57
Subscriber access provided by Basel University Library Journal of Chemical Theory and Computation is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties. Article Soft-sphere continuum solvation in electronic-structure calculations Giuseppe Fisicaro, Luigi Genovese, Oliviero Andreussi, Sagarmoy Mandal, Nisanth N. Nair, Nicola Marzari, and Stefan Goedecker J. Chem. Theory Comput., Just Accepted Manuscript • DOI: 10.1021/acs.jctc.7b00375 • Publication Date (Web): 19 Jun 2017 Downloaded from http://pubs.acs.org on June 26, 2017 Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Subscriber access provided by Basel University Library

Journal of Chemical Theory and Computation is published by the American ChemicalSociety. 1155 Sixteenth Street N.W., Washington, DC 20036Published by American Chemical Society. Copyright © American Chemical Society.However, no copyright claim is made to original U.S. Government works, or worksproduced by employees of any Commonwealth realm Crown government in the courseof their duties.

Article

Soft-sphere continuum solvation in electronic-structure calculationsGiuseppe Fisicaro, Luigi Genovese, Oliviero Andreussi, Sagarmoy

Mandal, Nisanth N. Nair, Nicola Marzari, and Stefan GoedeckerJ. Chem. Theory Comput., Just Accepted Manuscript • DOI: 10.1021/acs.jctc.7b00375 • Publication Date (Web): 19 Jun 2017

Downloaded from http://pubs.acs.org on June 26, 2017

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are postedonline prior to technical editing, formatting for publication and author proofing. The American ChemicalSociety provides “Just Accepted” as a free service to the research community to expedite thedissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscriptsappear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have beenfully peer reviewed, but should not be considered the official version of record. They are accessible to allreaders and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offeredto authors. Therefore, the “Just Accepted” Web site may not include all articles that will be publishedin the journal. After a manuscript is technically edited and formatted, it will be removed from the “JustAccepted” Web site and published as an ASAP article. Note that technical editing may introduce minorchanges to the manuscript text and/or graphics which could affect content, and all legal disclaimersand ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errorsor consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Page 2: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Soft-sphere continuum solvation in

electronic-structure calculations

G. Fisicaro,∗,† L. Genovese,‡ O. Andreussi,¶,§ S. Mandal,‖ N. N. Nair,‖ N.

Marzari,§ and S. Goedecker†

†Department of Physics, University of Basel, Klingelbergstrasse 82, CH-4056 Basel,

Switzerland

‡Laboratoire de simulation atomistique (L_Sim), SP2M, INAC, CEA-UJF, F-38054

Grenoble, France

¶Institute of Computational Science, Universita’ della Svizzera Italiana,Via Giuseppe Buffi

13, CH-6904 Lugano, Switzerland

§Theory and Simulations of Materials (THEOS) and National Centre for Computational

Design and Discovery of Novel Materials (MARVEL), École Polytechnique Fédérale de

Lausanne, Station 12, CH-1015 Lausanne, Switzerland

‖Department of Chemistry, Indian Institute of Technology Kanpur, Kanpur 208016, India

E-mail: [email protected]

Abstract

We present an implicit solvation approach where the interface between the quantum-

mechanical solute and the surrounding environment is described by a fully continuous

permittivity built up with atomic-centered “soft” spheres. This approach combines

many of the advantages of the self-consistent continuum solvation model in handling

solutes and surfaces in contact with complex dielectric environments or electrolytes

in electronic-structure calculations. In addition it is able to describe accurately both

1

Page 1 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 3: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

neutral and charged systems. The continuous function, describing the variation of the

permittivity, allows to compute analytically the non-electrostatic contributions to the

solvation free energy that are described in terms of the quantum surface. The whole

methodology is computationally stable, provides consistent energies and forces, and

keeps the computational efforts and runtimes comparable to those of standard vacuum

calculations. The capabilitiy to treat arbitrary molecular or slab-like geometries as

well as charged molecules is key to tackle electrolytes within mixed explicit/implicit

frameworks. We show that, with given, fixed atomic radii, two parameters are sufficient

to give a mean absolute error of only 1.12 kca/mol with respect to the experimental

aqueous solvation energies for a set of 274 neutral solutes. For charged systems, the

same set of parameters provides solvation energies for a set of 60 anions and 52 cations

with an error of 2.96 and 2.13 kcal/mol, respectively, improving upon previous literature

values. To tackle elements not present in most solvation databases, a new benchmark

scheme on wettability and contact angles is proposed for solid-liquid interfaces, and

applied to the investigation of the stable terminations of a CdS (1120) surface in an

electrochemical medium.

