuum solvation models for methanol and ammonium ions. A few years later, a similar model was used in Nancy to revisit the above mentioned problem, the irregular order of the basicity of amines. 108 The discrete-continuum model in the Nancy version was also used by Sánchez-Marcos et al. 178 to study the coordination of water around metallic cations in aqueous solutions. 178 The work was to anticipate a very fruitful avenue of research in Seville on the development of intermolecular potentials for statistical simulations of metal ions using the “hydrated ion” concept. 179 Among the large number of works using similar discrete-continuum combined models, let us mention a few pioneers. Aleman et al. 180 used the model of Pisa to study the solvent effect in aqueous solution on the activation barrier of the SN2 reaction. Pappalardo and Sánchez-Marcos used the model of Nancy to analyze the influence of solvation on the isomeric equilibria of nitroenamines. The same model was used by Tuñón et al. to study the hydration of the proton 181 and the hydroxide ion 182 in aqueous solution. Similar approaches focused on diverse hydrated cations. 183 , 184 , 185 The work by Pliego 185 established the basis for the calculation of thermodynamic properties with this model, which opened the door to the calculation of acid dissociation constants. 186 , 187 Another important achievement was the work by Antonczak et al. in 1994 96 devoted to the mechanism of the formamide hydrolysis reaction, which to our knowledge was the first study of water-assisted reactions using a discrete-continuum approach. Many other key studies using the discrete-continuum model could be cited. But, despite its attractiveness, the model also showed important limitations concerning, in particular, the small number of explicit solvent molecules that can be efficiently handled in a single calculation.

At this point, we must return to the 1970s. A fundamental contribution in this decade came from Martin Karplus, Michael Levitt and Arieh Warshel, the 2013 Nobel Laureates in Chemistry. In 1972, 188 as part of the postdoctoral project of Arieh Warshel with Martin Karplus at Harvard, they published a method to calculate potential energy surfaces of large conjugated systems. The π electronic system was described by a semiempirical quantum chemical method while a σ bonding orbital representing the connection to the rest of the system was described through an empirical potential. The idea of combing Quantum Mechanics (QM) and Molecular Mechanics (MM) approaches was incipient in this work! It will be further explored and extended by Warshel with the aim of studying enzymatic reactions. Thus, in 1976, now in collaboration with Levitt (sharing their time between the Weizmann Institute, Rehovot, Israel, and the Laboratory of Molecular Biology, Cambridge, England) they proposed a hybrid classical/quantum mechanical approach in which the atoms of the substrate involved in the reaction are described quantum mechanically, while the rest of the enzyme and the surrounding water solvent are treated classically. 189 The polarization (dielectric behavior) of the surroundings was accounted for by calculating the electrostatic field stemming from induced dipoles on the protein atoms and water molecules, the latter being represented by a grid of polarizable Langevin type dipoles. The model of course needed to be calibrated, using for instance some reference solvation free energies. The two articles mentioned in this paragraph are considered to be the foundational works of the hybrid QM/MM methodology which will be described in the next section.

In the following years, Warshel will report several improvements of the approach aimed at obtaining accurate solvation free energies for the study of chemical reactions in solution. Thus, the surface constrained soft sphere dipoles (SCSSD) model will be published in 1979, 190 and the more elaborated surface constraint all atom solvent (SCAAS) model 191 a few years later, in 1985. The methods were broadly applied in connection with Warshel’s Empirical Valence Bond (EVB) approach, 192 , 193 in which a reacting system is described by a calibrated Hamiltonian in terms of diabatic states corresponding to classical valence bond structures (typically ionic and covalent resonance states). The strengths of this approach were demonstrated in numerous applications, as for instance in the study of SN2 reactions in aqueous solution, 194 in which the importance of solvent fluctuations in driving the system toward the transition state was revealed. Later, the EVB approach will also be used in QM/MM simulations, as presented below. Additional material on the applications of the above mentioned methods and on other contributions of Warshel to solvation modeling can be found in Ref. 195], in his Nobel lecture, 196 and in the references cited therein.

Another innovative approach to calculate the free energy profile of chemical reactions in solution was the one developed by William Jorgensen and his group, 197 , 198 which basically has three steps. First, the minimum energy reaction path for the process of interest in the gas phase is calculated using any quantum chemical method chosen. Then, for the series of structures along the reaction coordinate calculated in the previous step, suitable solute-solvent interaction potentials are defined in order to carry out statistical simulations. In principle, this can be done by fitting the potential to ab initio calculations on complexes containing the reactant system and solvent molecules in different configurations. However, in many cases, the problem is reduced to calculating the net charges on the atoms of the reactant system along the reaction path, taking other necessary parameters from a universal force field such as OPLS 199 , 200 , 201 , 202 (solvent parameters and Lennard-Jones parameters for the reactant system). Finally, statistical simulations (usually Monte Carlo) are carried out to calculate the potential of mean force in solution. The use of techniques such as perturbation theory allows this objective to be achieved by obtaining solvation free energy differences between nearby structures. The method was used to study the SN2 reaction, 203 which was a process extensively investigated by theoreticians due to the remarkable solvation effect observed in experiments and the possibility of defining simple model systems (its variant, the Menshutkin reaction displays even more spectacular effects and became also popular thanks in particular to works by Bertrán and coworkers 204 ). The approach of Jorgensen aimed at study general organic reactions and was applied to many different processes, such as the Diels-Alder 205 and Claisen rearrangement 206 reactions. The methodology revealed to be very useful to calculate free energy profiles in solution provided the reaction path in the gas phase does not differ much from the reaction path in solution, and the challenge was to know when this is really the case. Calculations with the continuum model, which were shown to produce solvation free energies similar to simulations using the free energy perturbation method, 146 offered the possibility to optimize transition structures and reaction pathways in solution, and thus to check the greater or lesser degree of variation with respect to gas-phase reaction pathways (see, e.g., Refs. 94], 117]).

Meanwhile, classical Monte Carlo and Molecular Dynamics simulations were being widely used to calculate equilibrium and non-equilibrium properties in the liquid phase. The understanding of reaction dynamics received a great deal of attention in the 1970s and 1980s, and the availability of computer simulations was particularly attractive for testing new theories and analytical models. The work by James T. Hynes and coworkers on the theory of chemical reaction dynamics in solution 207 occupies a prominent place in this context. The theory of activated reactions known as Grote-Hynes theory (GHT) was published in two papers 208 , 209 and the predictions of the theory were confirmed via molecular dynamics simulations for the SN1 and SN2 reactions, 210 , 211 , 212 and others. An important finding in the simulations was that the transmission coefficients are typically not very far from 1, as assumed in Transition State Theory (TST), with values in the range 0.5–0.8 being usual in GHT. A note by Hynes in a recent personal perspective on the subject offers a reading of those findings: “my interpretation was, and is, that these results show that TST is in fact an excellent rate constant theory for reactions with notable barriers and that GHT shows why this is so”. 213 In addition to the SN1 and SN2 reactions, pioneer studies on reaction dynamics using simulations by Hynes and other authors were devoted to the study of fundamental processes such as ion pair association, electron and proton transfer, etc. (see for instance Refs. 210], 212], [214], [215], [216], [217], [218], [219], [220], [221], [222], [223) and have strongly contributed to the understanding of chemical reaction dynamics in solution. The simulation techniques and applications up to 1993 were reviewed in the excellent book chapter written by Whitnell and Wilson. 224

By the mid-1980s, much experience had been accumulated using continuum models, classical simulations and mixed quantum/classical partitions. Things were mature enough to take a step forward.

The success of models combining quantum chemical methods and statistical mechanics simulations

Ab initio molecular dynamics and related methods

In 1985, Roberto Car and Michele Parrinello working at SISSA (Trieste, Italy) reported an innovating method that will mark a milestone in the field of the theoretical study of complex systems. 225 The Car-Parrinello (CP) method offered the possibility of combining a classical molecular dynamics of the nuclear motions with a full quantum chemical description of the electronic structure of the system using density functional theory. The method was applied to the study of crystalline silicon. In CP molecular dynamics simulations, the interactions in the system (forces on atoms) are computed on the fly from first principles using the Kohn–Sham formulation of DFT. The electronic orbitals are expanded in a plane-wave basis set and in contrast to standard DFT calculations, in the CP method, they are not obtained from a minimization procedure via a self-consistent field calculation. Instead, the electronic degrees of freedom are treated as fictitious classical variables, having a fictitious time-dependence and a parametric fictitious mass. The whole dynamics of the system (nuclei, electronic orbitals) is described by an extended Lagrangian operator. The orbitals are propagated from an initially fully minimized set. They are maintained at a fictitious temperature which is well below the real nuclear temperature but allows the orbitals to relax rapidly under the effect of nuclear movements. The CP approach bears similarities with the principles of the simulated annealing minimization technique published a couple of years before, 226 as is clearly depicted in the review by Dominik Marx. 227 Since the CP method accounts for the electronic structure of the whole system, global polarization effects are explicitly included, and of course, the method can describe chemical events such as the formation and breaking of chemical bonds. Computational resources required to carry out simulations of this type are, however, considerably large, and the simulations were initially limited to small systems (tens of atoms) and short times (a few ps). Although the CP method was developed assuming classical dynamics for the nuclei, and is generally used that way, it was extended to incorporate nuclear quantum effects. via the Feynman path integral approach. 228 , 229 , 230

Early applications of the CP method were devoted to amorphous and liquid carbon and silicon, liquid metals, and semiconductors. 231 But the CP method will soon be applied by Parrinello and coworkers to the study of the structure and properties of liquid water and of water autoionization, 232 , 233 , 234 , 235 , 236 , 237 , 238 in some cases including quantum effects. 239 , 240 , 241 The study of liquid water with the CP method was an important breakthrough in the field of solvation and initiated new avenues in simulations of liquids and solutions (see the review by Sprik 242 ), chemical reactions (see for instance Refs. [243], [244], [245], [246], [247], [248), surfaces, 249 , 250 , 251 , 252 , 253 systems of biochemical interest, 254 , 255 etc. The study of reactions having a significant activation barrier (typically much greater than kT) by molecular dynamics simulations is confronted to the low probability of sampling the transitions state region within the time scale explored in usual calculations. The CP method did not escape of course to this issue and ad hoc solutions based on biased simulations were proposed. 256 , 257 , 258

The development of the Car-Parrinello method incited the development of many other related approaches that include electronic structure calculations in molecular dynamics simulations. In the Born-Oppenheimer molecular dynamics approach, the DFT energy is fully minimized at each time step. In other words, the electronic structure is fully converged by solving the Kohn–Sham equations for each nuclear configuration along the simulation. The electronic structure still depends on time but its time-dependence is linked to the movements of the nuclei and is not intrinsic, as in the CP method. A method combining the advantages of the Car-Parrinello and Born-Oppenheimer molecular dynamics was proposed by Parrinello and coworkers aiming at accelerating the simulations. 259 The choice of the basis set type has also generated considerable discussion. The use of plane waves in the original CP method had various advantages (orthogonality, independence of nuclei positions) and appeared as a natural choice for treating periodic systems. However, properties of chemical systems are usually local, and this prompted out different groups to explore simulations using other basis sets with localized functions. Gaussians 260 , 261 and discrete variable representations 262 , 263 were proposed, each choice presenting its own advantages and drawbacks.

Born–Oppenheimer and extended Lagrangian molecular dynamics methods are now grouped under the general name of ab initio molecular dynamics (AIMD). Furthermore, the basic principles of AIMD simulations are not limited to the use of DFT-based electronic structure methods and similar simulations can be done using other QM theories. References to numerous works using AIMD simulations at the density functional, Hartree–Fock, generalized valence bond, complete active space SCF (CASSCF), full configuration interaction, semiempirical, and other approximate QM levels until 2006 were collected in Ref. 227] More recently, simulations using the self-consistent charge density functional tight-binding method have attracted interest too. 264 , 265 , 266 The term “direct molecular dynamics” (DMD) is increasingly being used to refer to all those simulations 267 in which the time-independent Schrödinger equation is solved at each time step to calculate the forces on the nuclei. It is worth mentioning that the first DMD simulations at the semiempirical 268 or ab initio 269 levels of simple systems were carried more than 50 years ago.

For more details on the formalisms used and examples of the application of CP, AIMD and related methods, the reader is invited to consult the many general reviews available in the literature. 231 , 242 , 262 , 270 , 271 , 272 , 273 , 274 , 275 , 276 , 277

The rise of AIMD simulations and the growing interest in complex systems led to the appearance of several techniques for performing full QM calculations of systems of hundreds or thousands of atoms at an affordable computational cost. The combination of such techniques with statistical mechanics to perform DMD simulations has had a variable degree of success but for completeness, I will briefly mention some of them in this subsection.

Wesolowski and Warshel developed a frozen density functional approach 278 , 279 for solute-solvent systems in which solvent molecules are represented by the nuclei and a frozen electron density. The eigenvalue problem is solved for the solute interacting with the frozen charge distribution of the solvent molecules. The interaction energy of the system is calculated in the DFT framework using the sum of the densities of the fragments, which in addition to the exchange-correlation functional, requires the use of a kinetic energy functional. Wesolowski, Weber and Chermette 280 , 281 , 282 made a comparative study using several kinetic energy functionals. Besides, important efforts were made by different groups to develop linear-scaling algorithms, i. e. approximate quantum-mechanical techniques where the computational cost scales linearly with the increasing size of the system (instead of ∼N 3 scaling in DFT or semiempirical QM computations and even much higher in correlated ab initio computations). In the Divide and Conquer (D&C) method first developed by Yang, 283 , 284 , 285 , 286 the molecular system is partitioned into subsystems for which the electronic problem is solved, and the electron density or density matrix of the whole system is then assembled appropriately. Merz and coworkers 287 , 288 made important developments of the model, specially at the semiempirical level. They also developed a composite Hamiltonian where part of the system is described at the DFT level while the rest of the system is described at a lower level. 289 Our study of liquid water is an example of application of the D&C approach in the field of solvation by Born-Oppenheimer semiempirical molecular dynamics. 290

A similar philosophy to D&C was adopted by Gadre and coworkers 291 , 292 in the Molecular Tailoring Approach (MTA), where the large system is divided also into small overlapping fragments and the ab initio energy of the whole system is obtained from the energies of the fragments using the so-called inclusion-exclusion principle. The Fragment Molecular Orbital (FMO) approach of Kitaura, Komeiji and coworkers 293 , 294 , 295 exploits the theory of molecular interactions and defines the total energy of the system as a sum of the fragment energies and intermolecular interaction energies. Finally, some methods make use of the transferability principle, such as the Lego method developed by Mezey, 296 , 297 which uses a library of fragment density matrices, or the ELMO method of Genoni and coworkers 298 , 299 which uses tail-free strictly localized molecular orbitals on small molecular subunits.

Combined QM/MM methods

Slightly after the emergence of the Car-Parrinello method, another cutting-edge approach for simulations of large systems based on hybrid quantum/classical potentials was going to emerge. The idea of coupling quantum mechanics and molecular mechanics (QM/MM) went back to the works of Levitt, Warshel and Karplus in the 1970s but papers by Singh and Kollman in 1986, 300 and by Bash, Field and Karplus in 1987 301 made key steps forward. In another paper published in 1990 by Field, Bash and Karplus, 302 the use of such a hybrid potential in molecular dynamics simulations was extensively discussed and the approach now known as QM/MM was definitively established.

In the QM/MM approach, a large system is divided into two subsystems or regions: the chemically interesting subsystem is described quantum mechanically (QM), while its environment is described by a classical molecular mechanics (MM) force-field. In the case of a solution, which is our focus here, the solute will be typically chosen as the QM subsystem and the solvent as the MM subsystem. The two subsystems are allowed to interact through electrostatic and non-electrostatic terms. The non-electrostatic QM/MM term is calculated using a Lennard-Jones potential (or equivalent expression), as in classical simulations. The electrostatic QM/MM term takes into account the interaction of the classical charges in the MM subsystem with the nuclei and electrons in the QM subsystem. In the original papers of Karplus and coworkers, the authors used a semi-empirical approach. Since in these methods there is no precise definition of the electrostatic potential, QM/MM electrostatic interactions are not simply expressed by Coulombic formulae but require a special treatment. Several possibilities were explored, although the question raised a lot of discussion during several years (see for instance Refs. [303], [304], [305], [306 and references therein). Whatever the approximation used, the part of the QM/MM interaction involving electronic coordinates is introduced into the Hamiltonian of the QM subsystem in the SCF procedure.

The QM/MM approach knew an immediate success. Tapia et al. 307 developed a generalized reaction-field theory where the fluctuations of the solvent response were accounted for through QM/MM calculations, though they used solute-solvent configurations from classical simulations. The first QM/MM simulations were carried out at the semiempirical level, as in the original papers by Karplus and coworkers. In the early 1990s, Jiali Gao made fundamental contributions in this field using Monte Carlo simulations, in particular assessing the role of polarization effects 308 , 309 and developing methods to calculate solvation free energies 310 and free energy surfaces of chemical reactions. 311 Some other QM/MM semiempirical studies in the 1990s were reported by Vasilyev et al., 312 Bash et al., 313 Bakowies and Thiel, 305 , 314 Thompson, 315 Field, 316 Cummins and Gready, 317 and Li et al. 318 Simulations using the EVB methodology were also performed by Warshel and coworkers. 195 , 319 The first DFT/MM simulations were carried out by Kenny Merz 320 , 321 , 322 and by us 323 , 324 , 325 , 326 , 327 and were quickly followed by studies of Freindorf and Gao, 328 Lyne et al., 329 Amara et al., 330 among others. Wei and Salahub 331 proposed a kind of sequential scheme in which DFT/MM calculations were done on snapshots obtained in a classical simulation. Part of our DFT/MM molecular dynamics simulations were devoted to exploring reaction dynamics in solution. Simulating reactive trajectories at the DFT theoretical level was a major challenge being undertaken for the first time, and its realization required an enormous computational effort. Two different kinds of processes were studied. The first one was a proton transfer in a strongly hydrogen-bonded system in aqueous solution presenting a low activation barrier, so that reactive trajectories were spontaneously obtained during the simulations. 326 The second one was the charge separation process in water occurring in the second step of ethylene bromination, a process displaying a significant activation barrier. 327 Since in this case the process cannot be observed spontaneously at the time scale of the simulations (tens of ps), we applied a “rare event” sampling technique. 332 It allowed us to obtain a set of reactive and non-reactive trajectories from which barrier recrossings and non-equilibrium solvent effects could be quantified, leading finally to the calculation of the overall transmission coefficient of the process.

There are some technical issues in QM/MM simulations (semi-empirical or ab initio) that have generated considerable debate in the literature. We will briefly comment on some of them. One concerns the non-electrostatic QM/MM interaction energy. Typically, a Lennard-Jones potential or similar is used to calculate its contribution, but the parameters for QM atoms are not defined a priori and, in principle, should be optimized for each application and QM level in use. 328 In addition, the optimum parameters for QM atoms could change during the course of a chemical reaction and there is no reason to keep them constant throughout the process. A systematic technique to optimize Lennard-Jones parameters for QM/MM simulations was reported in Ref. 333] It should be noted that, in some cases, the use of parametric expressions for non-electrostatic QM/MM energy can also be seen as an advantage of the method over full ab initio MD simulations, where solute-solvent and solvent-solvent dispersion energy terms are very approximate at commonly used QM levels. Another issue is related to electronic polarization effects. It was noted 334 that the interaction of point charges with a QM molecule can lead to non-physical distortions of the electronic cloud, and that this distortion tends to be worse as the basis set increases. At short distances, this can create artifacts in QM/MM simulations. A possible solution was proposed 335 consisting of replacing the short-range repulsion term in the Lennard-Jones potential (or its equivalent) by pseudopotentials that include an explicit dependence on electronic coordinates. Another point of intense discussion was the treatment of the QM/MM frontier when the QM and MM subsystems are linked by a covalent bond. This is a key question for the study of enzymatic reactions, or the study of large molecules in general. Different solutions were proposed to deal with a QM/MM boundary. The simplest one was the use of a “link-atom” (typically a monovalent atom, H or halogen, a pseudoatom, etc.) to saturate the valency of the QM system. 300 , 302 In the local-SCF (LSCF) method of Rivail and coworkers 336 , 337 the frontier bond is a strictly localized bond obtained in a reference calculation, a concept extended by Gao, Field and coworkers 338 , 339 through the generalized hybrid orbital (GHO) approach, in which the frontier orbital is a hybrid orbital calculated on the fly during simulations. Finally, though former QM/MM works did not include explicit polarization effects on the classical solvent partition, it is possible to do so by using either self-consistent of matrix inversion methods. 315 , 340 The method known as Direct Reaction Field (DRF) developed by van Duijnen and co-workers 341 , 342 , 343 is an elegant and general approach to deal with solvent polarization. Depending on how QM/MM methods take electrostatic interactions into account (or not), Bakowies and Thiel 314 made a classification into methods involving mechanical embedding, electrostatic embedding and polarized embedding, a definition that is still used today. It is not possible to provide here an exhaustive list of papers devoted to these and related issues, so the reader is referred to the various reviews on QM/MM methods presented below.