1 Introduction

The computational study of matter in various environments is a continuously growing field

in solid state physics and chemistry. Systems of interest are for instance molecules, clusters

or surfaces in contact with solvents.1 A possibility is to include explicitly in the simulation

domain all the atoms and molecules of the solution. Such an explicit treatment is in principle

the natural way to account for solvent effects in a first-principles scheme. However, this

approach enormously increases the computational cost and limits at the same time the size

of the system contained in the explicit dielectric medium.2 As a consequence, the study of

the solute-solvent interactions at length scales larger than molecular sizes, or the generation

of long molecular dynamics trajectories, require big computational efforts.

2

Page 2 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 4: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Numerically expensive investigations in material science such as structure predictions3

or reaction path determinations4 would become very time-consuming if not altogether unaf-

fordable. During an exploration of the potential energy surface, most of the minima would

arise from rearrangements of solvent molecule. For an hydrated surface, a lot of time would

consequently be spent moving water molecules rather than sampling the solid-water interface

that is of interest. The explicit description of water also suffers from some well-known lim-

itations of current state-of-the-art ab-initio methods, in particular regarding its structural

and dielectric properties.5–8

The alternative is an implicit description of the solvent effects, while still treating the

other parts of the system explicitly on an atomic level. Such an explicit/implicit treatment

requires three main ingredients:

• A dielectric cavity represented by a proper function ǫ(r) mimicking the surrounding

solvent of a solute as a continuum dielectric;

• A solver for the generalized Poisson equation9

∇ · ǫ(r)∇φ(r) = −4πρ(r), (1)

where φ(r) is the potential generated by a given charge density ρ(r);

• A model for the non-electrostatic terms to the total free energy of solvation.

The dielectric function ǫ(r) has to take the value of one where the solute is placed to

solve a vacuum-like quantum problem, and the bulk dielectric constant ǫ0 outside.

Several implicit solvation models have been reported in the quantum chemistry literature,

starting from the earliest work of Onsager10 to the widespread polarizable continuum model

(PCM) developed by Tomasi and co-workers.11–13

In the PCM formulation, the cavity surrounding the solute is described by a sharp and

discontinuous dielectric function ǫ(r), and a polarization charge density is exactly local-

3

Page 3 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 5: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

ized at the interface between the vacuum and the dielectric. In this way, the dielectric

environment is represented by an effective surface polarization charge, reducing the three-

dimensional dielectric problem into a two-dimensional one. In its original formulation this

approaches suffers from discontinuities in the atomic forces and numerical singularities, that

are a consequence of sharp cavities. This drawback hindered a PCM extension to ab-initio

molecular dynamics (MD). Modified versions have been proposed in the literature,14,15 as

well as advanced definitions of the dielectric cavity in terms of an isodensity of the solute

charge density.16,17

Another possibility is to represent the interface between the region inside the cavity

and the implicit solvent region outside the cavity by a smooth function of the solute elec-

tronic density ρ(r).18–20 Non-electrostatic contributions to the solvation free energy can be

expressed as function of the cavity volume and surface area.19,21,22 For a charge-dependent

approach ǫ[ρ](r) the functional derivative with respect to ρ(r) of the Hamiltonian delivers

new additional terms to the Kohn-Sham (KS) potential, both from the electrostatic energy

and the non-electrostatic contributions.22 These formulations are elegant and easy to be

parametrized. The cavity changes self-consistently during the wave function optimization

and forces do not depend explicitly on the shape of the cavity. However, there are a number

of drawbacks. The smooth transition between the quantum and continuum regions gives rise

to an inherently three-dimensional problem, while the self-consistent nature of the approach

requires solving a new problem at each step of electronic optimization: thus the approach is

intrinsically computationally more demanding than the rigid 2D alternatives such as PCM.