Despite the QM/MM partition represented a tremendous advancement for the theoretical study of chemical properties in solution, simulations remained quite expensive and different groups worked on algorithms to reduce the computational cost. Enhanced sampling techniques and free energy perturbation theory, for instance, appeared promising. 344 , 345 From a different perspective, some authors developed QM/MM methods within the mean field approximation, also assumed in continuum models, which resulted in a drastic reduction of the number of quantum chemical calculations. In the mean field approximation, an average solvent potential is used to calculate the solute wavefunction, instead of calculating this potential and the wavefunction at each time step along the simulation, then making appropriate averages. In the RISM-SCF model (RISM 346 stands for reference interaction site method) developed by Hirata, Kato, and coworkers 347 , 348 , 349 , 350 the average solvent structure was obtained by using an integral equation method for atomic charges of the solute coming from quantum chemical calculations with a modified Hamiltonian. Instead, in the average solvent electrostatic potential (ASEP) approach developed by Aguilar, Olivares del Valle and coworkers, 351 , 352 , 353 the average solvent potential acting on the solute is obtained from classical molecular dynamics simulations. The ASEP method is iterative. In the simulations, the solute is represented by a set of atomic charges obtained in the QM calculation of the previous iteration (plus a suitable set of Lennard-Jones parameters). From the simulation, the average solvent electrostatic potential acting on the solute is calculated and introduced into the solute QM Hamiltonian to calculate a new wavefunction and a new set of charges. The ASEP can also be replaced by an average solvent electrostatic configuration (ASEC), and this possibility has successfully been explored by the group of Canuto and Coutinho. 354 , 355 Both ASEP and ASEC can efficiently be used to study reaction mechanisms and spectroscopic properties in solution 356 , 357 in connection with the free energy gradient (FEG) theory developed by Nagaoka and coworkers, 358 , 359 , 360 which allows to calculate first and second derivatives of the free energy of a solvated system within a QM/MM scheme. By means of free energy perturbation theory, in addition, the calculation of such derivatives can be done at very high ab initio levels. 361 , 362

Of course, the QM/MM approach has also been used in combination with other solvation models. A multilayer QM/MM/continuum approach was explored using either multipole expansions 363 or apparent surface charge 364 alternatives for describing the reaction field. The idea is similar to that developed in the discrete-continuum model presented at the beginning of this section, but in this case solvent molecules are described with a classical force-field.

A combination of QM/MM and CP methods was first proposed by Ziegler and coworkers to study transition metal catalysis with isolated systems 365 but also model solvated systems, 366 and by Eichinger et al. 367 The approach has extensively been developed and used by Ursula Rothlisberger and coworkers to study a variety of chemical and biochemical problems. 368 , 369 , 370 , 371 , 372 , 373 , 374

Before closing this section, it is worth mentioning the innovative family of methods developed by Keiji Morokuma and coworkers 375 , 376 , 377 , 378 , 379 , 380 , 381 which eventually led to an approach known presently as the ONIOM method (for our own n-layered integrated molecular orbital and molecular mechanics). 378 , 382 In ONIOM, the molecular system is divided into two or three layers which are described using theoretical levels of different complexity, by either quantum mechanics or molecular mechanics. The energy of the system is obtained by the combination of calculations with the different methods and layers, and may include or not electrostatic embedding. Though in principle it can be used to simulate systems in solution, 383 , 384 , 385 , 386 , 387 the method is principally oriented towards the treatment of very large molecular systems in organic chemistry or biochemistry.

For additional details on the QM/MM approach, many reviews 340 , 388 , 389 , 390 , 391 , 392 , 393 are available and can be consulted by the reader.

Concluding remarks and perspectives

During the past 50 years, the development of various quantum chemical models has allowed to do an immense progress on the understanding of solvation phenomena. The wide variety of solvation models that are routinely used today to study the spectroscopy of molecules and the dynamics of reactions in solution have different levels of elaboration, so they can be chosen according to the degree of accuracy required and the computational resources available. For the sake of conciseness, this brief historical perspective has focused on the main steps that led to the models most commonly used today, although, of course, there have been many other interesting contributions that could be mentioned. For example, the development of reactive force-fields is a very useful tool for the simulation of chemical reactions in solution. 394 , 395 , 396

However, despite enormous theoretical advances in the field of solvation phenomena, there are important challenges ahead. On one hand, the study of some complex processes requires a change of scale, from the typical molecular scale (nm, ns) of present models, to the mesoscale (μm, ms), or even the macroscale. Examples of such complex processes are the reactions occurring at surfaces or interfaces, as in heterogeneous catalysis 397 or water microdroplets, 398 where various difficult reactions have been shown to proceed spontaneously. 399 , 400 On the other hand, the need of the chemical industry to move to greener synthetic reactions by reducing the use of organic solvents is promoting the development of non-conventional solvents such as ionic liquids, deep eutectic solvents, or supercritical fluids. 401 , 402 Solvation and chemical reactivity in these media are still poorly understood and much progress need to be done. Some examples are discussed in Refs. [403], [404], [405], [406], [407], [408], [409.

Like many other fields of science and technology, computer simulation of chemical processes in solution has recently benefited from the rapid expansion of artificial intelligence. Machine learning techniques, and deep learning based on neural networks, in particular, have been used to improve the efficiency (computing time, accuracy, functionality) of ab initio molecular dynamics, QM//MM methods, reactive force-fields, etc. The reader can already consult several reviews on this subject. 410 , 411 , 412 , 413 , 414 , 415 , 416 Likewise, the emerging field of quantum computing offers very interesting perspectives for the future of quantum chemistry in general, 417 , 418 , 419 and some preliminary studies using the variational quantum eigensolver algorithm 420 have been done for solvated molecules using the PCM model, 421 and for very simple systems using ab initio molecular dynamics simulations. 422

We are undoubtedly on the threshold of a new era of computational chemistry, and it is likely that the advancement of techniques and computers in the next 10–20 years will mark a qualitative leap in the way of thinking and working in molecular modeling. This formidable evolution should help theoretical chemists solve complex problems that cannot be addressed today and contribute to tackling the major challenges that chemistry will be facing in the transition to a post-fossil fuel world. The history of the quantum chemistry of molecules in solution is certainly not over yet.

Dedicatory

This perspective article is dedicated to Professors Juan Bertrán, Antonio Oliva and Jean-Louis Rivail, as well as to the memory of Dr. Daniel Rinaldi, who were pioneers in the field of solvation and who taught me the essentials. It is also dedicated to the long list of doctoral students and postdoctoral researchers with whom I have had the privilege of working on the subject throughout my career.


Corresponding author: Manuel F. Ruiz-López, Laboratoire de Physique et Chimie Théoriques, UMR CNRS 7019, University of Lorraine, CNRS, BP 70239, 54506 Vandoeuvre-lès-Nancy, France, e-mail:
Article note: A collection of invited papers to celebrate the UN’s proclamation of 2025 as the International Year of Quantum Science and Technology.

Acknowledgments

The author is grateful to the University of Lorraine and the French CNRS for their continuous support.

  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: The author has accepted responsibility for the entire content of this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The author states no conflict of interest.

  6. Research funding: None declared.

  7. Data availability: Not applicable.

References

1. Camilleri, K. How Quantum Mechanics Emerged in a Few Revolutionary Months 100 Years ago. Nature 2025, 637, 269–271; https://doi.org/10.1038/d41586-024-04217-0.Search in Google Scholar PubMed

2. Heisenberg, W. Über Quantentheoretische Umdeutung Kinematischer und Mechanischer Beziehungen. Z. Phys. 1925, 33, 879–893; https://doi.org/10.1007/bf01328377.Search in Google Scholar

3. Schrödinger, E. Quantisierung als Eigenwertproblem (Erste Mitteilung). Ann. Phys. 1926, 79, 361–376.10.1002/andp.19263840404Search in Google Scholar

4. Schrödinger, E. Quantisierung als Eigenwertproblem. (Zweite Mitteilung). Ann. Phys. 1926, 79, 489–527.10.1002/andp.19263840602Search in Google Scholar

5. Boyd, R. J. The Nobel History of Computational Chemistry. A Personal Perspective. J. Comput. Chem. 2024, 45, 1921–1935; https://doi.org/10.1002/jcc.27383.Search in Google Scholar PubMed

6. Moore, T. S.; Winmill, T. F. CLXXVII.—The State of Amines in Aqueous Solution. J. Chem. Soc. Trans. 1912, 101, 1635–1676; https://doi.org/10.1039/ct9120101635.Search in Google Scholar

7. Latimer, W. M.; Rodebush, W. H. Polarity and Ionization from the Standpoint of the Lewis Theory of Valence. J. Am. Chem. Soc. 1920, 42, 1419–1433; https://doi.org/10.1021/ja01452a015.Search in Google Scholar

8. Smith, D. A. A Brief-history of the Hydrogen-bond. In Modeling the Hydrogen Bond; Smith, D. A., Ed.; American Chemical Society: Washington DC, Vol. 569, 1994; pp 1–5.10.1021/bk-1994-0569.ch001Search in Google Scholar

9. Bratož, S. Electronic Theories of Hydrogen Bonding. In Advances in Quantum Chemistry; Löwdin, P.-O., Ed.; Academic Press: London, Vol. 3, 1967; pp 209–237.10.1016/S0065-3276(08)60090-8Search in Google Scholar

10. Kollman, P. A.; Allen, L. C. Theory of the Hydrogen Bond. Chem. Rev. 1972, 72, 283–303; https://doi.org/10.1021/cr60277a004.Search in Google Scholar

11. Szalewicz, K. Hydrogen Bond. In Encyclopedia of Physical Science and Technology; Meyers, R. A., Ed.; Academic Press: New York, 2003, 3rd ed.; pp. 505–538.10.1016/B0-12-227410-5/00322-7Search in Google Scholar

12. Pimentel, G. C. The Bonding of Trihalide and Bifluoride Ions by the Molecular Orbital Method. J. Chem. Phys. 1951, 19, 446–448; https://doi.org/10.1063/1.1748245.Search in Google Scholar

13. Nukasawa, K.; Tanaka, J.; Nagakura, S. A Note on the Hydrogen Bond. J. Phys. Soc. Jpn. 1953, 8, 792–793; https://doi.org/10.1143/jpsj.8.792.Search in Google Scholar

14. Lennard-Jones, J. E.; Pople, J. A. Molecular Association in Liquids I. Molecular Association Due to Lone-pair Electrons. Proc. R. Soc. Lond. A: Math. Phys. Sci. 1951, 205, 155–162.10.1098/rspa.1951.0023Search in Google Scholar

15. Pople, J. A. Molecular Association in Liquids. 2. A Theory of the Structure of Water. Proc. R. Soc. Lond. A: Math. Phys. Sci. 1951, 205, 163–178.10.1098/rspa.1951.0024Search in Google Scholar

16. Coulson, C. The Hydrogen Bond: A Review of the Present Interpretations and an Introduction to the Theoretical Papers Presented at the Congress. In Hydrogen Bonding; Hadzi, D.; Thompson, H. W., Eds.; Pergamon Pres: London, 1959; pp. 339–360.10.1016/B978-0-08-009140-2.50041-5Search in Google Scholar

17. Clementi, E. Study of the Electronic Structure of Molecules. II. Wavefunctions for the NH3+ HCl→ NH4Cl Reaction. J. Chem. Phys. 1967, 46, 3851–3880; https://doi.org/10.1063/1.1840458.Search in Google Scholar

18. Morokuma, K.; Pedersen, L. Molecular-orbital Studies of Hydrogen Bonds. An Ab Initio Calculation for Dimeric H2O. J. Chem. Phys. 1968, 48, 3275–3282; https://doi.org/10.1063/1.1669604.Search in Google Scholar

19. Del Bene, J. E. Molecular Orbital Theory of the Hydrogen Bond. II. Dimers Containing H2O2 and H2O. J. Chem. Phys. 1972, 56, 4923–4929; https://doi.org/10.1063/1.1676970.Search in Google Scholar

20. Alagona, G.; Pullman, A.; Scrocco, E.; Tomasi, J. Quantum-mechanical Studies of Environmental Effects on Biomolecules. I. Hydration of Formamide. Int. J. Pept. Protein Res. 1973, 5, 251–259; https://doi.org/10.1111/j.1399-3011.1973.tb03459.x.Search in Google Scholar PubMed

21. Pimentel, G. C.; McClellan, A. L. The Hydrogen Bond; W.H. Freeman and Co.: San Francisco and London, 1960.Search in Google Scholar

22. Pullman, B.; Pullman, A. The Electronic Structure of the Purine-pyrimidine Pairs of DNA. Biochim. Biophys. Acta 1959, 36, 343–350; https://doi.org/10.1016/0006-3002(59)90176-3.Search in Google Scholar PubMed

23. Pullman, B.; Pullman, A. Some Electronic Aspects of Biochemistry. Rev. Mod. Phys. 1960, 32, 428–436; https://doi.org/10.1103/revmodphys.32.428.Search in Google Scholar

24. Pullman, A. The Supermolecule Approach to the Solvation Problem. In Quantum Theory of Chemical Reactions: II: Solvent Effect, Reaction Mechanisms, Photochemical Processes; Daudel, R.; Pullman, A.; Salem, L.; Veillard, A., Eds.; Springer Netherlands: Dordrecht, 1981; pp. 1–24.10.1007/978-94-010-9716-1_1Search in Google Scholar

25. Pullman, A.; Pullman, B. New Paths in the Molecular Orbital Approach to Solvation of Biological Molecules. Q. Rev. Biophys. 1974, 7, 505–566; https://doi.org/10.1017/s0033583500001529.Search in Google Scholar PubMed

26. Pullman, A. The Solvent Effect: Recent Developments. In The New World of Quantum Chemistry; Pullman, B.; Parr, R., Eds.; Springer Netherlands: Dordrecht, 1976; pp. 149–187.10.1007/978-94-010-1523-3_9Search in Google Scholar

27. Lewis, T. CNDO/2 Calculations on the Acidities of Alkanes and Saturated Alcohols and the Basicities of Methylamines. Tetrahedron 1969, 25, 4117–4126; https://doi.org/10.1016/s0040-4020(01)82944-6.Search in Google Scholar PubMed

28. Graffeuil, M.; Labarre, J.-F.; Leibovici, C. Interprétation Quantique des Propriétés Physicochimiques des Alkylamines. J. Mol. Struct. 1974, 22, 97–108; https://doi.org/10.1016/0022-2860(74)80070-0.Search in Google Scholar

29. Hehre, W.; Pople, J. The Methyl Inductive Effect on Acid-base Strengths. Tetrahedron Lett. 1970, 11, 2959–2962; https://doi.org/10.1016/s0040-4039(01)98385-6.Search in Google Scholar

30. Tollenaere, J. P.; Moereels, H. All-valence Electron Calculations and Proton Affinity of Amines. Tetrahedron Lett. 1978, 19, 1347–1350; https://doi.org/10.1016/0040-4039(78)80125-7.Search in Google Scholar

31. Pullman, A.; Brochen, P. On the Origin of the Reverse Order of Intrinsic Affinities Towards H+ and Li+ in the Series NH3 to N(CH3)3. Chem. Phys. Lett. 1975, 34, 7–10; https://doi.org/10.1016/0009-2614(75)80189-8.Search in Google Scholar

32. Umeyama, H.; Morokuma, K. Origin of Alkyl Substituent Effect in the Proton Affinity of Amines, Alcohols, and Ethers. J. Am. Chem. Soc. 1976, 98, 4400–4404; https://doi.org/10.1021/ja00431a011.Search in Google Scholar

33. Pullman, A.; Armbruster, A.-M. Ab Initio Investigation of the Energy and Electronic Evolution Upon Progressive Solvation of Ammonium Ions. Chem. Phys. Lett. 1975, 36, 558–563; https://doi.org/10.1016/0009-2614(75)85337-1.Search in Google Scholar

34. Claverie, P.; Daudey, J. P.; Langlet, J.; Pullman, B.; Piazzola, D.; Huron, M. J. Studies of Solvent Effects. 1. Discrete, Continuum, and Discrete-continuum Models and their Comparison for Some Simple Cases : NH4+, CH3OH, and Substituted NH4+. J. Phys. Chem. 1978, 82, 405–418; https://doi.org/10.1021/j100493a008.Search in Google Scholar

35. Ruiz, M.; Oliva, A.; Fernandez-Alonso, J. I.; Bertrán, J. Estudio Teórico del Efecto de Solvatación Sobre la Basicidad del Amoniaco y de las Metilaminas. Anal. Quím. 1982, 78, 202–205.Search in Google Scholar

36. Ruiz-López, M. F. Estudio teórico del efecto de la solvatación sobre la basicidad del amoniaco y de las metilaminas. Master Thesis, Autonomous University of Madrid: Madrid, 1980.Search in Google Scholar

37. Frisch, M. J; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 16 Rev. C.01; Gaussian Inc.: Wallingford, CT, 2016.Search in Google Scholar

38. Barca, G. M. J.; Bertoni, C.; Carrington, L.; Datta, D.; De Silva, N.; Deustua, J. E.; Fedorov, D. G.; Gour, J. R.; Gunina, A. O.; Guidez, E.; Harville, T.; Irle, S.; Ivanic, J.; Kowalski, K.; Leang, S. S.; Li, H.; Li, W.; Lutz, J. J.; Magoulas, I.; Mato, J.; Mironov, V.; Nakata, H.; Pham, B. Q.; Piecuch, P.; Poole, D.; Pruitt, S. R.; Rendell, A. P.; Roskop, L. B.; Ruedenberg, K.; Sattasathuchana, T.; Schmidt, M. W.; Shen, J.; Slipchenko, L.; Sosonkina, M.; Sundriyal, V.; Tiwari, A.; Galvez Vallejo, J. L.; Westheimer, B.; Włoch, M.; Xu, P.; Zahariev, F.; Gordon, M. S. Recent Developments in the General Atomic and Molecular Electronic Structure System. J. Chem. Phys. 2020, 152, 152. https://doi.org/10.1063/5.0005188.Search in Google Scholar PubMed

39. Epifanovsky, E.; Gilbert, A. T. B.; Feng, X.; Lee, J.; Mao, Y.; Mardirossian, N.; Pokhilko, P.; White, A. F.; Coons, M. P.; Dempwolff, A. L.; Gan, Z.; Hait, D.; Horn, P. R.; Jacobson, L. D.; Kaliman, I.; Kussmann, J.; Lange, A. W.; Lao, K. U.; Levine, D. S.; Liu, J.; McKenzie, S. C.; Morrison, A. F.; Nanda, K. D.; Plasser, F.; Rehn, D. R.; Vidal, M. L.; You, Z.-Q.; Zhu, Y.; Alam, B.; Albrecht, B. J.; Aldossary, A.; Alguire, E.; Andersen, J. H.; Athavale, V.; Barton, D.; Begam, K.; Behn, A.; Bellonzi, N.; Bernard, Y. A.; Berquist, E. J.; Burton, H. G. A.; Carreras, A.; Carter-Fenk, K.; Chakraborty, R.; Chien, A. D.; Closser, K. D.; Cofer-Shabica, V.; Dasgupta, S.; de Wergifosse, M.; Deng, J.; Diedenhofen, M.; Do, H.; Ehlert, S.; Fang, P.-T.; Fatehi, S.; Feng, Q.; Friedhoff, T.; Gayvert, J.; Ge, Q.; Gidofalvi, G.; Goldey, M.; Gomes, J.; González-Espinoza, C. E.; Gulania, S.; Gunina, A. O.; Hanson-Heine, M. W. D.; Harbach, P. H. P.; Hauser, A.; Herbst, M. F.; Hernández Vera, M.; Hodecker, M.; Holden, Z. C.; Houck, S.; Huang, X.; Hui, K.; Huynh, B. C.; Ivanov, M.; Jász, Á.; Ji, H.; Jiang, H.; Kaduk, B.; Kähler, S.; Khistyaev, K.; Kim, J.; Kis, G.; Klunzinger, P.; Koczor-Benda, Z.; Koh, J. H.; Kosenkov, D.; Koulias, L.; Kowalczyk, T.; Krauter, C. M.; Kue, K.; Kunitsa, A.; Kus, T.; Ladjánszki, I.; Landau, A.; Lawler, K. V.; Lefrancois, D.; Lehtola, S.; Li, R. R.; Li, Y.-P.; Liang, J.; Liebenthal, M.; Lin, H.-H.; Lin, Y.-S.; Liu, F.; Liu, K.-Y.; Loipersberger, M.; Luenser, A.; Manjanath, A.; Manohar, P.; Mansoor, E.; Manzer, S. F.; Mao, S.-P.; Marenich, A. V.; Markovich, T.; Mason, S.; Maurer, S. A.; McLaughlin, P. F.; Menger, M. F. S. J.; Mewes, J.-M.; Mewes, S. A.; Morgante, P.; Mullinax, J. W.; Oosterbaan, K. J.; Paran, G.; Paul, A. C.; Paul, S. K.; Pavošević, F.; Pei, Z.; Prager, S.; Proynov, E. I.; Rák, Á.; Ramos-Cordoba, E.; Rana, B.; Rask, A. E.; Rettig, A.; Richard, R. M.; Rob, F.; Rossomme, E.; Scheele, T.; Scheurer, M.; Schneider, M.; Sergueev, N.; Sharada, S. M.; Skomorowski, W.; Small, D. W.; Stein, C. J.; Su, Y.-C.; Sundstrom, E. J.; Tao, Z.; Thirman, J.; Tornai, G. J.; Tsuchimochi, T.; Tubman, N. M.; Veccham, S. P.; Vydrov, O.; Wenzel, J.; Witte, J.; Yamada, A.; Yao, K.; Yeganeh, S.; Yost, S. R.; Zech, A.; Zhang, I. Y.; Zhang, X.; Zhang, Y.; Zuev, D.; Aspuru-Guzik, A.; Bell, A. T.; Besley, N. A.; Bravaya, K. B.; Brooks, B. R.; Casanova, D.; Chai, J.-D.; Coriani, S.; Cramer, C. J.; Cserey, G.; DePrince, A. E.III; DiStasio, R. A.Jr.; Dreuw, A.; Dunietz, B. D.; Furlani, T. R.; Goddard, W. A.III; Hammes-Schiffer, S.; Head-Gordon, T.; Hehre, W. J.; Hsu, C.-P.; Jagau, T.-C.; Jung, Y.; Klamt, A.; Kong, J.; Lambrecht, D. S.; Liang, W.; Mayhall, N. J.; McCurdy, C. W.; Neaton, J. B.; Ochsenfeld, C.; Parkhill, J. A.; Peverati, R.; Rassolov, V. A.; Shao, Y.; Slipchenko, L. V.; Stauch, T.; Steele, R. P.; Subotnik, J. E.; Thom, A. J. W.; Tkatchenko, A.; Truhlar, D. G.; Van Voorhis, T.; Wesolowski, T. A.; Whaley, K. B.; Woodcock, H. L.III; Zimmerman, P. M.; Faraji, S.; Gill, P. M. W.; Head-Gordon, M.; Herbert, J. M.; Krylov, A. I. Software for the Frontiers of Quantum Chemistry: An Overview of Developments in the Q-Chem 5 Package. J. Chem. Phys. 2021, 155, 154102. https://doi.org/10.1063/5.0055522.Search in Google Scholar PubMed PubMed Central

40. Aquilante, F.; Autschbach, J.; Carlson, R. K.; Chibotaru, L. F.; Delcey, M. G.; De Vico, L.; Fdez. Galván, I.; Ferré, N.; Frutos, L. M.; Gagliardi, L.; Garavelli, M.; Giussani, A.; Hoyer, C. E.; Li Manni, G.; Lischka, H.; Ma, D.; Malmqvist, P. Å.; Müller, T.; Nenov, A.; Olivucci, M.; Pedersen, T. B.; Peng, D.; Plasser, F.; Pritchard, B.; Reiher, M.; Rivalta, I.; Schapiro, I.; Segarra-Martí, J.; Stenrup, M.; Truhlar, D. G.; Ungur, L.; Valentini, A.; Vancoillie, S.; Veryazov, V.; Vysotskiy, V. P.; Weingart, O.; Zapata, F.; Lindh, R. Molcas 8: New Capabilities for Multiconfigurational Quantum Chemical Calculations Across the Periodic Table. J. Comput. Chem. 2016, 37, 506–541. https://doi.org/10.1002/jcc.24221.Search in Google Scholar PubMed

41. Mossotti, O. F. Memorie di matematica e fisica in Modena, Vol. 24 II, 1850; p. 49.Search in Google Scholar

42. Clausius, R. J. Die mechanische Wärmetheorie: Die mechanische Behandlung der Electricität; Druck und Verlag von Friedrich Vieweg und Sohn: Braunschweig, Vol. 2, 1879.10.1007/978-3-663-20232-5Search in Google Scholar

43. Debye, P. Einige Resultate Einer Kinetischen Theorie der Isolatoren. Physik Z. 1912, 13, 97.Search in Google Scholar

44. Born, M. Volumen und Hydratationswärme der Ionen. Z. Phys. 1920, 1, 45–48; https://doi.org/10.1007/bf01881023.Search in Google Scholar

45. Kirkwood, J. G. Theory of Solutions of Molecules Containing Widely Separated Charges with Special Application to Zwitterions. J. Chem. Phys. 1934, 2, 351–361; https://doi.org/10.1063/1.1749489.Search in Google Scholar

46. Onsager, L. Electric Moments of Molecules in Liquids. J. Am. Chem. Soc. 1936, 58, 1486–1493; https://doi.org/10.1021/ja01299a050.Search in Google Scholar

47. Böttcher, C. J. F. Theory of Electric Polarization, 2nd ed.; Elsevier: Amsterdam, 1973.10.1016/B978-0-444-41019-1.50009-2Search in Google Scholar

48. Blondel-Mégrelis, M. Between Disciplines: Jean Barriol and the Theoretical Chemistry Laboratory in Nancy. In Chemical Sciences in the 20th Century; Reinhardt, C., Ed.; Wiley VCH: Weinheim, 2001; pp. 105–118.10.1002/9783527612734.ch05Search in Google Scholar

49. Barriol, J. Les Moments Dipolaires; Gauthier-Villars: Paris, 1957.Search in Google Scholar

50. Barriol, J.; Weisbecker, A. Contribution à la Théorie de la Constante Dielectrique des Solutions. CR Acad. Sci. Paris 1964, 259, 2831–2833.Search in Google Scholar

51. Thiébaut, J.; Rivail, J.; Barriol, J. Dielectric Studies of Non‐electrolytic Solutions. J. Chem. Soc., Faraday Trans. II 1972, 68, 1253–1264.10.1039/F29726801253Search in Google Scholar

52. Rinaldi, D. GEOMO, Program N° 290, Quantum Chemistry Program Exchange (QCPE); Indiana University: Bloomington, IN, 1972.Search in Google Scholar

53. Rinaldi, D. Versatile Techniques for Semi-empirical SCF-LCAO Calculations Including Minimization of Energy. Comput. Chem. 1977, 1, 109–114; https://doi.org/10.1016/0097-8485(77)80009-0.Search in Google Scholar

54. Pople, J. A.; Segal, G. A. Approximate Self‐Consistent Molecular Orbital Theory. III. CNDO Results for AB2 and AB3 Systems. J. Chem. Phys. 1966, 44, 3289–3296; https://doi.org/10.1063/1.1727227.Search in Google Scholar

55. Pople, J. A.; Beveridge, D. L.; Dobosh, P. A. Approximate Self‐Consistent Molecular‐Orbital Theory. V. Intermediate Neglect of Differential Overlap. J. Chem. Phys. 1967, 47, 2026–2033; https://doi.org/10.1063/1.1712233.Search in Google Scholar

56. Bingham, R. C.; Dewar, M. J. S.; Lo, D. H. Ground States of Molecules. XXV. MINDO/3. Improved Version of the MINDO Semiempirical SCF-MO Method. J. Am. Chem. Soc. 1975, 97, 1285–1293; https://doi.org/10.1021/ja00839a001.Search in Google Scholar

57. Rinaldi, D.; Rivail, J. L. Recherche Rapide de la Géométrie d’une Molécule à l’aide des Méthodes LCAO Semi-empiriques ne Faisant Intervenir que des Intégrales Mono et Bicentriques. C.R. Acad. Sci. Paris 1972, 274, 1664–1667.Search in Google Scholar

58. Ruiz-López, M. F.; Rivail, J.-L. Quantum Chemistry in Solution: Fiftieth Anniversary. ChemPhysChem 2023, 24, e202300176; https://doi.org/10.1002/cphc.202300176.Search in Google Scholar PubMed

59. Granger, P.; University of Nancy. unpublished results, 1972.Search in Google Scholar

60. Botskor, I.; Rudolph, H. D.; Roussy, G. The Microwave Spectrum of One N-gauche Rotamer of Allylamine. J. Mol. Spectrosc. 1974, 53, 15–36; https://doi.org/10.1016/0022-2852(74)90257-4.Search in Google Scholar

61. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2003.Search in Google Scholar

62. Rinaldi, D.; Rivail, J. L. Polarisabilités Moléculaires et Effet Diélectrique de Milieu à l’état Liquide. Étude Théorique de la Molécule d’eau et de ses Dimères. Theor. Chim. Acta 1973, 32, 57–70; https://doi.org/10.1007/bf01209416.Search in Google Scholar

63. Rivail, J. L.; Rinaldi, D. A Quantum Chemical Approach to Dielectric Solvent Effects in Molecular Liquids. Chem. Phys. 1976, 18, 233–242; https://doi.org/10.1016/0301-0104(76)87050-4.Search in Google Scholar

64. Yomosa, S. Nonlinear Schrödinger Equation on the Molecular Complex in Solution–Towards a Biophysics. J. Phys. Soc. Jpn. 1973, 35, 1738–1746; https://doi.org/10.1143/jpsj.35.1738.Search in Google Scholar

65. Johnston, M. D.Jr.; Barfield, M. Theoretical Studies of Solvent Effects on Nuclear Spin–Spin Coupling Constants. I. The Reaction Field Model. J. Chem. Phys. 1971, 54, 3083–3096; https://doi.org/10.1063/1.1675296.Search in Google Scholar

66. Huron, M. J.; Claverie, P. Calculation of the Interaction Energy of One Molecule with its Whole Surrounding. I. Method and Application to Pure Nonpolar Compounds. J. Phys. Chem. 1972, 76, 2123–2133; https://doi.org/10.1021/j100659a011.Search in Google Scholar

67. Langlet, J.; Claverie, P.; Caillet, J.; Pullman, A. Improvements of the Continuum Model. 1. Application to the Calculation of the Vaporization Thermodynamic Quantities of Nonassociated Liquids. J. Phys. Chem. 1988, 92, 1617–1631; https://doi.org/10.1021/j100317a048.Search in Google Scholar

68. Tapia, O.; Goscinski, O. Self-consistent Reaction Field Theory of Solvent Effects. Mol. Phys. 1975, 29, 1653–1661; https://doi.org/10.1080/00268977500101461.Search in Google Scholar

69. Constanciel, R.; Tapia, O. On the Theory of Solvent Effects: The Virtual Charge Model to Represent the Solvent Polarization. Theor. Chim. Acta 1978, 48, 75–86; https://doi.org/10.1007/bf00550242.Search in Google Scholar

70. Newton, M. D. Ab Initio Hartree‐Fock Calculations with Inclusion of a Polarized Dielectric; Formalism and Application to the Ground State Hydrated Electron. J. Chem. Phys. 1973, 58, 5833–5835; https://doi.org/10.1063/1.1679211.Search in Google Scholar

71. McCreery, J. H.; Christoffersen, R. E.; Hall, G. G. The Development of Quantum Mechanical Solvent Effect Models. Macroscopic Electrostatic Contributions. J. Am. Chem. Soc. 1976, 98, 7191–7197; https://doi.org/10.1021/ja00439a015.Search in Google Scholar

72. Felder, C. E.; Applequist, J. Energies of Solute Molecules from an Atom Charge–dipole Interaction Model with a Surrounding Dielectric: Application to Gibbs Energies of Proton Transfer Between Carboxylic Acids in Water. J. Chem. Phys. 1981, 75, 2390–2398; https://doi.org/10.1063/1.442302.Search in Google Scholar

73. Brown, R. D.; Coulson, C. A. In Calcul des Fonctions d’Onde Moléculaires; Colloques Internationaux du Centre Nationale de la Recherche Scientifique: Paris, 1958; pp. 311–327.Search in Google Scholar

74. Coulson, C. A. Theoretical Considerations of the Effects of the Medium Upon Molecular Energy Levels. Proc. R. Soc. Lond. A: Math. Phys. Sci. 1960, 255, 69–72.10.1098/rspa.1960.0051Search in Google Scholar

75. Jortner, J.; Coulson, C. Environmental Effects on Atomic Energy Levels. Mol. Phys. 1961, 4, 451–464; https://doi.org/10.1080/00268976100100611.Search in Google Scholar

76. Bertran, J.; Oliva, A.; Rinaldi, D.; Rivail, J. L. Investigations on the Role of Electrostatic Interactions on Frontier Orbitals and Chemical-reactivity in the Liquid-state. Nouv. J. Chim. 1980, 4, 209–217.Search in Google Scholar

77. Binkley, J. S.; Whiteside, R. A.; Hariharan, P. C.; Seeger, R.; Pople, J. A.; Hehre, W. J.; Newton, M. D. Gaussian 76; Carnegie-Mellon University: Pittsburgh, PA, 1976.Search in Google Scholar

78. Ruiz-Lopez, M. F.; Rinaldi, D.; Rivail, J. L. Electrostatic Solvent Effect and Molecular Conformation in Liquids. An Ab Initio Approach Using a Deformable Ellipsoidal Cavity. In 4th International Congress of Quantum Chemistry: Uppsala, Sweden, 1982.Search in Google Scholar

79. Rinaldi, D.; Ruiz-Lopez, M. F.; Rivail, J. L. Ab Initio SCF Calculations on Electrostatically Solvated Molecules Using a Deformable 3-axes Ellipsoidal Cavity. J. Chem. Phys. 1983, 78, 834–838; https://doi.org/10.1063/1.444783.Search in Google Scholar

80. Ruiz-López, M. F. Contribution à l’étude théorique des effets électrostatiques de solvant en spectroscopies moléculaires. Thesis. Université de Nancy I: Vandoeuvre-lès-Nancy, 1985.Search in Google Scholar

81. Ruiz-Lopez, M.; Rinaldi, D.; Rivail, J. L.; Oliva, A. Electrostatic Solvent Effect on Spin-spin Nuclear Coupling-constants. J. Chem. Res.-S 1982, 329.Search in Google Scholar

82. Ruiz-López, M. F.; Rinaldi, D.; Rivail, J. L. Computation of Nuclear Quadrupole Coupling Constants by Semiempirical Quantum Chemical Techniques. J. Mol. Struct.: THEOCHEM 1983, 91, 373–385; https://doi.org/10.1016/0166-1280(83)80081-5.Search in Google Scholar

83. Ruiz-López, M. F.; Rinaldi, D. Electrostatic Solvent Effect on the Circular Dichroism of the Carbonyl np* Transition. J. Mol. Struct.: THEOCHEM 1983, 93, 277–281.10.1016/0166-1280(83)80113-4Search in Google Scholar

84. Ruiz-Lopez, M. F.; Rinaldi, D. Modifications of Circular-dichroism by Electrostatic Influence of the Solvent – A Theoretical Approach in the Random Phase Approximation. Chem. Phys. 1984, 86, 367–373; https://doi.org/10.1016/0301-0104(84)80025-7.Search in Google Scholar

85. Flament, J. P.; Gervais, H. P. Perturbative Solution of the RPA Equation – An Application to the Calculation of Rotational Strengths. Theor. Chim. Acta 1982, 61, 149–162; https://doi.org/10.1007/bf00549140.Search in Google Scholar

86. Ruiz-Lopez, M. F.; Rinaldi, D.; Rivail, J. L. Circular-dichroism of (1S,4R)-norcamphor and its Methyl-derivatives – Theoretical Calculation of Solvent Effects and Vibronic Coupling Perturbations. Chem. Phys. 1986, 110, 403–414; https://doi.org/10.1016/0301-0104(86)87095-1.Search in Google Scholar

87. Bertrán, J.; Rinaldi, D.; Rivail, J. L. Effet de Solvant Dans la Substitutionnucléophile du Fluorure de Méthyle par les Ions Hydroxyle et Cyanure. C. R. Acad. Sci. Paris 1979, 289, 195–198.Search in Google Scholar

88. Bertrán, J.; Oliva, A.; Rinaldi, D.; Rivail, J. L. Investigations on the Role of Electrostatic Interactions on Frontier Orbitals and Chemical-reactivity in the Liquid-state. Nouv. J. Chim. 1980, 4, 209–217.Search in Google Scholar

89. Pappalardo, R. R.; Sánchez-Marcos, E.; Ruiz-López, M. F.; Rinaldi, D.; Rivail, J.-L. Theoretical Study of Simple Push-pull Ethylenes in Solution. J. Phys. Org. Chem. 1991, 4, 141–148.10.1002/poc.610040304Search in Google Scholar

90. Rinaldi, D.; Rivail, J.-L.; Rguini, N. Fast Geometry Optimization in Self-cosistent Reaction Field Computations on Solvated Molecules. J. Comput. Chem. 1992, 13, 675–680; https://doi.org/10.1002/jcc.540130602.Search in Google Scholar

91. Sanchez Marcos, E.; Pappalardo, R. R.; Rinaldi, D. Effects of the Solvent Reaction Field on the Geometrical Structures of Hexahydrate Metallic Cations. J. Phys. Chem. 1991, 95, 8928–8932; https://doi.org/10.1021/j100175a091.Search in Google Scholar

92. Rinaldi, D.; Pappalardo, R. R. SCRFPAC, Program No. 622 Quantum Chemistry Program Exchange; Indiana University: Bloomington, IN, 1992.Search in Google Scholar

93. Bertrán, J.; Ruiz-López, M. F.; Rinaldi, D.; Rivail, J. L. Water Dimer in Liquid Water. Theoret. Chim. Acta 1992, 84, 181–194; https://doi.org/10.1007/bf01113207.Search in Google Scholar

94. Pappalardo, R. R.; Sánchez-Marcos, E.; Ruiz-López, M. F.; Rinaldi, D.; Rivail, J. L. Solvent Effects on Molecular Geometries and Isomerization Processes: A Study of Push-pull Ethylenes in Solution. J. Am. Chem. Soc. 1993, 115, 3722–3730; https://doi.org/10.1021/ja00062a043.Search in Google Scholar

95. Ruiz-López, M. F.; Assfeld, X.; García, J. I.; Mayoral, J. A.; Salvatella, L. Solvent Effects on the Mechanism and Selectivities of Asymmetric Diels-Alder Reactions. J. Am. Chem. Soc. 1993, 115, 8780–8787; https://doi.org/10.1021/ja00072a035.Search in Google Scholar

96. Antonczak, S.; Ruiz-López, M. F.; Rivail, J. L. Ab Initio Analysis of Water-assisted Reaction Mechanisms in Amide Hydrolysis. J. Am. Chem. Soc. 1994, 116, 3912–3921; https://doi.org/10.1021/ja00088a030.Search in Google Scholar

97. Ruiz-López, M. F.; Bohr, F.; Martins-Costa, M. T. C.; Rinaldi, D. Studies of Solvent Effects Using Density Functional Theory. Co-operative Interactions in H3N…HBr Proton Transfer. Chem. Phys. Lett. 1994, 221, 109–116; https://doi.org/10.1016/0009-2614(94)87025-x.Search in Google Scholar

98. Dillet, V.; Ángyán, J. G.; Rinaldi, D.; Rivail, J. L. Reaction Field Factors for a Multipole Distribution in a Cavity Surrounded by a Continuum. Chem. Phys. Lett. 1993, 202, 18–22.10.1016/0009-2614(93)85344-NSearch in Google Scholar

99. Rinaldi, D.; Bouchy, A.; Rivail, J. L.; Dillet, V. A Self-consistent Reaction Field Model of Solvation Using Distributed Multipoles. I. Energy and Energy Derivatives. J. Chem. Phys. 2004, 120, 2343–2350; https://doi.org/10.1063/1.1635355.Search in Google Scholar

100. Costa Cabral, B. J.; Rinaldi, D.; Rivail, J. L. On the Calculation of the Dispersion Contribution to the Free-energy of Solvation by Using Cavity Models. CR Acad. Sci. Paris 1984, 298, 675–678.Search in Google Scholar

101. Rinaldi, D.; Costa Cabral, B. J.; Rivail, J.-L. Influence of Dispersion Forces on the Electronic Structure of a Solvated Molecule. Chem. Phys. Lett. 1986, 125, 495–499; https://doi.org/10.1016/0009-2614(86)87087-7.Search in Google Scholar

102. Tuñón, I.; Bertrán, J.; Ruiz-López, M. F.; Rinaldi, D. Computation of Hydration Free Energies Using a Parametrized Continuum Model. Study of Equilibrium Geometries and Reactive Processes in Water Solution. J. Comput. Chem. 1996, 17, 148–155.10.1002/(SICI)1096-987X(19960130)17:2<148::AID-JCC2>3.0.CO;2-WSearch in Google Scholar

103. Costa Cabral, B. J.; Rinaldi, D.; Rivail, J. L. A Monte-Carlo Study of Electrostatic Solvation Energies in Molecular Liquids. Chem. Phys. Lett. 1982, 93, 157–161; https://doi.org/10.1016/0009-2614(82)83684-1.Search in Google Scholar

104. Rivail, J.-L.; Rinaldi, D.; Ruiz-Lopez, M. F. The Self-Consistent Reaction Field Model for Molecular Computations in Solution. In Theoretical and Computational Models for Organic Chemistry; Formosinho, S. J.; Csizmadia, I. G.; Arnaut, L. G., Eds.; Springer Netherlands: Dordrecht, 1991; pp. 79–92.10.1007/978-94-011-3584-9_5Search in Google Scholar

105. Wong, M. W.; Frisch, M. J.; Wiberg, K. B. Solvent Effects. 1. The Mediation of Electrostatic Effects by Solvents. J. Am. Chem. Soc. 1991, 113, 4776–4782; https://doi.org/10.1021/ja00013a010.Search in Google Scholar

106. Marcos, E. S.; Maraver, J.; Ruiz-López, M. F.; Bertrán, J. Electrostatic Interactions as a Factor in the Determination of the HOMO in the Liquid State. Can. J. Chem. 1986, 64, 2353–2358; https://doi.org/10.1139/v86-388.Search in Google Scholar

107. Ruiz-López, M. F.; Oliva, A.; Tuñón, I.; Bertrán, J. Self-consistent Reaction Field Calculations of Nonequilibrium Solvent Effects on Proton Transfer Processes Through Low-barrier Hydrogen Bonds. J. Phys. Chem. A 1998, 102, 10728–10735; https://doi.org/10.1021/jp983288b.Search in Google Scholar

108. Sánchez-Marcos, E.; Terryn, B.; Rivail, J. L. Protonation of Nitrogen-containing Bases in Solution: Continuum vs. Discrete-continuum Models for Aqueous Solutions. J. Phys. Chem. 1985, 89, 4695–4700; https://doi.org/10.1021/j100268a010.Search in Google Scholar

109. Pappalardo, R. R.; Martínez, J. M.; Sánchez Marcos, E. Geometrical Structure of the Cis- and Trans-isomers of 1,2-dihaloethylenes and the Energetics of their Chemical Equilibrium in Solution. Chem. Phys. Lett. 1994, 225, 202–207.10.1016/0009-2614(94)00601-6Search in Google Scholar

110. Assfeld, X.; Sordo, J. A.; González, J.; Ruiz-López, M. F.; Sordo, T. L. Electrostatic Solvent Effect on the Ketene-imine Cycloaddition Reaction. J. Mol. Struct.: THEOCHEM 1993, 287, 193–199.10.1016/0166-1280(93)87222-YSearch in Google Scholar

111. Assfeld, X.; Ruiz-López, M. F.; Gonzalez, J.; López, R.; Sordo, J. A.; Sordo, T. L. Theoretical Analysis of the Role of the Solvent on the Reaction Mechanisms: One-step Versus Two-step Ketene-imine Cycloaddition. J. Comput. Chem. 1994, 15, 479–487; https://doi.org/10.1002/jcc.540150502.Search in Google Scholar

112. López, R.; Suárez, D.; Ruiz-López, M. F.; González, J.; Sordo, J. A.; Sordo, T. L. Solvent Effects on the Stereoselectivity of Ketene-imine Cycloaddition Reactions. J. Chem. Soc. Chem. Comm. 1995, 1677–1678; https://doi.org/10.1039/c39950001677.Search in Google Scholar

113. López, R.; Ruiz-López, M. F.; Rinaldi, D.; Sordo, J. A.; Sordo, T. L. Synthesis of b-Lactams from Fluoroketenes and Imines: Ab Initio Potential Energy Surfaces in Gas Phase and in Solution. J. Phys. Chem. 1996, 100, 10600–10608; https://doi.org/10.1021/jp960691e.Search in Google Scholar

114. Del Río, E.; López, R.; Menéndez, M. I.; Sordo, T. L.; Ruiz-López, M. F. Theoretical Study of Ester Enolate-imine Condensation Route to b-Lactams. J. Comput. Chem. 1998, 19, 1826–1833; https://doi.org/10.1002/(sici)1096-987x(199812)19:16<1826::aid-jcc4>3.0.co;2-n.10.1002/(SICI)1096-987X(199812)19:16<1826::AID-JCC4>3.0.CO;2-NSearch in Google Scholar

115. López, R.; Suarez, D.; Sordo, T. L.; Ruiz-López, M. F. Theoretical Study of the [2+2] Cycloaddition of Thioketenes with Imines to form b-Thiolactams. Chem. Eur. J. 1998, 4, 328–334; https://doi.org/10.1002/(sici)1521-3765(19980210)4:2<328::aid-chem328>3.0.co;2-q.10.1002/(SICI)1521-3765(19980210)4:2<328::AID-CHEM328>3.0.CO;2-QSearch in Google Scholar

116. Cativiela, C.; García, J. I.; Mayoral, J. A.; Royo, A. J.; Salvatella, L.; Assfeld, X.; Ruiz-López, M. F. Experimental and Theoretical Study of the Influence of the Solvent on Asymmetric Diels-Alder Reactions. J. Phys. Org. Chem. 1992, 5, 230–238.10.1002/poc.610050503Search in Google Scholar

117. Assfeld, X.; Ruiz-López, M. F.; García, J. I.; Mayoral, J. A.; Salvatella, L. Importance of Electronic and Nuclear Polarization Energy on Diastereofacial Selectivity of Diels-Alder Reactions in Aqueous Solution. J. Chem. Soc., Chem. Commun. 1995, 0, 1371–1372; https://doi.org/10.1039/c39950001371.Search in Google Scholar

118. Cativiela, C.; Dillet, V.; Garcia, J.; Mayoral, J.; Ruiz-López, M. F.; Salvatella, L. Solvent Effects on Diels-Alder Reactions. A Semi-empirical Study. J. Mol. Struct.: THEOCHEM 1995, 331, 37–50; https://doi.org/10.1016/0166-1280(94)03795-m.Search in Google Scholar

119. Assfeld, X.; García, J.; García, J. I.; Mayoral, J. A.; Proietti, M. G.; Ruiz-López, M. F.; Sánchez, M. C. Investigation of Dienophile-TiCl4 Complexation by Means of X-ray Absorption and 13C-NMR Spectroscopies. J. Org. Chem. 1996, 61, 1636–1642; https://doi.org/10.1021/jo950945c.Search in Google Scholar

120. García, J.; Mayoral, J.; Salvatella, L.; Assfeld, X.; Ruiz-López, M. F. On the Conformational Preferences of a,b-Unsaturated Carbonyl Compounds. An Ab Initio Study. J. Mol. Struct.: THEOCHEM 1996, 362, 187–197.10.1016/0166-1280(95)04391-8Search in Google Scholar

121. Cancès, E.; Mennucci, B. The Escaped Charge Problem in Solvation Continuum Models. J. Chem. Phys. 2001, 115, 6130–6135; https://doi.org/10.1063/1.1401157.Search in Google Scholar

122. Ruiz-López, M. F. The Multipole Moment Expansion Solvent Continuum Model: A Brief Review. In Solvation Effects on Molecules and Biomolecules: Computational Methods and Applications; Canuto, S., Ed.; Springer: Dordrecht, Vol. 6, 2008; pp. 23–38.10.1007/978-1-4020-8270-2_2Search in Google Scholar

123. Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic Interaction of a Solute with a Continuum. A Direct Utilization of Ab Initio Molecular Potentials for the Prevision of Solvent Effects. Chem. Phys. 1981, 55, 117–129.10.1016/0301-0104(81)85090-2Search in Google Scholar

124. Bonaccorsi, R.; Petrongolo, C.; Scrocco, E.; Tomasi, J. Theoretical Investigations on the Solvation Process. Theor. Chim. Acta 1971, 20, 331–342; https://doi.org/10.1007/bf00527188.Search in Google Scholar

125. Floris, F.; Tomasi, J. Evaluation of the Dispersion Contribution to the Solvation Energy. A Simple Computational Model in the Continuum Approximation. J. Comput. Chem. 1989, 10, 616–627; https://doi.org/10.1002/jcc.540100504.Search in Google Scholar

126. Floris, F. M.; Tomasi, J.; Pascual-Ahuir, J. L. Dispersion and Repulsion Contributions to the Solvation Energy : Refinements to a Simple Computational Model in the Continuum Approximation. J. Comp. Chem. 1991, 12, 784–791; https://doi.org/10.1002/jcc.540120703.Search in Google Scholar

127. Bianco, R.; Miertus, S.; Persico, M.; Tomasi, J. Molecular Reactivity in Solution. Modelling of the Effects of the Solvent and of its Stochastic Fluctuation on an SN2 Reaction. Chem. Phys. 1992, 168, 281–292.10.1016/0301-0104(92)87162-3Search in Google Scholar

128. Cossi, M.; Persico, M.; Tomasi, J. A Simple Solvent Model for Electron Transfer Reactions. J. Mol. Liquids 1994, 60, 87–105; https://doi.org/10.1016/0167-7322(94)00741-1.Search in Google Scholar

129. Frisch, M. J.; T, G. W.; Schegel, H. B.; Gill, P. M. W.; Johnson, B. G.; Robb, M. A.; Cheseman, J. R.; Keith, T. A.; Petersson, G. A.; Montgomery, J. A.; Raghavachari, K.; Al-Laham, M. A.; Zakrzeski, V. G.; Ortiz, J. V.; Foresman, J. B.; Ciolowski, J.; Stefanov, B. B.; Nanayakkara, A.; Challacombe, M.; Peng, C. Y.; Ayala, P. Y.; Chen, W.; Wong, M. W.; Andrés, J. L.; Reploge, E. S.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Binkley, J. S.; Defrees, D. J.; Baker, J.; Stewart, J. P.; Head-Gordon, M.; Gonzalez, C.; Pople, J. A. Gaussian 94; Gaussian, Inc.: Pittsburgh, PA, 1995.Search in Google Scholar

130. Cossi, M.; Barone, V.; Cammi, R.; Tomasi, J. Ab Initio Study of Solvated Molecules: A New Implementation of the Polarizable Continuum Model. Chem. Phys. Lett. 1996, 255, 327–335; https://doi.org/10.1016/0009-2614(96)00349-1.Search in Google Scholar

131. Pomelli, C. S.; Tomasi, J. Ab Initio Study of the SN2 Reaction CH3Cl+ Cl-→ Cl-+ CH3Cl in Supercritical Water with the Polarizable Continuum Model. J. Phys. Chem. A 1997, 101, 3561–3568; https://doi.org/10.1021/jp962358g.Search in Google Scholar

132. Cancès, E.; Mennucci, B.; Tomasi, J. A New Integral Equation Formalism for the Polarizable Continuum Model: Theoretical Background and Applications to Isotropic and Anisotropic Dielectrics. J. Chem. Phys. 1997, 107, 3032–3041; https://doi.org/10.1063/1.474659.Search in Google Scholar

133. Tomasi, J.; Mennucci, B.; Cancès, E. The IEF Version of the PCM Solvation Method: An Overview of a New Method Addressed to Study Molecular Solutes at the QM Ab Initio Level. J. Mol. Struct.: THEOCHEM 1999, 464, 211–226.10.1016/S0166-1280(98)00553-3Search in Google Scholar

134. Cancès, E.; Mennucci, B. New Applications of Integral Equations Methods for Solvation Continuum Models: Ionic Solutions and Liquid Crystals. J. Math. Chem. 1998, 23, 309–326.10.1023/A:1019133611148Search in Google Scholar

135. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001; https://doi.org/10.1021/jp9716997.Search in Google Scholar

136. Klamt, A.; Schüürmann, G. COSMO: A New Approach to Dielectric Screening in Solvents with Explicit Expressions for the Screening Energy and its Gradient. J. Chem. Soc. Perkin Trans. 1993, 2, 799–805.10.1039/P29930000799Search in Google Scholar

137. Cancès, E.; Mennucci, B.; Tomasi, J. Analytical Derivatives for Geometry Optimization in Solvation Continuum Models. II. Numerical Applications. J. Chem. Phys. 1998, 109, 260–266; https://doi.org/10.1063/1.476559.Search in Google Scholar

138. Mennucci, B.; Cammi, R.; Tomasi, J. Analytical Free Energy Second Derivatives with Respect to Nuclear Coordinates: Complete Formulation for Electrostatic Continuum Solvation Models. J. Chem. Phys. 1999, 110, 6858–6870; https://doi.org/10.1063/1.478591.Search in Google Scholar

139. Olivares del Valle, F. J.; Tomasi, J. A Simple Model for Molecular Vibrations in Solution: Application to Hydrogen Fluoride and its Dimer in Polar and Non-polar Solvents. Chem. Phys. 1987, 114, 231–239.10.1016/0301-0104(87)80112-XSearch in Google Scholar

140. Aguilar, M. A.; Bianco, R.; Miertus, S.; Persico, M.; Tomasi, J. Chemical Reactions in Solution : Modelling of the Delay of Solvent Synchronism (dielectric Friction) Along the Reaction Path of an SN2 Reaction. Chem. Phys. 1993, 174, 397–407; https://doi.org/10.1016/0301-0104(93)80006-u.Search in Google Scholar

141. Aguilar, M. A.; Olivares del Valle, F. J.; Tomasi, J. Non-equilibrium Solvation: An Ab Initio Quantum-mechanical Method the Continuum Cavity Model Approximation. J. Chem. Phys. 1993, 98, 7375–7384; https://doi.org/10.1063/1.464728.Search in Google Scholar

142. Pascual‐Ahuir, J.; Silla, E.; Tomasi, J.; Bonaccorsi, R. Electrostatic Interaction of a Solute with a Continuum. Improved Description of the Cavity and of the Surface Cavity Bound Charge Distribution. J. Comput. Chem. 1987, 8, 778–787; https://doi.org/10.1002/jcc.540080605.Search in Google Scholar

143. Pascual-Ahuir, J. L.; Silla, E. GEPOL – An Improved Description of Molecular-surfaces 1. Building the Spherical Surface Set. J. Comput. Chem. 1990, 11, 1047–1060; https://doi.org/10.1002/jcc.540110907.Search in Google Scholar

144. Silla, E.; Tuñón, I.; Pascual-Ahuir, J. L. GEPOL – An Improved Description of Molecular-surfaces 2. Computing the Molecular Area and Volume. J. Comput. Chem. 1991, 12, 1077–1088.10.1002/jcc.540120905Search in Google Scholar

145. Pascual-Ahuir, J. L.; Silla, E.; Tuñón, I. GEPOL – an Improved Description of Molecular-surfaces 3. A New Algorithm for the Computation of a Solvent-excluding Surface. J. Comput. Chem. 1994, 15, 1127–1138; https://doi.org/10.1002/jcc.540151009.Search in Google Scholar

146. Orozco, M.; Jorgensen, W. L.; Luque, F. J. Comparison of 6-31G*-based MST/SCRF and FEP Evaluations of the Free Energies of Hydration for Small Neutral Molecules. J. Comp. Chem. 1993, 14, 1498–1503; https://doi.org/10.1002/jcc.540141212.Search in Google Scholar

147. Bachs, M.; Luque, F. J.; Orozco, M. Optimization of Solute Cavities and Van Der Waals Parameters in Ab Initio MST-SCRF Calculations of Neutral Molecules. J. Comp. Chem. 1994, 15, 446–454; https://doi.org/10.1002/jcc.540150408.Search in Google Scholar

148. Orozco, M.; Bachs, M.; Luque, F. J. Development of Optimized MST/SCRF Methods for Semiempirical Calculations – The MNDO and PM3 Hamiltonians. J. Comput. Chem. 1995, 16, 563–575; https://doi.org/10.1002/jcc.540160505.Search in Google Scholar

149. Luque, F. J.; Bachs, M.; Alemán, C.; Orozco, M. Extension of MST/SCRF Method to Organic Solvents: Ab Initio and Semiempirical Parametrization for Neutral Solutes in CCl4. J. Comp. Chem. 1996, 17, 806–820; https://doi.org/10.1002/(sici)1096-987x(199605)17:7<806::aid-jcc5>3.0.co;2-w.10.1002/(SICI)1096-987X(199605)17:7<806::AID-JCC5>3.0.CO;2-WSearch in Google Scholar

150. Orozco, M.; Luque, F. J. Theretical Methods for the Description of the Solvent Effect in Biomolecular Systems. Chem. Rev. 2000, 100, 4187–4225.10.1021/cr990052aSearch in Google Scholar

151. Tomasi, J. Selected Features of the Polarizable Continuum Model for the Representation of Solvation. WIRes Comput. Mol. Sci. 2011, 1, 855–867; https://doi.org/10.1002/wcms.54.Search in Google Scholar

152. Mennucci, B. Polarizable Continuum Model. WIRes Comput. Mol. Sci. 2012, 2, 386–404; https://doi.org/10.1002/wcms.1086.Search in Google Scholar

153. Tomasi, J. PCM: A Short History of the Method (1981–2001). https://www.dcci.unipi.it/images/varie/PCM_A_short_history.pdf (accessed 2025-03-04).Search in Google Scholar

154. Karelson, M. M.; Katritzky, A. R.; Zerner, M. C. Reaction Field Effects on the Electron Distribution and Chemical Reactivity of Molecules. Int. J. Quantum Chem. 1986, 30, 521–527; https://doi.org/10.1002/qua.560300745.Search in Google Scholar

155. Mikkelsen, K. V.; Dalgaard, E.; Swanstroem, P. Electron-transfer Reactions in Solution: An Ab Initio Approach. J. Phys. Chem. 1987, 91, 3081–3092; https://doi.org/10.1021/j100295a088.Search in Google Scholar

156. Karlström, G. New Approach to the Modeling of Dielectric Media Effects in Ab Initio Quantum Chemical Calculations. J. Phys. Chem. 1988, 92, 1315–1318; https://doi.org/10.1021/j100316a060.Search in Google Scholar

157. van Duijnen, P. T.; Juffer, A. H.; Dijkman, H. P. Quantum Chemistry in the Condensed Phase: An Extended Direct Reaction Field Approach. J. Mol. Struct.: THEOCHEM 1992, 260, 195–205; https://doi.org/10.1016/0166-1280(92)87043-y.Search in Google Scholar

158. Chipman, D. M. Reaction Field Treatment of Charge Penetration. J. Chem. Phys. 2000, 112, 5558–5565; https://doi.org/10.1063/1.481133.Search in Google Scholar

159. Cramer, C. J.; Truhlar, D. G. General Parameterized SCF Model for Free Energies of Solvation in Aqueous Solution. J. Am. Chem. Soc. 1991, 113, 8305–8311; https://doi.org/10.1021/ja00022a017.Search in Google Scholar

160. Klamt, A. The COSMO and COSMO-RS Solvation Models. WIRes Comput. Mol. Sci. 2011, 1, 699–709; https://doi.org/10.1002/wcms.56.Search in Google Scholar

161. Klamt, A. COSMO-RS for Aqueous Solvation and Interfaces. Fluid Phase Equilib. 2016, 407, 152–158; https://doi.org/10.1016/j.fluid.2015.05.027.Search in Google Scholar

162. Cramer, C. J.; Truhlar, D. G. A Universal Approach to Solvation Modeling. Acc. Chem. Res. 2008, 41, 760–768; https://doi.org/10.1021/ar800019z.Search in Google Scholar PubMed

163. Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396; https://doi.org/10.1021/jp810292n.Search in Google Scholar PubMed

164. Martínez, J. M.; Pappalardo, R. R.; Sánchez Marcos, E.; Mennucci, B.; Tomasi, J. Analysis of the Opposite Solvent Effects Caused by Different Solute Cavities on the Metal−Water Distance of Monoatomic Cation Hydrates. J. Phys. Chem. B 2002, 106, 1118–1123; https://doi.org/10.1021/jp012404z.Search in Google Scholar

165. Curutchet, C.; Cramer, C. J.; Truhlar, D. G.; Ruiz-López, M. F.; Rinaldi, D.; Orozco, M.; Luque, F. J. Electrostatic Component of Solvation: Comparison of SCRF Continuum Models. J. Comput. Chem. 2003, 24, 284–297; https://doi.org/10.1002/jcc.10143.Search in Google Scholar PubMed

166. Terryn, B.; Barriol, J. Calcul des Grandeurs et Coefficients Liés à la Forme d’une Molécule Définie à Partir des Rayons de van der Waals des Atomes. Cas Particulier du Volume Moléculaire. J. Chim. Phys. 1981, 78, 207–212; https://doi.org/10.1051/jcp/1981780207.Search in Google Scholar

167. Curutchet, C.; Orozco, M.; Luque, F. J. Solvation in Octanol: Parametrization of the Continuum MST Model. J. Comput. Chem. 2001, 22, 1180–1193; https://doi.org/10.1002/jcc.1076.abs.Search in Google Scholar

168. Barone, V.; Cossi, M.; Tomasi, J. A New Definition of Cavities for the Computation of Solvation Free Energies by the Polarizable Continuum Model. J. Chem. Phys. 1997, 107, 3210–3221; https://doi.org/10.1063/1.474671.Search in Google Scholar

169. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093; https://doi.org/10.1021/cr9904009.Search in Google Scholar PubMed

170. Tomasi, J. Thirty Years of Continuum Solvation Chemistry: A Review, and Prospects for the Near Future. Theoret. Chem. Acc. 2004, 112, 184–203; https://doi.org/10.1007/s00214-004-0582-3.Search in Google Scholar

171. Tomasi, J.; Persico, M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 94, 2027–2094; https://doi.org/10.1021/cr00031a013.Search in Google Scholar

172. Tomasi, J.; Cancès, E.; Pomelli, C. S.; Caricato, M.; Scalmani, G.; Frisch, M. J.; Cammi, R.; Basilevsky, M. V.; Chuev, G. N.; Mennucci, B. Modern Theories of Continuum Models. In Continuum Solvation Models in Chemical Physics; John Wiley & Sons Ltd: Chichester, 2007; pp 1–123.10.1002/9780470515235.ch1Search in Google Scholar

173. Cramer, C. J.; Truhlar, D. G. Continuum Solvation Models: Classical and Quantum Implementations. In Reviews in Computational Chemistry; Lipkowitz, K. B.; Boyd, D. B., Eds.; VCH Publishers Inc.: New York, Vol. 6, 1995; pp. 1–72.10.1002/9780470125830.ch1Search in Google Scholar

174. Cramer, C. J.; Truhlar, D. G. Implicit Solvation Models: Equilibria, Structure, Spectra, and Dynamics. Chem. Rev. 1999, 99, 2161–2200; https://doi.org/10.1021/cr960149m.Search in Google Scholar PubMed

175. Cramer, C. J.; Truhlar, D. G. Continuum Solvation Models. In Solvent Effects and Chemical Reactivity; Tapia, O.; Bertrán, J., Eds.; Springer Netherlands: Dordrecht, 2002; pp. 1–80.10.1007/0-306-46931-6_1Search in Google Scholar

176. Mennucci, B. Continuum Solvation Models: What Else Can We Learn from Them? J. Phys. Chem. Lett. 2010, 1, 1666–1674; https://doi.org/10.1021/jz100506s.Search in Google Scholar

177. Rivail, J. L.; Terryn, B.; Rinaldi, D.; Ruiz-López, M. F. Liquid State Quantum Chemistry : A Cavity Model. J. Mol. Struct.: THEOCHEM 1985, 120, 387–400; https://doi.org/10.1016/0166-1280(85)85133-2.Search in Google Scholar

178. Sánchez Marcos, E.; Pappalardo, R. R.; Rinaldi, D. Effects of the Solvent Reaction Field on the Geometrical Structures of Hexahydrate Metallic Cations. J. Phys. Chem. 1991, 95, 8928–8932; https://doi.org/10.1021/j100175a091.Search in Google Scholar

179. Martínez, J. M.; Pappalardo, R. R.; Sánchez Marcos, E. First-Principles Ion−Water Interaction Potentials for Highly Charged Monatomic Cations. Computer Simulations of Al3+, Mg2+, and Be2+ in Water. J. Am. Chem. Soc. 1999, 121, 3175–3184; https://doi.org/10.1021/ja9830748.Search in Google Scholar

180. Alemán, C.; Maseras, F.; Lledós, A.; Duran, M.; Bertrán, J. Analysis of Solvent Effect on SN2 Reactions by Different Theoretical Models. J. Phys. Org. Chem. 1989, 2, 611–622; https://doi.org/10.1002/poc.610020804.Search in Google Scholar

181. Tuñón, I.; Silla, E.; Bertrán, J. Proton Solvation in Liquid Water: An Ab Initio Study Using the Continuum Model. J. Phys. Chem. 1993, 97, 5547–5552; https://doi.org/10.1021/j100123a016.Search in Google Scholar

182. Tuñón, I.; Rinaldi, D.; Ruiz-López, M. F.; Rivail, J. L. Hydroxide Ion in Liquid Water: Structure, Energetics, and Proton Transfer Using a Mixed Discrete-continuum Ab Initio Model. J. Phys. Chem. 1995, 99, 3798–3805; https://doi.org/10.1021/j100011a056.Search in Google Scholar

183. Floris, F.; Persico, M.; Tani, A.; Tomasi, J. Free Energies and Structures of Hydrated Cations, Based on Effective Pair Potentials. Chem. Phys. 1995, 195, 207–220; https://doi.org/10.1016/0301-0104(95)00094-5.Search in Google Scholar

184. Grimm, A. R.; Bacskay, G. B.; Haymet, A. D. J. Quantum Chemical Studies of the Solvation of the Hydroxide Ion. Mol. Phys. 1995, 86, 369–384; https://doi.org/10.1080/00268979500102081.Search in Google Scholar

185. Pliego, J. R.; Riveros, J. M. The Cluster-continuum Model for the Calculation of the Solvation Free Energy of Ionic Species. J. Phys. Chem. A 2001, 105, 7241–7247; https://doi.org/10.1021/jp004192w.Search in Google Scholar

186. Pliego, J. R.; Riveros, J. M. Theoretical Calculation of pKa Using the Cluster−Continuum Model. J. Phys. Chem. A 2002, 106, 7434–7439; https://doi.org/10.1021/jp025928n.Search in Google Scholar

187. Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Adding Explicit Solvent Molecules to Continuum Solvent Calculations for the Calculation of Aqueous Acid Dissociation Constants. J. Phys. Chem. A 2006, 110, 2493–2499; https://doi.org/10.1021/jp055336f.Search in Google Scholar PubMed PubMed Central

188. Warshel, A.; Karplus, M. Calculation of Ground and Excited State Potential Surfaces of Conjugated Molecules. I. Formulation and Parametrization. J. Am. Chem. Soc. 1972, 94, 5612–5625; https://doi.org/10.1021/ja00771a014.Search in Google Scholar

189. Warshel, A.; Levitt, M. Theoretical Studies of Enzymic Reactions : Dielectric, Electrostatic and Steric Stabilization of the Carbonium Ion in the Reaction of Lysozyme. J. Mol. Biol. 1976, 103, 227–249; https://doi.org/10.1016/0022-2836(76)90311-9.Search in Google Scholar PubMed

190. Warshel, A. Calculations of Chemical Processes in Solutions. J. Phys. Chem. 1979, 83, 1640–1652; https://doi.org/10.1021/j100475a014.Search in Google Scholar

191. Warshel, A.; King, G. Polarization Constraints in Molecular Dynamics Simulation of Aqueous Solutions: The Surface Constraint All Atom Solvent (SCAAS) Model. Chem. Phys. Lett. 1985, 121, 124–129; https://doi.org/10.1016/0009-2614(85)87168-2.Search in Google Scholar

192. Warshel, A.; Weiss, R. M. An Empirical Valence Bond Approach for Comparing Reactions in Solutions and in Enzymes. J. Am. Chem. Soc. 1980, 102, 6218–6226; https://doi.org/10.1021/ja00540a008.Search in Google Scholar

193. Kamerlin, S. C. L.; Warshel, A. The Empirical Valence Bond Model: Theory and Applications. WIRes Comput. Mol. Sci. 2011, 1, 30–45; https://doi.org/10.1002/wcms.10.Search in Google Scholar

194. Warshel, A. Dynamics of Reactions in Polar Solvents. Semiclassical Trajectory Studies of Electron-transfer and Proton-transfer Reactions. J. Phys. Chem. 1982, 86, 2218–2224; https://doi.org/10.1021/j100209a016.Search in Google Scholar

195. Åqvist, J.; Warshel, A. Simulation of Enzyme Reactions Using Valence Bond Force Fields and Other Hybrid Quantum/classical Approaches. Chem. Rev. 1993, 93, 2523–2544; https://doi.org/10.1021/cr00023a010.Search in Google Scholar

196. Warshel, A. Nobel Lecture: Multiscale Modeling of Biological Functions: From Enzymes to Molecular Machines. https://www.nobelprize.org/uploads/2018/06/warshel-lecture.pdf (accessed 2025-03-17).Search in Google Scholar

197. Jorgensen, W. L. Energy Profiles for Organic Reactions in Solution. Adv. Chem. Phys. 2009, 70, 469–488; https://doi.org/10.1002/9780470122693.ch10.Search in Google Scholar

198. Jorgensen, W. L. Free Energy Calculations: A Breakthrough for Modeling Organic Chemistry in Solution. Acc. Chem. Res. 1989, 22, 184–189; https://doi.org/10.1021/ar00161a004.Search in Google Scholar

199. Jorgensen, W. L.; Chandrashekar, J.; Madura, J. D.; Impey, W. R.; Klein, M. L. Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926–935.10.1063/1.445869Search in Google Scholar

200. Jorgensen, W. L.; Madura, J. D.; Swenson, C. J. Optimized Intermolecular Potential Functions for Liquid Hydrocarbons. J. Am. Chem. Soc. 1984, 106, 6638–6646; https://doi.org/10.1021/ja00334a030.Search in Google Scholar

201. Jorgensen, W. L. Optimized Intermolecular Potential Functions for Liquid Alcohols. J. Phys. Chem. 1986, 90, 1276–1284; https://doi.org/10.1021/j100398a015.Search in Google Scholar

202. Jorgensen, W. L.; Tirado-Rives, J. The OPLS Potential Functions for Proteins, Energy Minimizations for Crystals of Cyclic Peptides and Crambin. J. Am. Chem. Soc. 1988, 110, 1657–1666; https://doi.org/10.1021/ja00214a001.Search in Google Scholar PubMed

203. Chandrasekhar, J.; Smith, S. F.; Jorgensen, W. L. SN2 Reaction Profiles in the Gas Phase and Aqueous Solution. J. Am. Chem. Soc. 1984, 106, 3049–3050; https://doi.org/10.1021/ja00322a059.Search in Google Scholar

204. Solà, M.; Lledós, A.; Duran, M.; Bertrán, J.; Abboud, J.-L. M. Analysis of Solvent Effects on the Menshutkin Reaction. J. Am. Chem. Soc. 1991, 113, 2873–2879; https://doi.org/10.1021/ja00008a013.Search in Google Scholar

205. Blake, J. F.; Jorgensen, W. L. Solvent Effects on a Diels-Alder Reaction from Computer-simulations. J. Am. Chem. Soc. 1991, 113, 7430–7432; https://doi.org/10.1021/ja00019a055.Search in Google Scholar

206. Severance, D. L.; Jorgensen, W. L. Effects of Hydration on the Claisen Rearrangement of Allyl Vinyl Ether from Computer-simulations. J. Am. Chem. Soc. 1992, 114, 10966–10968; https://doi.org/10.1021/ja00053a046.Search in Google Scholar

207. Hynes, J. T. Chemical-reaction Dynamics in Solution. Annu. Rev. Phys. Chem. 1985, 36, 573–597; https://doi.org/10.1146/annurev.pc.36.100185.003041.Search in Google Scholar

208. Northrup, S. H.; Hynes, J. T. The Stable States Picture of Chemical Reactions. I. Formulation for Rate Constants and Initial Condition Effects. J. Chem. Phys. 1980, 73, 2700–2714; https://doi.org/10.1063/1.440484.Search in Google Scholar

209. Grote, R. F.; Hynes, J. T. The Stable States Picture of Chemical Reactions. II. Rate Constants for Condensed and Gas-phase Reaction Models. J. Chem. Phys. 1980, 73, 2715–2732; https://doi.org/10.1063/1.440485.Search in Google Scholar

210. Bergsma, J. P.; Gertner, B. J.; Wilson, K. R.; Hynes, J. T. Molecular Dynamics of a Model SN2 Reaction in Water. J. Chem. Phys. 1987, 86, 1356–1376; https://doi.org/10.1063/1.452224.Search in Google Scholar

211. Gertner, B. J.; Wilson, K. R.; Zichi, D. A.; Lee, S.; Hynes, J. T. Non-equilibrium Solvation in SN1 and SN2 Reactions in Polar-solvents. Faraday Discuss. 1988, 85, 297–308; https://doi.org/10.1039/dc9888500297.Search in Google Scholar

212. Gertner, B. J.; Wilson, K. R.; Hynes, J. T. Nonequilibrium Solvation Effects on Reaction-rates for Model SN2 Reactions in Water. J. Chem. Phys. 1989, 90, 3537–3558; https://doi.org/10.1063/1.455864.Search in Google Scholar

213. Hynes, J. T. Molecules in Motion: Chemical Reaction and Allied Dynamics in Solution and Elsewhere. Annu. Rev. Phys. Chem. 2015, 66, 1–20; https://doi.org/10.1146/annurev-physchem-040214-121833.Search in Google Scholar PubMed

214. Berkowitz, M.; Karim, O. A.; McCammon, J. A.; Rossky, P. J. Sodium Chloride Ion Pair Interaction in Water: Computer Simulation. Chem. Phys. Lett. 1984, 105, 577–580; https://doi.org/10.1016/0009-2614(84)85660-2.Search in Google Scholar

215. Karim, O. A.; McCammon, J. A. Dynamics of a Sodium Chloride Ion Pair in Water. J. Am. Chem. Soc. 1986, 108, 1762–1766; https://doi.org/10.1021/ja00268a008.Search in Google Scholar

216. Ciccotti, G.; Ferrario, M.; Hynes, J. T.; Kapral, R. Molecular Dynamics Simulation of Ion Association Reactions in a Polar Solvent. J. Chim. Phys. 1988, 85, 925–929; https://doi.org/10.1051/jcp/1988850925.Search in Google Scholar

217. Hwang, J. K.; King, G.; Creighton, S.; Warshel, A. Simulation of Free Energy Relationships and Dynamics of SN2 Reactions in Aqueous Solution. J. Am. Chem. Soc. 1988, 110, 5297–5311; https://doi.org/10.1021/ja00224a011.Search in Google Scholar

218. Gertner, B. J.; Whitnell, R. M.; Wilson, K. R.; Hynes, J. T. Activation to the Transition State: Reactant and Solvent Energy Flow for a Model SN2 Reaction in Water. J. Am. Chem. Soc. 1991, 113, 74–87; https://doi.org/10.1021/ja00001a014.Search in Google Scholar PubMed

219. Borgis, D. C.; Lee, S.; Hynes, J. T. A Dynamical Theory of Nonadiabatic Proton and Hydrogen Atom Transfer Reaction Rates in Solution. Chem. Phys. Lett. 1989, 162, 19–26; https://doi.org/10.1016/0009-2614(89)85059-6.Search in Google Scholar

220. Warshel, A.; Chu, Z. Quantum Corrections for Rate Constants of Diabatic and Adiabatic Reactions in Solutions. J. Chem. Phys. 1990, 93, 4003–4015; https://doi.org/10.1063/1.458785.Search in Google Scholar

221. Nagaoka, M.; Okuno, Y.; Yamabe, T. The Chemical Reaction Molecular Dynamics Method and the Dynamic Transition State: Proton Transfer Reaction in Formamidine and Water Solvent System. J. Am. Chem. Soc. 1991, 113, 769–778; https://doi.org/10.1021/ja00003a007.Search in Google Scholar

222. Borgis, D.; Tarjus, G.; Azzouz, H. An Adiabatic Dynamical Simulation Study of the Zundel Polarization of Strongly H-bonded Complexes in Solution. J. Chem. Phys. 1992, 97, 1390–1400; https://doi.org/10.1063/1.463265.Search in Google Scholar

223. Borgis, D.; Tarjus, G.; Azzouz, H. Solvent-induced Proton Transfer in Strongly H-bonded Complexes: An Adiabatic Dynamical Simulation Study. J. Phys. Chem. 1992, 96, 3188–3191; https://doi.org/10.1021/j100187a003.Search in Google Scholar

224. Whitnell, R. M.; Wilson, K. R. Computational Molecular Dynamics of Chemical Reactions in Solution. In Reviews in Computational Chemistry; Lipkowitz, K. B.; Boyd, D. B., Eds.; VCH Publishers, Inc.: New York, Vol. 4, 1993; pp. 67–148; https://doi.org/10.1002/9780470125816.ch3.Search in Google Scholar

225. Car, R.; Parrinello, M. Unified Approach for Molecular Dynamics and Density-functional Theory. Phys. Rev. Lett. 1985, 55, 2471–2474; https://doi.org/10.1103/physrevlett.55.2471.Search in Google Scholar PubMed

226. Kirkpatrick, S.; Gelatt, C. D.; Vecchi, M. P. Optimization by Simulated Annealing. Science 1983, 220, 671–680; https://doi.org/10.1126/science.220.4598.671.Search in Google Scholar PubMed

227. Marx, D. An Introduction to Ab Initio Molecular Dynamics Simulations. In Computational Nanoscience: Do It Yourself; Grotendorst, J.; Marx, S. B. D., Eds.; John von Neumann Institute for Computing: Jülich, Vol. 31, 2006; pp. 195–244.Search in Google Scholar

228. Marx, D.; Parrinello, M. Ab Initio Path-integral Molecular Dynamics. Z. Phys. B: Condens. Matter 1994, 95, 143–144; https://doi.org/10.1007/bf01312185.Search in Google Scholar

229. Marx, D.; Parrinello, M. Ab Initio Path Integral Molecular Dynamics: Basic Ideas. J. Chem. Phys. 1996, 104, 4077–4082; https://doi.org/10.1063/1.471221.Search in Google Scholar

230. Tuckerman, M. E.; Marx, D.; Klein, M. L.; Parrinello, M. Efficient and General Algorithms for Path Integral Car–Parrinello Molecular Dynamics. J. Chem. Phys. 1996, 104, 5579–5588; https://doi.org/10.1063/1.471771.Search in Google Scholar

231. Rothlisberger, U. 15 Years of Car-Parrinello Simulations in Physics, Chemistry and Biology. In Computational Chemistry: Reviews of Current Trends; Leszczynski, J., Ed.; World Scientific Publishing Co Pte Ltd: Singapour, Vol. 6, 2001; pp. 33–68; https://doi.org/10.1142/9789812799937_0002.Search in Google Scholar

232. Laasonen, K.; Sprik, M.; Parrinello, M.; Car, R. Ab Initio Liquid Water. J. Chem. Phys. 1993, 99, 9080–9089; https://doi.org/10.1063/1.465574.Search in Google Scholar

233. Tuckerman, M.; Laasonen, K.; Sprik, M.; Parrinello, M. Ab Initio Molecular Dynamics Simulation of the Solvation and Transport of Hydronium and Hydroxyl Ions in Water. J. Chem. Phys. 1995, 103, 150–161; https://doi.org/10.1063/1.469654.Search in Google Scholar

234. Sprik, M.; Hutter, J.; Parrinello, M. Ab Initio Molecular Dynamics Simulation of Liquid Water: Comparison of Three Gradient-corrected Density Functionals. J. Chem. Phys. 1996, 105, 1142–1152; https://doi.org/10.1063/1.471957.Search in Google Scholar

235. Silvestrelli, P. L.; Bernasconi, M.; Parrinello, M. Ab Initio Infrared Spectrum of Liquid Water. Chem. Phys. Lett. 1997, 277, 478–482; https://doi.org/10.1016/s0009-2614(97)00930-5.Search in Google Scholar

236. Trout, B. L.; Parrinello, M. The Dissociation Mechanism of H2O in Water Studied by First-principles Molecular Dynamics. Chem. Phys. Lett. 1998, 288, 343–347; https://doi.org/10.1016/s0009-2614(98)00286-3.Search in Google Scholar

237. Silvestrelli, P. L.; Parrinello, M. Water Molecule Dipole in the gas and in the Liquid Phase. Phys. Rev. Lett. 1999, 82, 3308–3311; https://doi.org/10.1103/physrevlett.82.3308.Search in Google Scholar

238. Silvestrelli, P. L.; Parrinello, M. Structural, Electronic, and Bonding Properties of Liquid Water from First Principles. J. Chem. Phys. 1999, 111, 3572–3580; https://doi.org/10.1063/1.479638.Search in Google Scholar

239. Tuckerman, M. E.; Marx, D.; Klein, M. L.; Parrinello, M. On the Quantum Nature of the Shared Proton in Hydrogen Bonds. Science 1997, 275, 817–820; https://doi.org/10.1126/science.275.5301.817.Search in Google Scholar PubMed

240. Marx, D.; Tuckerman, M. E.; Hutter, J.; Parrinello, M. The Nature of the Hydrated Excess Proton in Water. Nature 1999, 397, 601–604; https://doi.org/10.1038/17579.Search in Google Scholar

241. Marx, D.; Tuckerman, M. E.; Parrinello, M. Solvated Excess Protons in Water: Quantum Effects on the Hydration Structure. J. Phys.: Condens. Matt. 2000, 12, A153–A159; https://doi.org/10.1088/0953-8984/12/8a/317.Search in Google Scholar

242. Sprik, M. Ab Initio Molecular Dynamics Simulation of Liquids and Solutions. J. Phys.: Condens. Matt. 1996, 8, 9405–9409; https://doi.org/10.1088/0953-8984/8/47/036.Search in Google Scholar

243. Curioni, A.; Andreoni, W.; Hutter, J.; Schiffer, H.; Parrinello, M. Density-functional-theory-based Molecular Dynamics Study of 1, 3, 5-trioxane and 1, 3-dioxolane Protolysis. J. Am. Chem. Soc. 1994, 116, 11251–11255; https://doi.org/10.1021/ja00104a006.Search in Google Scholar

244. Curioni, A.; Sprik, M.; Andreoni, W.; Schiffer, H.; Hutter, J.; Parrinello, M. Density Functional Theory-based Molecular Dynamics Simulation of Acid-catalyzed Chemical Reactions in Liquid Trioxane. J. Am. Chem. Soc. 1997, 119, 7218–7229; https://doi.org/10.1021/ja970935o.Search in Google Scholar

245. Meijer, E. J.; Sprik, M. A Density Functional Study of the Addition of Water to SO3 in the Gas Phase and in Aqueous Solution. J. Phys. Chem. A 1998, 102, 2893–2898; https://doi.org/10.1021/jp972146z.Search in Google Scholar

246. Liu, Z.; Siu, C.; John, S. T. Ab Initio Molecular Dynamics Study on the Hydrolysis of Molecular Chlorine. Chem. Phys. Lett. 1999, 311, 93–101.10.1016/S0009-2614(99)00809-XSearch in Google Scholar

247. Doclo, K.; Rothlisberger, U. Conformational Equilibria of Peroxynitrous Acid in Water: A First-Principles Molecular Dynamics Study. J. Phys. Chem. A 2000, 104, 6464–6469; https://doi.org/10.1021/jp0012193.Search in Google Scholar

248. Meijer, E. J.; Sprik, M. Ab Initio Molecular Dynamics Study of the Reaction of Water with Formaldehyde in Sulfuric Acid Solution. J. Am. Chem. Soc. 1998, 120, 6345–6355; https://doi.org/10.1021/ja972935u.Search in Google Scholar

249. Lindan, P.; Harrison, N.; Holender, J.; Gillan, M. First-principles Molecular Dynamics Simulation of Water Dissociation on TiO2 (110). Chem. Phys. Lett. 1996, 261, 246–252; https://doi.org/10.1016/0009-2614(96)00934-7.Search in Google Scholar

250. Boero, M.; Parrinello, M.; Terakura, K. First Principles Molecular Dynamics Study of Ziegler-Natta Heterogeneous Catalysis. J. Am. Chem. Soc. 1998, 120, 2746–2752; https://doi.org/10.1021/ja972367i.Search in Google Scholar

251. Coussens, B. B.; Buda, F.; Oevering, H.; Meier, R. J. Simulations of Ethylene Insertion in the PtII−H Bond of (H)Pt(PX3)2+ Species. Organometallics 1998, 17, 795–801; https://doi.org/10.1021/om970080o.Search in Google Scholar

252. Hass, K. C.; Schneider, W. F.; Curioni, A.; Andreoni, W. The Chemistry of Water on Alumina Surfaces: Reaction Dynamics from First Principles. Science 1998, 282, 265–268; https://doi.org/10.1126/science.282.5387.265.Search in Google Scholar PubMed

253. Zhang, C.; Hu, P.; Alavi, A. A Density Functional Theory Study of CO Oxidation on Ru (0001) at Low Coverage. J. Chem. Phys. 2000, 112, 10564–10570; https://doi.org/10.1063/1.481690.Search in Google Scholar

254. Carloni, P.; Bloechl, P. E.; Parrinello, M. Electronic Structure of the Cu, Zn Superoxide Dismutase Active Site and its Interactions with the Substrate. J. Phys. Chem. 1995, 99, 1338–1348; https://doi.org/10.1021/j100004a039.Search in Google Scholar

255. Rothlisberger, U.; Carloni, P.; Doclo, K.; Parrinello, M. A Comparative Study of Galactose Oxidase and Active Site Analogs Based on QM/MM Car-Parrinello Simulations. J. Biol. Inorg. Chem. 2000, 5, 236–250; https://doi.org/10.1007/s007750050368.Search in Google Scholar PubMed

256. VandeVondele, J.; Rothlisberger, U. Efficient Multidimensional Free Energy Calculations for Ab Initio Molecular Dynamics Using Classical Bias Potentials. J. Chem. Phys. 2000, 113, 4863–4868; https://doi.org/10.1063/1.1289527.Search in Google Scholar

257. Röthlisberger, U.; Sprik, M.; Hutter, J. Time and Length Scales in Ab Initio Molecular Dynamics. In Bridging Time Scales: Molecular Simulations for the Next Decade; Nielaba, P.; Mareschal, M.; Ciccotti, G., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2002; pp. 413–442.10.1007/3-540-45837-9_15Search in Google Scholar

258. Iannuzzi, M.; Laio, A.; Parrinello, M. Efficient Exploration of Reactive Potential Energy Surfaces Using Car-Parrinello Molecular Dynamics. Phys. Rev. Lett. 2003, 90, 238302; https://doi.org/10.1103/physrevlett.90.238302.Search in Google Scholar PubMed

259. Kühne, T. D.; Krack, M.; Mohamed, F. R.; Parrinello, M. Efficient and Accurate Car-Parrinello-like Approach to Born-Oppenheimer Molecular Dynamics. Phys. Rev. Lett. 2007, 98, 066401; https://doi.org/10.1103/physrevlett.98.066401.Search in Google Scholar

260. Lippert, G.; Hutter, J.; Parrinello, M. The Gaussian and Augmented-plane-wave Density Functional Method for Ab Initio Molecular Dynamics Simulations. Theoret. Chem. Acc. 1999, 103, 124–140; https://doi.org/10.1007/s002140050523.Search in Google Scholar

261. Lippert, G.; Hutter, J.; Parrinello, M. A Hybrid Gaussian and Plane Wave Density Functional Scheme. Mol. Phys. 1997, 92, 477–488; https://doi.org/10.1080/00268979709482119.Search in Google Scholar

262. Tuckerman, M. E. Ab Initio Molecular Dynamics: Basic Concepts, Current Trends and Novel Applications. J. Phys.: Condens. Matt. 2002, 14, R1297–R1355; https://doi.org/10.1088/0953-8984/14/50/202.Search in Google Scholar

263. Liu, Y.; Yarne, D. A.; Tuckerman, M. E. Ab Initio Molecular Dynamics Calculations with Simple, Localized, Orthonormal Real-space Basis Sets. Phys. Rev. B 2003, 68, 125110; https://doi.org/10.1103/physrevb.68.125110.Search in Google Scholar

264. Maupin, C. M.; Aradi, B.; Voth, G. A. The Self-consistent Charge Density Functional Tight Binding Method Applied to Liquid Water and the Hydrated Excess Proton: Benchmark Simulations. J. Phys. Chem. B 2010, 114, 6922–6931; https://doi.org/10.1021/jp1010555.Search in Google Scholar PubMed

265. Goyal, P.; Elstner, M.; Cui, Q. Application of the SCC-DFTB Method to Neutral and Protonated Water Clusters and Bulk Water. J. Phys. Chem. B 2011, 115, 6790–6805; https://doi.org/10.1021/jp202259c.Search in Google Scholar PubMed PubMed Central

266. Cuny, J.; Cerda Calatayud, J.; Ansari, N.; Hassanali, A. A.; Rapacioli, M.; Simon, A. Simulation of Liquids with the Tight-Binding Density-Functional Approach and Improved Atomic Charges. J. Phys. Chem. B 2020, 124, 7421–7432; https://doi.org/10.1021/acs.jpcb.0c04167.Search in Google Scholar PubMed

267. Gao, J.; Rossky, P. J. The Age of Direct Chemical Dynamics. Acc. Chem. Res. 2022, 55, 471–472; https://doi.org/10.1021/acs.accounts.2c00021.Search in Google Scholar PubMed

268. Wang, I. S. Y.; Karplus, M. Dyanmics of Organic Reactions. J. Am. Chem. Soc. 1973, 95, 8160–8164; https://doi.org/10.1021/ja00805a033.Search in Google Scholar

269. Leforestier, C. Classical Trajectories Using the Full Ab Initio Potential Energy Surface H-+CH4→CH4+H-. J. Chem. Phys. 1978, 68, 4406–4410; https://doi.org/10.1063/1.435520.Search in Google Scholar

270. Tuckerman, M. E.; Ungar, P. J.; Von Rosenvinge, T.; Klein, M. L. Ab Initio Molecular Dynamics Simulations. J. Phys. Chem. 1996, 100, 12878–12887; https://doi.org/10.1021/jp960480+.10.1021/jp960480+Search in Google Scholar

271. Krack, M.; Parrinello, M. All-electron ab-initio Molecular Dynamics. Phys. Chem. Chem. Phys. 2000, 2, 2105–2112; https://doi.org/10.1039/b001167n.Search in Google Scholar

272. Tse, J. S. Ab Initio Molecular Dynamics with Density Functional Theory. Annu. Rev. Phys. Chem. 2002, 53, 249–290; https://doi.org/10.1146/annurev.physchem.53.090401.105737.Search in Google Scholar PubMed

273. Iftimie, R.; Minary, P.; Tuckerman, M. E. Ab Initio Molecular Dynamics: Concepts, Recent Developments, and Future Trends. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 6654–6659; https://doi.org/10.1073/pnas.0500193102.Search in Google Scholar PubMed PubMed Central

274. Marx, D.; Hutter, J. Ab Initio Molecular Dynamics: Basic Theory and Advanced Methods; Cambridge University Press: Cambridge, 2009.10.1017/CBO9780511609633Search in Google Scholar

275. Paquet, E.; Viktor, H. L. Computational Methods for Ab Initio Molecular Dynamics. Adv. Chem. 2018, 2018, 9839641–14; https://doi.org/10.1155/2018/9839641.Search in Google Scholar

276. Schlegel, H. B. Ab Initio Direct Dynamics. Acc. Chem. Res. 2021, 54, 3749–3759; https://doi.org/10.1021/acs.accounts.1c00390.Search in Google Scholar PubMed

277. Mouvet, F.; Villard, J.; Bolnykh, V.; Rothlisberger, U. Recent Advances in First-Principles Based Molecular Dynamics. Acc. Chem. Res. 2022, 55, 221–230; https://doi.org/10.1021/acs.accounts.1c00503.Search in Google Scholar

278. Wesolowski, T. A.; Warshel, A. Frozen Density-functional Approach for ab-initio Calculations of Solvated Molecules. J. Phys. Chem. 1993, 97, 8050–8053; https://doi.org/10.1021/j100132a040.Search in Google Scholar

279. Wesolowski, T.; Warshel, A. Ab Initio Free Energy Perturbation Calculations of Solvation Free Energy Using the Frozen Density Functional Approach. J. Phys. Chem. 1994, 98, 5183–5187; https://doi.org/10.1021/j100071a003.Search in Google Scholar

280. Wesolowski, T. A.; Chermette, H.; Weber, J. Accuracy of Approximate Kinetic Energy Functionals in the Model of Kohn-Sham Equations with Constrained Electron Density: The FH Center Dot Center Dot Center Dot NCH Complex as a Test Case. J. Chem. Phys. 1996, 105, 9182–9190; https://doi.org/10.1063/1.472823.Search in Google Scholar

281. Wesolowski, T. A.; Weber, J. Kohn-Sham Equations with Constrained Electron Density: An Iterative Evaluation of the Ground-state Electron Density of Interacting Molecules. Chem. Phys. Lett. 1996, 248, 71–76; https://doi.org/10.1016/0009-2614(95)01281-8.Search in Google Scholar

282. Wesolowski, T. A.; Weber, J. Kohn-Sham Equations with Constrained Electron Density: The Effect of Various Kinetic Energy Functional Parametrizations on the Ground-state Molecular Properties. Int. J. Quantum Chem. 1997, 61, 303–311; https://doi.org/10.1002/(sici)1097-461x(1997)61:2<303::aid-qua13>3.0.co;2-c.10.1002/(SICI)1097-461X(1997)61:2<303::AID-QUA13>3.3.CO;2-LSearch in Google Scholar

283. Yang, W. Direct Calculation of Electron Density in Density-functional Theory: Implementation for Benzene and a Tetrapeptide. Phys. Rev. A 1991, 44, 7823–7826; https://doi.org/10.1103/physreva.44.7823.Search in Google Scholar

284. Yang, W.; Lee, T.-S. A Density-matrix Divide-and-conquer Approach for Electronic Structure Calculations of Large Molecules. J. Chem. Phys. 1995, 103, 5674–5678; https://doi.org/10.1063/1.470549.Search in Google Scholar

285. York, D. M.; Lee, T.-S.; Yang, W. Parameterization and Efficient Implementation of a Solvent Model for Linear-scaling Semiempirical Quantum Mechanical Calculations of Biological Macromolecules. Chem. Phys. Lett. 1996, 263, 297–304; https://doi.org/10.1016/s0009-2614(96)01198-0.Search in Google Scholar

286. Lee, T. S.; York, D. M.; Yang, W. Linear‐scaling Semiempirical Quantum Calculations for Macromolecules. J. Chem. Phys. 1996, 105, 2744–2750; https://doi.org/10.1063/1.472136.Search in Google Scholar

287. Dixon, S. L.; Merz, Jr, K. M. Semiempirical Molecular Orbital Calculations with Linear System Size Scaling. J. Chem. Phys. 1996, 104, 6643–6649; https://doi.org/10.1063/1.471382.Search in Google Scholar

288. Dixon, S. L.; Merz, K. M. J. The Divide and Conquer Methodology Applied to Emiempirical SCF MO Calculations. In Encyclopedia of Computational Chemistry; Schleyer, P. V. R., Ed.; Wiley: Baffins Lane: Chichester, UK, Vol. 1, 1998; pp. 762–776.Search in Google Scholar

289. Gogonea, V.; Westerhoff, L. M.; Merz, K. M.Jr Quantum Mechanical/quantum Mechanical Methods. I. A Divide and Conquer Strategy for Solving the Schrödinger Equation for Large Molecular Systems Using a Composite Density Functional–semiempirical Hamiltonian. J. Chem. Phys. 2000, 113, 5604–5613; https://doi.org/10.1063/1.1290608.Search in Google Scholar

290. Monard, G.; Bernal-Uruchurtu, M. I.; Van Der Vaart, A.; Merz Jr, K. M.; Ruiz-López, M. F. Simulation of Liquid Water Using Semiempirical Hamiltonians and the Divide and Conquer Approach. J. Phys. Chem. A 2005, 109, 3425–3432; https://doi.org/10.1021/jp0459099.Search in Google Scholar PubMed

291. Gadre, S. R.; Shirsat, R. N.; Limaye, A. C. Molecular Tailoring Approach for Simulation of Electrostatic Properties. J. Phys. Chem. 1994, 98, 9165–9169; https://doi.org/10.1021/j100088a013.Search in Google Scholar

292. Sahu, N.; Gadre, S. R. Molecular Tailoring Approach: A Route for Ab Initio Treatment of Large Clusters. Acc. Chem. Res. 2014, 47, 2739–2747; https://doi.org/10.1021/ar500079b.Search in Google Scholar PubMed

293. Kitaura, K.; Ikeo, E.; Asada, T.; Nakano, T.; Uebayasi, M. Fragment Molecular Orbital Method: An Approximate Computational Method for Large Molecules. Chem. Phys. Lett. 1999, 313, 701–706; https://doi.org/10.1016/s0009-2614(99)00874-x.Search in Google Scholar

294. Komeiji, Y.; Nakano, T.; Fukuzawa, K.; Ueno, Y.; Inadomi, Y.; Nemoto, T.; Uebayasi, M.; Fedorov, D. G.; Kitaura, K. Fragment Molecular Orbital Method: Application to Molecular Dynamics Simulation, ‘Ab Initio FMO-MD’. Chem. Phys. Lett. 2003, 372, 342–347; https://doi.org/10.1016/s0009-2614(03)00430-5.Search in Google Scholar

295. Komeiji, Y.; Ishikawa, T.; Mochizuki, Y.; Yamataka, H.; Nakano, T. Fragment Molecular Orbital Method-Based Molecular Dynamics (FMO-MD) as a Simulator for Chemical Reactions in Explicit Solvation. J. Comput. Chem. 2009, 30, 40–50; https://doi.org/10.1002/jcc.21025.Search in Google Scholar PubMed

296. Walker, P. D.; Mezey, P. G. Molecular Electron Density Lego Approach to Molecule Building. J. Am. Chem. Soc. 1993, 115, 12423–12430; https://doi.org/10.1021/ja00079a025.Search in Google Scholar

297. Exner, T. E.; Mezey, P. G. Ab Initio Quality Properties for Macromolecules Using the ADMA Approach. J. Comput. Chem. 2003, 24, 1980–1986; https://doi.org/10.1002/jcc.10340.Search in Google Scholar PubMed

298. Macetti, G.; Wieduwilt, E. K.; Assfeld, X.; Genoni, A. Localized Molecular Orbital-Based Embedding Scheme for Correlated Methods. J. Chem. Theor. Comput. 2020, 16, 3578–3596; https://doi.org/10.1021/acs.jctc.0c00084.Search in Google Scholar

299. Meyer, B.; Guillot, B.; Ruiz-López, M. F.; Genoni, A. Libraries of Extremely Localized Molecular Orbitals. 1. Model Molecules Approximation and Molecular Orbitals Transferability. J. Chem. Theor. Comput. 2016, 12, 1052–1067.10.1021/acs.jctc.5b01007Search in Google Scholar

300. Singh, U. C.; Kollman, P. A. A Combined Ab Initio Quantum Mechanical and Molecular Mechanical Method for Carrying Out Simulations on Complex Molecular Systems : Applications to the CH3Cl + Cl- Exchange Reaction and Gas Phase Protonation and Polyethers. J. Comput. Chem. 1986, 7, 718–730; https://doi.org/10.1002/jcc.540070604.Search in Google Scholar

301. Bash, P. A.; Field, M. J.; Karplus, M. Free Energy Perturbation Method for Chemical Reactions in the Condensed Phase: A Dynamic Approach Based on a Combined Quantum and Molecular Mechanics Potential. J. Am. Chem. Soc. 1987, 109, 8092–8094; https://doi.org/10.1021/ja00260a028.Search in Google Scholar

302. Field, M. J.; Bash, P. A.; Karplus, M. A Combined Quantum Mechanical and Molecular Mechanical Potential for Molecular Dynamics Simulations. J. Comput. Chem. 1990, 11, 700–733; https://doi.org/10.1002/jcc.540110605.Search in Google Scholar

303. Ford, G. P.; Wang, B. New Approach to the Rapid Semiempirical Calculation of Molecular Electrostatic Potentials Based on the am1 Wave Function: Comparison with Ab Initio HF/6-31g* Results. J. Comput. Chem. 1993, 14, 1101–1111; https://doi.org/10.1002/jcc.540140911.Search in Google Scholar

304. Luque, F.; Orozco, M. Reliability of the AM1 Wavefunction to Compute Molecular Electrostatic Potentials. Chem. Phys. Lett. 1990, 168, 269–275; https://doi.org/10.1016/0009-2614(90)85609-g.Search in Google Scholar

305. Bakowies, D.; Thiel, W. Semiempirical Treatment of Electrostatic Potentials and Partial Charges in Combined Quantum Mechanical and Molecular Mechanical Approaches. J. Comp. Chem. 1996, 17, 87–108; https://doi.org/10.1002/(sici)1096-987x(19960115)17:1<87::aid-jcc8>3.3.co;2-t.10.1002/(SICI)1096-987X(19960115)17:1<87::AID-JCC8>3.3.CO;2-TSearch in Google Scholar

306. Luque, F. J.; Reuter, N.; Cartier, A.; Ruiz-López, M. F. Calibration of the Quantum/classical Hamiltonian in Semiempirical QM/MM AM1 and PM3 Methods. J. Phys. Chem. A 2000, 104, 10923–10931; https://doi.org/10.1021/jp001974g.Search in Google Scholar

307. Tapia, O.; Colonna, F.; Angyan, J. Generalized Self-consistent Reaction Field Theory in a Multicenter-multipole ab-initio LCGO Framework. I. Electronic Properties of the Water Molecule in a Monte Carlo Sample of Liquid Water Molecules Studied with Standard Basis Sets. J. Chim. Phys. 1990, 87, 875–903; https://doi.org/10.1051/jcp/1990870875.Search in Google Scholar

308. Gao, J.; Xia, X. A Priori Evaluation of Aqueous Polarization Effects Through Monte Carlo QM-MM Simulations. Science 1992, 258, 631–635; https://doi.org/10.1126/science.1411573.Search in Google Scholar

309. Gao, J. Potential of Mean Force for the Isomerization of DMF in Aqueous Solution: A Monte Carlo QM/MM Simulation Study. J. Am. Chem. Soc. 1993, 115, 2930–2935; https://doi.org/10.1021/ja00060a047.Search in Google Scholar

310. Gao, J. Absolute Free Energy of Solvation from Monte Carlo Simulations Using Combined Quantum and Molecular Mechanical Potentials. J. Phys. Chem. 1992, 96, 537–540; https://doi.org/10.1021/j100181a009.Search in Google Scholar

311. Gao, J.; Xia, X. A Two-dimensional Energy Surface fr a Type II SN2 Reaction in Aqueous Solution. J. Am. Chem. Soc. 1993, 115, 9667–9675; https://doi.org/10.1021/ja00074a036.Search in Google Scholar

312. Vasilyev, V. V.; Bliznyuk, A. A.; Voityuk, A. A. A Combined Quantum Chemical/molecular Mechanical Study of Hydrogen-bonded Systems. Int. J. Quantum Chem. 1992, 44, 897–930; https://doi.org/10.1002/qua.560440517.Search in Google Scholar

313. Bash, P. A.; Ho, L. L.; MacKerell, A. D.; Levine, D.; Hallstrom, P. Progress Toward Chemical Accuracy in the Computer Simulation of Condensed Phase Reactions. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 3698–3703; https://doi.org/10.1073/pnas.93.8.3698.Search in Google Scholar

314. Bakowies, D.; Thiel, W. Hybrid Models for Combined Quantum Mechanical and Molecular Mechanical Approaches. J. Phys. Chem. 1996, 100, 10580–10594; https://doi.org/10.1021/jp9536514.Search in Google Scholar

315. Thompson, M. A. QM/MMpol: A Consistent Model for Solute/solvent Polarization. Application to the Aqueous Solvation and Spectroscopy of Formaldehyde, Acetaldehyde, and Acetone. J. Phys. Chem. 1996, 100, 14492–14507; https://doi.org/10.1021/jp960690m.Search in Google Scholar

316. Field, M. J. Hybrid Quantum Mechanical/molecular Mechanical Fluctuating Charge Models for Condensed Phase Simulations. Mol. Phys. 1997, 91, 835–845; https://doi.org/10.1080/00268979709482774.Search in Google Scholar

317. Cummins, P. L.; Gready, J. E. Coupled Semiempirical Molecular Orbital and Molecular Mechanics Model (QM/MM) for Organic Molecules in Aqueous Solution. J. Comp. Chem. 1997, 18, 1496–1512; https://doi.org/10.1002/(sici)1096-987x(199709)18:12<1496::aid-jcc7>3.0.co;2-e.10.1002/(SICI)1096-987X(199709)18:12<1496::AID-JCC7>3.0.CO;2-ESearch in Google Scholar

318. Li, G. S.; Martins-Costa, M. T. C.; Millot, C.; Ruiz-Lopez, M. F. AM1/TIP3P Molecular Dynamics Simulation of Imidazole Proton-relay Processes in Aqueous Solution. Chem. Phys. Lett. 1998, 297, 38–44; https://doi.org/10.1016/s0009-2614(98)01128-2.Search in Google Scholar

319. Muller, R. P.; Warshel, A. Ab Initio Calculations of Free Energy Barriers for Chemical Reactions in Solution. J. Phys. Chem. 1995, 99, 17516–17254; https://doi.org/10.1021/j100049a009.Search in Google Scholar

320. Stanton, R. V.; Hartsough, D. S.; Merz, K. M. Calculation of Solvation Free Energies Using a Density Functional/molecular Dynamics Coupled Potential. J. Phys. Chem. 1993, 97, 11868–11870; https://doi.org/10.1021/j100148a005.Search in Google Scholar

321. Stanton, R. V.; Hartsough, D. S.; Merz, K. M. An Examination of a Density Functional/molecular Mechanical Coupled Potential. J. Comp. Chem. 1995, 16, 113–128; https://doi.org/10.1002/jcc.540160110.Search in Google Scholar

322. Stanton, R. V.; Little, L. R.; Merz, K. M. An Examination of a Hartree-Fock/molecular Mechanical Coupled Potential. J. Phys. Chem. 1995, 99, 17344–17348; https://doi.org/10.1021/j100048a006.Search in Google Scholar

323. Tuñón, I.; Martins-Costa, M. T. C.; Millot, C.; Ruiz-López, M. F. A Hybrid Density Functional-Classical Molecular Dynamics Simulation of a Water Molecule in Liquid Water. J. Mol. Mod. 1995, 1, 196–201; https://doi.org/10.1007/s008940050016.Search in Google Scholar

324. Tuñón, I.; Martins-Costa, M. T. C.; Millot, C.; Ruiz-López, M. F. Coupled Density Functional/molecular Mechanics Monte Carlo Simulations of Ions in Water. The Bromide Ion. Chem. Phys. Lett. 1995, 241, 450–456; https://doi.org/10.1016/0009-2614(95)00615-b.Search in Google Scholar

325. Tuñón, I.; Martins-Costa, M. T. C.; Millot, C.; Ruiz-López, M. F.; Rivail, J. L. A Coupled Density Functional-molecular Mechanics Monte Carlo Simulation Method: The Water Molecule in Liquid Water. J. Comp. Chem. 1996, 17, 19–29; https://doi.org/10.1002/(sici)1096-987x(19960115)17:1<19::aid-jcc2>3.0.co;2-3.10.1002/(SICI)1096-987X(19960115)17:1<19::AID-JCC2>3.3.CO;2-BSearch in Google Scholar

326. Tuñón, I.; Martins-Costa, M. T. C.; Millot, C.; Ruiz-López, M. F. Molecular Dynamics Simulations of Elementary Chemical Processes in Liquid Water Using Combined Density Functional and Molecular Mechanics Potentials. I. Proton Transfer in Strongly H-bonded Complexes. J. Chem. Phys. 1997, 106, 3633–3642; https://doi.org/10.1063/1.473457.Search in Google Scholar

327. Strnad, M.; Martins-Costa, M. T. C.; Millot, C.; Tuñón, I.; Ruiz-López, M. F.; Rivail, J. L. Molecular Dynamics Simulations of Elementary Chemical Processes in Liquid Water Using Combined Density Functional and Molecular Mechanics Potentials. II. Charge Separation Processes. J. Chem. Phys. 1997, 106, 3643–3657; https://doi.org/10.1063/1.473458.Search in Google Scholar

328. Freindorf, M.; Gao, J. Optimization of the Lennard-Jones Parameters for a Combined Ab Initio Quantum Mechanical and Molecular Mechanical Potential Using the 3-21G Basis Set. J. Comp. Chem. 1996, 17, 386–395; https://doi.org/10.1002/(sici)1096-987x(199603)17:4<386::aid-jcc1>3.0.co;2-q.10.1002/(SICI)1096-987X(199603)17:4<386::AID-JCC1>3.0.CO;2-QSearch in Google Scholar

329. Lyne, P. D.; Hodoscek, M.; Karplus, M. A Hybrid QM-MM Potential Employing Hartree-Fock or Density Functional Methods in the Quantum Region. J. Phys. Chem. A 1999, 103, 3462–3471; https://doi.org/10.1021/jp982115j.Search in Google Scholar

330. Amara, P.; Volbeda, A.; Fontecilla-Camps, J. C.; Field, M. J. A Hybrid Density Functional Theory/Molecular Mechanics Study of Nickel−Iron Hydrogenase: Investigation of the Active Site Redox States. J. Am. Chem. Soc. 1999, 121, 4468–4477; https://doi.org/10.1021/ja983971b.Search in Google Scholar

331. Wei, D.; Salahub, D. R. A Combined Density Functional and Molecular Dynamics Simulation of a Quantum Water Molecule in Aqueous Solution. Chem. Phys. Lett. 1994, 224, 291–296; https://doi.org/10.1016/0009-2614(94)00540-0.Search in Google Scholar

332. Anderson, J. B. Predicting Rare Events in Molecular Dynamics. Adv. Chem. Phys. 1995, 91, 381–431; https://doi.org/10.1002/9780470141502.ch5.Search in Google Scholar

333. Martin, M. E.; Aguilar, M. A.; Chalmet, S.; Ruiz-López, M. F. An Iterative Procedure to Determine Lennard-Jones Parameters for their Use in Quantum Mechanics/Molecular Mechanics Liquid State Simulations. Chem. Phys. 2002, 284, 607–614; https://doi.org/10.1016/s0301-0104(02)00785-1.Search in Google Scholar

334. Surján, P.; Ángyán, J. G. The Reliability of the Point Charge Model Representing Intermolecular Effects in Ab Initio Calculations. Chem. Phys. Lett. 1994, 225, 258–264; https://doi.org/10.1016/0009-2614(94)00636-9.Search in Google Scholar

335. Chalmet, S.; Ruiz-López, M. F. New Approaches to the Description of Short-range Repulsion Interactions in Hybrid Quantum/classical Systems. Chem. Phys. Lett. 2000, 329, 154–159.10.1016/S0009-2614(00)00992-1Search in Google Scholar

336. Théry, V.; Rinaldi, D.; Rivail, J.-L.; Maigret, B.; Ferenczy, G. G. Quantum Mechanical Computations on Very Large Molecular Systems: The Local Self-consistent Field Method. J. Comp. Chem. 1994, 15, 269–282; https://doi.org/10.1002/jcc.540150303.Search in Google Scholar

337. Assfeld, X.; Rivail, J.-L. Quantum Chemical Computations on Parts of Large Molecules: The Ab Initio Local Self Consistent Field Method. Chem. Phys. Lett. 1996, 263, 100–106; https://doi.org/10.1016/s0009-2614(96)01165-7.Search in Google Scholar

338. Gao, J.; Amara, P.; Alhambra, C.; Field, M. J. A Generalized Hybrid Orbital (GHO) Method for the Treatment of Boundary Atoms in Combined QM/MM Calculations. J. Phys. Chem. A 1998, 102, 4714–4721; https://doi.org/10.1021/jp9809890.Search in Google Scholar

339. Amara, P.; Field, M. J.; Alhambra, C.; Gao, J. The Generalized Hybrid Orbital Method for Combined Quantum Mechanical/molecular Mechanical Calculations: Formulation and Tests of the Analytical Derivatives. Theoret. Chem. Acc. 2000, 104, 336–343; https://doi.org/10.1007/s002140000153.Search in Google Scholar

340. Ruiz-López, M. F.; Rivail, J. L. Combined Quantum Mechanics and Molecular Mechanics Approaches to Chemical and Biochemical Reactivity. In Encyclopedia of Computational Chemistry; Schleyer, P. V. R., Ed.; Wiley: New York, Vol. 1, 1998; pp 437–448.10.1002/0470845015.cca028Search in Google Scholar

341. Thole, B. T.; van Duijnen, P. T. On the Quantum Mechanical Treatment of Solvent Effects. Theor. Chim. Acta 1980, 55, 307–318; https://doi.org/10.1007/bf00549429.Search in Google Scholar

342. De Vries, A. H.; Van Duijnen, P. T.; Juffer, A. H.; Rullmann, J. A. C.; Dijkman, J. P.; Merenga, H.; Thole, B. T. Implementation of Reaction Field Methods in Quantum Chemistry Computer Codes. J. Comp. Chem. 1995, 16, 37–55; https://doi.org/10.1002/jcc.540161113.Search in Google Scholar

343. Van Duijnen, P. T.; de Vries, A. H. Direct Reaction Field Force Field: A Consistent Way to Connect and Combine Quantum-chemical and Classical Descriptions of Molecules. Int. J. Quantum Chem. 1996, 60, 1111–1132; https://doi.org/10.1002/(sici)1097-461x(1996)60:6<1111::aid-qua2>3.0.co;2-2.10.1002/(SICI)1097-461X(1996)60:6<1111::AID-QUA2>3.0.CO;2-2Search in Google Scholar

344. Martins-Costa, M. T. C.; Ruiz-López, M. F. Reaching Multi-Nanosecond Timescales in Combined QM/MM Molecular Dynamics Simulations Through Parallel Horsetail Sampling. J. Comput. Chem. 2017, 38, 659–668; https://doi.org/10.1002/jcc.24723.Search in Google Scholar

345. Martins-Costa, M. T. C.; Ruiz-López, M. F. Highly Accurate Computation of Free Energies in Complex Systems Through Horsetail QM/MM Molecular Dynamics Combined with Free-energy Perturbation Theory. Theoret. Chem. Acc. 2017, 136, 50.10.1007/s00214-017-2078-ySearch in Google Scholar

346. Chandler, D.; Andersen, H. C. Optimized Cluster Expansions for Classical Fluids. II. Theory of Molecular Liquids. J. Chem. Phys. 1972, 57, 1930–1937; https://doi.org/10.1063/1.1678513.Search in Google Scholar

347. Ten-no, S.; Hirata, F.; Kato, S. A Hybrid Approach for the Solvent Effect on the Electronic Structure of a Solute Based on the RISM and Hartree-Fock Equations. Chem. Phys. Lett. 1993, 214, 391–396; https://doi.org/10.1016/0009-2614(93)85655-8.Search in Google Scholar

348. Maw, S.; Sato, H.; Ten-no, S.; Hirata, F. Ab Initio Study of Water: Self-consistent Determination of Electronic Structure and Liquid State Properties. Chem. Phys. Lett. 1997, 276, 20–25; https://doi.org/10.1016/s0009-2614(97)88029-3.Search in Google Scholar

349. Naka, K.; Sato, H.; Morita, A.; Hirata, F.; Kato, S. RISM-SCF Study of the Free-energy Profile of the Menshutkin-type Reaction NH3+CH3Cl→NH3CH3++Cl− in Aqueous Solution. Theoret. Chem. Acc. 1999, 102, 165–169; https://doi.org/10.1007/s002140050487.Search in Google Scholar

350. Sato, H. A Modern Solvation Theory: Quantum Chemistry and Statistical Chemistry. Phys. Chem. Chem. Phys. 2013, 15, 7450–7465; https://doi.org/10.1039/c3cp50247c.Search in Google Scholar

351. Sánchez, M. L.; Aguilar, M. A.; Olivares del Valle, F. J. Study of Solvent Effects by Means of Averaged Solvent Electrostatic Potentials Obtained from Molecular Dynamics Data. J. Comput. Chem. 1997, 18, 313–322; https://doi.org/10.1002/(sici)1096-987x(199702)18:3<313::aid-jcc2>3.0.co;2-x.10.1002/(SICI)1096-987X(199702)18:3<313::AID-JCC2>3.3.CO;2-0Search in Google Scholar

352. Sánchez, M. L.; Martín, M. E.; Aguilar, M. A.; Olivares del Valle, F. J. Solvent Effects on the 1(n,p*) Transition of Formaldehyde in Liquid Water. A QM/MM Study Using the Mean Field Approximation. Chem. Phys. Lett. 1999, 310, 195–200.10.1016/S0009-2614(99)00735-6Search in Google Scholar

353. Muñoz-Losa, A.; Fdez-Galván, I.; Martín, M. E.; Aguilar, M. A. Theoretical Study of Liquid Hydrogen Fluoride. Application of the Averaged Solvent Electrostatic Potential/Molecular Dynamics Method. J. Phys. Chem. B 2003, 107, 5043–5047; https://doi.org/10.1021/jp022422w.Search in Google Scholar

354. Coutinho, K.; Georg, H. C.; Fonseca, T. L.; Ludwig, V.; Canuto, S. An Efficient Statistically Converged Average Configuration for Solvent Effects. Chem. Phys. Lett. 2007, 437, 148–152; https://doi.org/10.1016/j.cplett.2007.02.012.Search in Google Scholar

355. Coutinho, K.; Rivelino, R.; Georg, H. C.; Canuto, S. The Sequential QM/MM Method and its Applications to Solvent Effects in Electronic and Structural Properties of Solutes. In Solvation Effects on Molecules and Biomolecules; Canuto, S., Ed.; Springer: Dordrecht, 2008; pp. 159–189.10.1007/978-1-4020-8270-2_7Search in Google Scholar

356. Fdez Galván, I.; Martín, M. E.; Aguilar, M. A. A New Method to Locate Saddle Points for Reactions in Solution by Using the Free‐energy Gradient Method and the Mean Field Approximation. J. Comput. Chem. 2004, 25, 1227–1233; https://doi.org/10.1002/jcc.20048.Search in Google Scholar

357. Georg, H. C.; Canuto, S. Electronic Properties of Water in Liquid Environment. A Sequential QM/MM Study Using the Free Energy Gradient Method. J. Phys. Chem. B 2012, 116, 11247–11254; https://doi.org/10.1021/jp304201b.Search in Google Scholar

358. Okuyama‐Yoshida, N.; Nagaoka, M.; Yamabe, T. Transition‐state Optimization on Free Energy Surface: Toward Solution Chemical Reaction Ergodography. Int. J. Quantum Chem. 1998, 70, 95–103; https://doi.org/10.1002/(sici)1097-461x(1998)70:1<95::aid-qua7>3.0.co;2-0.10.1002/(SICI)1097-461X(1998)70:1<95::AID-QUA7>3.0.CO;2-0Search in Google Scholar

359. Okuyama-Yoshida, N.; Kataoka, K.; Nagaoka, M.; Yamabe, T. Structure Optimization Via Free Energy Gradient Method: Application to Glycine Zwitterion in Aqueous Solution. J. Chem. Phys. 2000, 113, 3519–3524; https://doi.org/10.1063/1.1287785.Search in Google Scholar

360. Kitamura, Y.; Takenaka, N.; Koyano, Y.; Nagaoka, M. Free Energy Gradient Method and its Recent Related Developments: Free Energy Optimization and Vibrational Frequency Analysis in Solution. In Quantum Modeling of Complex Molecular Systems; Rivail, J.-L.; Ruiz-López, M.; Assfeld, X., Eds.; Springer: Cham Heidelberg New York Dordrecht London, Vol. 21, 2015; pp. 219–252; https://doi.org/10.1007/978-3-319-21626-3_8.Search in Google Scholar

361. Bistafa, C.; Kitamura, Y.; Martins-Costa, M. T. C.; Nagaoka, M.; Ruiz-López, M. F. A Cost-Effective Method for Free-Energy Minimization in Complex Systems with Elaborated Ab Initio Potentials. J. Chem. Theor. Comp. 2018, 14, 3262–3271; https://doi.org/10.1021/acs.jctc.8b00271.Search in Google Scholar

362. Bistafa, C.; Kitamura, Y.; Martins-Costa, M. T. C.; Nagaoka, M.; Ruiz-López, M. F. Vibrational Spectroscopy in Solution Through Perturbative Ab Initio Molecular Dynamics Simulations. J. Chem. Theor. Comput. 2019, 15, 4615–4622; https://doi.org/10.1021/acs.jctc.9b00362.Search in Google Scholar

363. Chalmet, S.; Rinaldi, D.; Ruiz-López, M. F. A QM/MM/Continuum Model for Computations in Solution: Comparison with QM/MM Molecular Dynamics Simulations. Int. J. Quantum Chem. 2001, 84, 559–564.10.1002/qua.1410Search in Google Scholar

364. Cui, Q. Combining Implicit Solvation Models with Hybrid Quantum Mechanical/molecular Mechanical Methods: A Critical Test with Glycine. J. Chem. Phys. 2002, 117, 4720–4728; https://doi.org/10.1063/1.1499481.Search in Google Scholar

365. Woo, T. K.; Margl, P. M.; Blöchl, P. E.; Ziegler, T. A Combined Car−Parrinello QM/MM Implementation for Ab Initio Molecular Dynamics Simulations of Extended Systems: Application to Transition Metal Catalysis. J. Phys. Chem. B 1997, 101, 7877–7880; https://doi.org/10.1021/jp9717296.Search in Google Scholar

366. Woo, T. K.; Blöchl, P. E.; Ziegler, T. Towards Solvation Simulations with a Combined Ab Initio Molecular Dynamics and Molecular Mechanics Approach. J. Mol. Struct.: THEOCHEM 2000, 506, 313–334.10.1016/S0166-1280(00)00424-3Search in Google Scholar

367. Eichinger, M.; Tavan, P.; Hutter, J.; Parrinello, M. A Hybrid Method for Solutes in Complex Solvents : Density Functional Theory Combined with Empirical Force Fields. J. Chem. Phys. 1999, 110, 10452–10466; https://doi.org/10.1063/1.479049.Search in Google Scholar

368. Laio, A.; VandeVondele, J.; Rothlisberger, U. A Hamiltonian Electrostatic Coupling Scheme for Hybrid Car–Parrinello Molecular Dynamics Simulations. J. Chem. Phys. 2002, 116, 6941–6947; https://doi.org/10.1063/1.1462041.Search in Google Scholar

369. Laio, A.; VandeVondele, J.; Rothlisberger, U. D-RESP: Dynamically Generated Electrostatic Potential Derived Charges from Quantum Mechanics/molecular Mechanics Simulations. J. Phys. Chem. B 2002, 106, 7300–7307; https://doi.org/10.1021/jp0143138.Search in Google Scholar

370. Colombo, M. C.; Guidoni, L.; Laio, A.; Magistrato, A.; Maurer, P.; Piana, S.; Röhrig, U.; Spiegel, K.; Sulpizi, M.; VandeVondele, J.; Zumstein, M.; Röthlisberger, U. Hybrid QM/MM Car-Parrinello Simulations of Catalytic and Enzymatic Reactions. Chimia 2002, 56, 13–19. https://doi.org/10.2533/000942902777680865.Search in Google Scholar

371. Rothlisberger, U.; Carloni, P.; Doclo, K.; Parrinello, M. A Comparative Study of Galactose Oxidase and Active Site Analogs Based on QM/MM Car Parrinello Simulations. J. Biol. Inorg. Chem. 2000, 5, 236–250; https://doi.org/10.1007/s007750050368.Search in Google Scholar PubMed

372. Guidoni, L.; Maurer, P.; Piana, S.; Rothlisberger, U. Hybrid Car-Parrinello/molecular Mechanics Modelling of Transition Metal Complexes: Structure, Dynamics and Reactivity. Quant. Struct.-Act. Relat. 2002, 21, 119–127; https://doi.org/10.1002/1521-3838(200207)21:2<119::aid-qsar119>3.0.co;2-b.10.1002/1521-3838(200207)21:2<119::AID-QSAR119>3.0.CO;2-BSearch in Google Scholar

373. Moret, M. E.; Tavernelli, I.; Rothlisberger, U. Combined QM/MM and Classical Molecular Dynamics Study of Ru(bpy)32+ in Water. J. Phys. Chem. B 2009, 113, 7737–7744; https://doi.org/10.1021/jp900147r.Search in Google Scholar

374. Liberatore, E.; Meli, R.; Rothlisberger, U. A Versatile Multiple Time Step Scheme for Efficient Ab Initio Molecular Dynamics Simulations. J. Chem. Theor. Comput. 2018, 14, 2834–2842; https://doi.org/10.1021/acs.jctc.7b01189.Search in Google Scholar

375. Humbel, S.; Sieber, S.; Morokuma, K. The IMOMO Method: Integration of Different Levels of Molecular Orbital Approximations for Geometry Optimization of Large Systems: Test for n-butane Conformation and SN2 Reaction: RCl+Cl-. J. Chem. Phys. 1996, 105, 1959–1967; https://doi.org/10.1063/1.472065.Search in Google Scholar

376. Matsubara, T.; Maseras, F.; Koga, N.; Morokuma, K. Application of the New ‘‘integrated MO+MM’’ (IMOMM) Method to the Organometallic Reaction Pt(PR3)2+H2 (R=H, Me, t-Bu, and Ph). J. Phys. Chem. 1996, 100, 2573–2580; https://doi.org/10.1021/jp951762x.Search in Google Scholar

377. Matsubara, T.; Sieber, S.; Morokuma, K. A Test of the New ‘‘integrated MO+MM’’ (IMOMM) Method for the Conformational Energy of Ethane and n-butane. Int. J. Quantum Chem. 1996, 60, 1101–1109; https://doi.org/10.1002/(sici)1097-461x(1996)60:6<1101::aid-qua1>3.0.co;2-3.10.1002/(SICI)1097-461X(1996)60:6<1101::AID-QUA1>3.0.CO;2-3Search in Google Scholar

378. Svensson, M.; Humbel, S.; Froese, R. D. J.; Matsubara, T.; Sieber, S.; Morokuma, K. ONIOM: A Multilayered Integrated MO+MM Method for Geometry Optimizations and Single Point Energy Predictions. A Test for Diels-Alder Reactions and Pt(P(t-Bu)3)2+H2 Oxidative Addition. J. Phys. Chem. 1996, 100, 19357–19363; https://doi.org/10.1021/jp962071j.Search in Google Scholar

379. Svensson, M.; Humbel, S.; Morokuma, K. Energetics Using the Single Point IMOMO (integrated Molecular Orbital Plus Molecular Orbital) Calculations: Choices of Computational Levels and Model System. J. Chem. Phys. 1996, 105, 3654–3661; https://doi.org/10.1063/1.472235.Search in Google Scholar

380. Svensson, M.; Matsubara, T.; Morokuma, K. Theoretical Study of Pd(II)- and Ni(II)-catalyzed Alternating Copolymerization of Carbon Monoxide with Ethylene. Organometallics 1996, 15, 5568–5576; https://doi.org/10.1021/om960535u.Search in Google Scholar

381. Maseras, F.; Morokuma, K. IMOMM: A New Integrated Ab Initio + Molecular Mechanics Geometry Optimization Scheme of Equilibrium Structures and Transition States. J. Comp. Chem. 1995, 16, 1170–1179; https://doi.org/10.1002/jcc.540160911.Search in Google Scholar

382. Chung, L. W.; Sameera, W. M. C.; Ramozzi, R.; Page, A. J.; Hatanaka, M.; Petrova, G. P.; Harris, T. V.; Li, X.; Ke, Z.; Liu, F.; Li, H.-B.; Ding, L.; Morokuma, K. The ONIOM Method and Its Applications. Chem. Rev. 2015, 115, 5678–5796. https://doi.org/10.1021/cr5004419.Search in Google Scholar PubMed

383. Woo, T. K.; Cavallo, L.; Ziegler, T. Implementation of the IMOMM Methodology for Performing Combined QM/MM Molecular Dynamics Simulations and Frequency Calculations. Theoret. Chem. Acc. 1998, 100, 307–313; https://doi.org/10.1007/s002140050391.Search in Google Scholar

384. Rega, N.; Iyengar, S. S.; Voth, G. A.; Schlegel, H. B.; Vreven, T.; Frisch, M. J. Hybrid Ab-Initio/Empirical Molecular Dynamics: Combining the ONIOM Scheme with the Atom-Centered Density Matrix Propagation (ADMP) Approach. J. Phys. Chem. B 2004, 108, 4210–4220; https://doi.org/10.1021/jp0370829.Search in Google Scholar

385. Rega, N.; Iyengar, S. S.; Voth, G. A.; Schlegel, H. B.; Vreven, T.; Frisch, M. J. Hybrid ab-initio/empirical Molecular Dynamics: Combining the ONIOM Scheme with the Atom-centered Density Matrix Propagation (ADMP) Approach. J. Phys. Chem. B 2004, 108, 4210–4220; https://doi.org/10.1021/jp0370829.Search in Google Scholar

386. Matsubara, T.; Dupuis, M.; Aida, M. The ONIOM Molecular Dynamics Method for Biochemical Applications: Cytidine Deaminase. Chem. Phys. Lett. 2007, 437, 138–142; https://doi.org/10.1016/j.cplett.2007.01.085.Search in Google Scholar

387. Sumner, I.; Iyengar, S. S. Combining Quantum Wavepacket Ab Initio Molecular Dynamics with QM/MM and QM/QM Techniques: Implementation Blending ONIOM and Empirical Valence Bond Theory. J. Chem. Phys. 2008, 129, 054109; https://doi.org/10.1063/1.2956496.Search in Google Scholar PubMed

388. Gao, J. Hybrid Quantum and Molecular Mechanical Simulations: An Alternative Avenue to Solvent Effects in Organic Chemistry. Acc. Chem. Res. 1996, 29, 298–305; https://doi.org/10.1021/ar950140r.Search in Google Scholar

389. Gao, J.; Thompson, M. A. Combined Quantum Mechanical and Molecular Mechanical Methods, Vol. 712; ACS Publications: Washington, DC, 1998.10.1021/bk-1998-0712Search in Google Scholar

390. Sherwood, P. Hybrid Quantum Mechanics/molecular Mechanics Approaches. In Modern Methods and Algorithms of Quantum Chemistry; Grotendorst, J., Ed.; NIC-Directors: Jülich, Germany, 2000; pp. 285–305.Search in Google Scholar

391. Lin, H.; Truhlar, D. G. QM/MM: What Have we Learned, Where are we, and Where do we go from Here? Theoret. Chem. Acc. 2007, 117, 185–199; https://doi.org/10.1007/s00214-006-0143-z.Search in Google Scholar

392. Liu, M. Y.; Wang, Y. J.; Chen, Y. K.; Field, M. J.; Gao, J. L. QM/MM Through the 1990s: The First Twenty Years of Method Development and Applications. Isr. J. Chem. 2014, 54, 1250–1263; https://doi.org/10.1002/ijch.201400036.Search in Google Scholar PubMed PubMed Central

393. Rivail, J.-L.; Ruiz-López, M. F.; Assfeld, X. Quantum Modeling of Complex Molecular Systems, Vol. 21; Springer: Cham Heidelberg New York Dordrecht London, 2015.10.1007/978-3-319-21626-3Search in Google Scholar

394. van Duin, A. C.; Dasgupta, S.; Lorant, F.; Goddard, W. A. ReaxFF: A Reactive Force Field for Hydrocarbons. J. Phys. Chem. A 2001, 105, 9396–9409; https://doi.org/10.1021/jp004368u.Search in Google Scholar

395. Leven, I.; Hao, H.; Das, A. K.; Head-Gordon, T. A Reactive Force Field with Coarse-grained Electrons for Liquid Water. J. Phys. Chem. Lett. 2020, 11, 9240–9247; https://doi.org/10.1021/acs.jpclett.0c02516.Search in Google Scholar PubMed

396. van Duin, A. Reactive Force Fields: Concepts of ReaxFF. In Computational Methods in Catalysis and Materials Science: An Introduction for Scientists and Engineers; van Santen, R. A., Sautet, P., Eds.; Wiley‐VCH Verlag GmbH & Co. KGaA: Weinheim, 2009; pp 167–181.10.1002/9783527625482.ch9Search in Google Scholar

397. Hariharan, S.; Kinge, S.; Visscher, L. Modeling Heterogeneous Catalysis Using Quantum Computers: An Academic and Industry Perspective. J. Chem. Inf. Model. 2025, 65, 472–511; https://doi.org/10.1021/acs.jcim.4c01212.Search in Google Scholar PubMed PubMed Central

398. Ruiz-López, M. F.; Francisco, J. S.; Martins-Costa, M. T. C.; Anglada, J. M. Molecular Reactions at Aqueous Interfaces. Nat. Rev. Chem. 2020, 4, 459–475; https://doi.org/10.1038/s41570-020-0203-2.Search in Google Scholar PubMed

399. Qiu, L.; Cooks, R. G. Spontaneous Oxidation in Aqueous Microdroplets: Water Radical Cation as Primary Oxidizing Agent. Angew. Chem. Int. Ed. 2024, 136, e202400118; https://doi.org/10.1002/anie.202400118.Search in Google Scholar PubMed

400. Mofidfar, M.; Mehrgardi, M. A.; Zare, R. N. Water Microdroplets Surrounded by Alcohol Vapor Cause Spontaneous Oxidation of Alcohols to Organic Peroxides. J. Am. Chem. Soc. 2024, 146, 18498–18503; https://doi.org/10.1021/jacs.4c04092.Search in Google Scholar PubMed

401. Winterton, N. The Green Solvent: A Critical Perspective. Clean Technol. Environ. Policy 2021, 23, 2499–2522; https://doi.org/10.1007/s10098-021-02188-8.Search in Google Scholar PubMed PubMed Central

402. Comptes Rendus. Chimie. Special Issue “French Network on Solvation (GDR 2035 SolvATE)”; F. Ingrosso, Ed. https://comptes-rendus.academie-sciences.fr/chimie/item/CRCHIM_2024__27_S5/ (accessed 2025-04-02).Search in Google Scholar

403. Kirchner, B. Theory of Complicated Liquids: Investigation of Liquids, Solvents and Solvent Effects with Modern Theoretical Methods. Phys. Rep. 2007, 440, 1–111; https://doi.org/10.1016/j.physrep.2006.11.005.Search in Google Scholar

404. Kirchner, B. Ionic Liquids from Theoretical Investigations. In Ionic Liquids; Kirchner, B., Ed.; Springer: Berlin, Heidelberg, Vol. 290, 2009; pp. 213–262; https://doi.org/10.1007/128_2008_36.Search in Google Scholar PubMed

405. Bernales, V. S.; Marenich, A. V.; Contreras, R.; Cramer, C. J.; Truhlar, D. G. Quantum Mechanical Continuum Solvation Models for Ionic Liquids. J. Phys. Chem. B 2012, 116, 9122–9129; https://doi.org/10.1021/jp304365v.Search in Google Scholar PubMed

406. Zahn, S.; Kirchner, B.; Mollenhauer, D. Charge Spreading in Deep Eutectic Solvents. ChemPhysChem 2016, 17, 3354–3358; https://doi.org/10.1002/cphc.201600348.Search in Google Scholar PubMed

407. Ingrosso, F.; Ruiz-Lopez, M. F. Modeling Solvation in Supercritical CO2. ChemPhysChem 2017, 18, 2560–2572; https://doi.org/10.1002/cphc.201700434.Search in Google Scholar PubMed

408. Perkin, S.; Kirchner, B.; Fayer, M. D. Preface: Special Topic on Chemical Physics of Ionic Liquids. J. Chem. Phys. 2018, 148, 193501; https://doi.org/10.1063/1.5039492.Search in Google Scholar PubMed

409. Glossman-Mitnik, D.; Maciejewska, M. Solvents, Ionic Liquids and Solvent Effects; IntechOpen: London, 2020.10.5772/intechopen.82039Search in Google Scholar

410. Botu, V.; Ramprasad, R. Adaptive Machine Learning Framework to Accelerate Ab Initio Molecular Dynamics. Int. J. Quantum Chem. 2015, 115, 1074–1083; https://doi.org/10.1002/qua.24836.Search in Google Scholar

411. Behler, J.; Parrinello, M. Generalized Neural-network Representation of High-dimensional Potential-energy Surfaces. Phys. Rev. Lett. 2007, 98, 146401; https://doi.org/10.1103/physrevlett.98.146401.Search in Google Scholar

412. Gastegger, M.; Behler, J.; Marquetand, P. Machine Learning Molecular Dynamics for the Simulation of Infrared Spectra. Chem. Sci. 2017, 8, 6924–6935; https://doi.org/10.1039/c7sc02267k.Search in Google Scholar PubMed PubMed Central

413. Shen, L.; Yang, W. Molecular Dynamics Simulations with Quantum Mechanics/Molecular Mechanics and Adaptive Neural Networks. J. Chem. Theor. Comput. 2018, 14, 1442–1455; https://doi.org/10.1021/acs.jctc.7b01195.Search in Google Scholar PubMed PubMed Central

414. Unke, O. T.; Chmiela, S.; Sauceda, H. E.; Gastegger, M.; Poltavsky, I.; Schutt, K. T.; Tkatchenko, A.; Müller, K.-R. Machine Learning Force Fields. Chem. Rev. 2021, 121, 10142–10186; https://doi.org/10.1021/acs.chemrev.0c01111.Search in Google Scholar PubMed PubMed Central

415. Zeng, J.; Zhang, D.; Lu, D.; Mo, P.; Li, Z.; Chen, Y.; Rynik, M.; Huang, L. a.; Li, Z.; Shi, S.; Wang, Y.; Ye, H.; Tuo, P.; Yang, J.; Ding, Y.; Li, Y.; Tisi, D.; Zeng, Q.; Bao, H.; Xia, Y.; Huang, J.; Muraoka, K.; Wang, Y.; Chang, J.; Yuan, F.; Bore, S. L.; Cai, C.; Lin, Y.; Wang, B.; Xu, J.; Zhu, J.-X.; Luo, C.; Zhang, Y.; Goodall, R. E. A.; Liang, W.; Singh, A. K.; Yao, S.; Zhang, J.; Wentzcovitch, R.; Han, J.; Liu, J.; Jia, W.; York, D. M.; E, W.; Car, R.; Zhang, L.; Wang, H. DeePMD-kit v2: A Software Package for Deep Potential Models. J. Chem. Phys. 2023, 159, 054801. https://doi.org/10.1063/5.0155600.Search in Google Scholar PubMed PubMed Central

416. Gomez, A.; de la Puente, M.; David, R.; Laage, D. Neural Network Potentials for Exploring Condensed Phase Chemical Reactivity. CR Chim. 2024, 27, 1–17. https://doi.org/10.5802/crchim.315.Search in Google Scholar

417. Shao, Y. H.; Gan, Z. T.; Epifanovsky, E.; Gilbert, A. T. B.; Wormit, M.; Kussmann, J.; Lange, A. W.; Behn, A.; Deng, J.; Feng, X. T.; Ghosh, D.; Goldey, M.; Horn, P. R.; Jacobson, L. D.; Kaliman, I.; Khaliullin, R. Z.; Kus, T.; Landau, A.; Liu, J.; Proynov, E. I.; Rhee, Y. M.; Richard, R. M.; Rohrdanz, M. A.; Steele, R. P.; Sundstrom, E. J.; Woodcock, H. L.; Zimmerman, P. M.; Zuev, D.; Albrecht, B.; Alguire, E.; Austin, B.; Beran, G. J. O.; Bernard, Y. A.; Berquist, E.; Brandhorst, K.; Bravaya, K. B.; Brown, S. T.; Casanova, D.; Chang, C. M.; Chen, Y. Q.; Chien, S. H.; Closser, K. D.; Crittenden, D. L.; Diedenhofen, M.; DiStasio, R. A.; Do, H.; Dutoi, A. D.; Edgar, R. G.; Fatehi, S.; Fusti-Molnar, L.; Ghysels, A.; Golubeva-Zadorozhnaya, A.; Gomes, J.; Hanson-Heine, M. W. D.; Harbach, P. H. P.; Hauser, A. W.; Hohenstein, E. G.; Holden, Z. C.; Jagau, T. C.; Ji, H. J.; Kaduk, B.; Khistyaev, K.; Kim, J.; Kim, J.; King, R. A.; Klunzinger, P.; Kosenkov, D.; Kowalczyk, T.; Krauter, C. M.; Lao, K. U.; Laurent, A. D.; Lawler, K. V.; Levchenko, S. V.; Lin, C. Y.; Liu, F.; Livshits, E.; Lochan, R. C.; Luenser, A.; Manohar, P.; Manzer, S. F.; Mao, S. P.; Mardirossian, N.; Marenich, A. V.; Maurer, S. A.; Mayhall, N. J.; Neuscamman, E.; Oana, C. M.; Olivares-Amaya, R.; O’Neill, D. P.; Parkhill, J. A.; Perrine, T. M.; Peverati, R.; Prociuk, A.; Rehn, D. R.; Rosta, E.; Russ, N. J.; Sharada, S. M.; Sharma, S.; Small, D. W.; Sodt, A.; Stein, T.; Stück, D.; Su, Y. C.; Thom, A. J. W.; Tsuchimochi, T.; Vanovschi, V.; Vogt, L.; Vydrov, O.; Wang, T.; Watson, M. A.; Wenzel, J.; White, A.; Williams, C. F.; Yang, J.; Yeganeh, S.; Yost, S. R.; You, Z. Q.; Zhang, I. Y.; Zhang, X.; Zhao, Y.; Brooks, B. R.; Chan, G. K. L.; Chipman, D. M.; Cramer, C. J.; Goddard, W. A.; Gordon, M. S.; Hehre, W. J.; Klamt, A.; Schaefer, H. F.; Schmidt, M. W.; Sherrill, C. D.; Truhlar, D. G.; Warshel, A.; Xu, X.; Aspuru-Guzik, A.; Baer, R.; Bell, A. T.; Besley, N. A.; Chai, J. D.; Dreuw, A.; Dunietz, B. D.; Furlani, T. R.; Gwaltney, S. R.; Hsu, C. P.; Jung, Y. S.; Kong, J.; Lambrecht, D. S.; Liang, W. Z.; Ochsenfeld, C.; Rassolov, V. A.; Slipchenko, L. V.; Subotnik, J. E.; Van Voorhis, T.; Herbert, J. M.; Krylov, A. I.; Gill, P. M. W.; Head-Gordon, M. Advances in Molecular Quantum Chemistry Contained in the Q-Chem 4 Program Package. Mol. Phys. 2015, 113, 184–215. https://doi.org/10.1080/00268976.2014.952696.Search in Google Scholar

418. Cao, Y. D.; Romero, J.; Olson, J. P.; Degroote, M.; Johnson, P. D.; Kieferová, M.; Kivlichan, I. D.; Menke, T.; Peropadre, B.; Sawaya, N. P. D.; Sim, S.; Veis, L.; Aspuru-Guzik, A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 10856–10915. https://doi.org/10.1021/acs.chemrev.8b00803.Search in Google Scholar PubMed

419. Motta, M.; Rice, J. E. Emerging Quantum Computing Algorithms for Quantum Chemistry. WIRes Comput. Mol. Sci. 2022, 12, e1580; https://doi.org/10.1002/wcms.1580.Search in Google Scholar

420. Kassal, I.; Whitfield, J. D.; Perdomo-Ortiz, A.; Yung, M.-H.; Aspuru-Guzik, A. Simulating Chemistry Using Quantum Computers. Annu. Rev. Phys. Chem. 2011, 62, 185–207; https://doi.org/10.1146/annurev-physchem-032210-103512.Search in Google Scholar PubMed

421. Castaldo, D.; Jahangiri, S.; Delgado, A.; Corni, S. Quantum Simulation of Molecules in Solution. J. Chem. Theor. Comput. 2022, 18, 7457–7469; https://doi.org/10.1021/acs.jctc.2c00974.Search in Google Scholar PubMed PubMed Central

422. Fedorov, D. A.; Otten, M. J.; Gray, S. K.; Alexeev, Y. Ab Initio Molecular Dynamics on Quantum Computers. J. Chem. Phys. 2021, 154, 164103; https://doi.org/10.1063/5.0046930.Search in Google Scholar PubMed

Received: 2025-04-07
Accepted: 2025-05-06
Published Online: 2025-05-23
Published in Print: 2025-09-25

© 2025 IUPAC & De Gruyter

Downloaded on 13.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/pac-2025-0476/html?lang=en
Scroll to top button