Moreover, accuracy might be affected due to the low number of parameters; neutral and

charged molecules cannot be treated simultaneously, as these require different parametriza-

tions,23 which lead to different Hamiltonians; furthermore, it is not possible to locally modify

cavity sizes looking only to the electronic density.

Greater flexibility can be added to these smooth functional approaches by defining them

in terms of a fictitious electronic density, which is kept rigid and defined in terms of pa-

4

Page 4 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 6: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

rameterized functions centered at atomic positions.24 Although this approach may overcome

some of the computational limitations of self-consistent alternatives, its original formulation

only focused on the electrostatic terms of the solvation free energy.

To make an mixed explicit/implicit treatment attractive to material science, it is neces-

sary to fulfill several requirements:

• Accurate forces and a numerical cost comparable to standard vacuum calculations to

make molecular dynamics or extensive potential energy surface (PES) explorations

feasible;

• A small number of cavity parameters, in order to easily fit a new environment or find

the best solvent conditions in a given range;

• Exact treatment of molecular or slab-like geometries, typical of systems plunged in a

solvent;

• Ability to treat neutral and charged molecules simultaneusly, in order to tackle complex

interfaces (e.g. a double layer).

The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22

goes in this direction, satisfying the first three conditions. In the present work an improved

implicit solvation model is proposed. It preserves the flexibility of PCM to locally vary

the cavity’s size around each atom as well as the straightforward definition of the non-

electrostatic contributions to the total energy of the charge-dependent solvation models,

function of the cavity volume and surface.22 The dielectric cavity is based on analytic smooth

spheres centered on each atom, and it is fully continuous from the vacuum-like inner regions

to the external bulk solvent. Because of these features, we named our approach “soft-sphere”

model. The proposed formulation satisfies all the requirements listed above. It gives the same

accuracy for the ionic forces as an ordinary vacuum calculation. Compared to the similar

approache of Sanchez et al.,24 that is based on functionals of a fictitious solute density, our

5

Page 5 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 7: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

definition of the cavity in terms of atom-centered smooth functions simplifies significantly

the calculation of derivatives. In addition, the continuous function, describing the variation

of the permittivity, allows us to compute analytically all the non-electrostatic contributions

to the free energy. Once we fix the atomic radii, a given environment can be fitted with two

parameters in a straightforward way. Details of the model will be presented on Sec. 2.

Our implementation has a small numerical overhead, thanks to an efficient solver for

the generalized Poisson equation,9 based on a preconditioned conjugate gradient (PCG)

algorithm.

The method has been tested within Kohn-Sham density functional theory (DFT) as

implemented in the BigDFT package.25,26 Thanks to the underlining electrostatic solver of

this package,27 the solvation problem for slab or isolated systems is solved with the correct

boundary conditions. An extensive parametrization study has been carried out both for

neutral molecules and ions in water and two non-aqueous solvents. Results are reported in

Sec. 3.

Our tests suggest that our approach can well handle slab-like boundary conditions. In Sec.

4 we propose a benchmarking database for slab like geometries, that is based on the wetting

properties and solvent contact angles of solvated surfaces. Standard protocols parametrize

implicit solvation models on a database of experimental free solvation energies for neutral

and ionic solutes in several solvents.28 However, if applied to solid-liquid interfaces in ma-

terial science, several issues arise. The molecules of the standard benchmark sets have a

limited representation of some elements of the periodic table. Many of these underrepre-

sented elements, such as transition metal atoms, play however an important role in many

materials science problems. Solar-energy harvesting in dye-sensitized29 and hybrid cells30,31

or electro-catalytic water splitting32,33 are just two simple examples where such materials

are extensively exploited.

Finally, as an illustrative application of the approach we present a test case of relevance

in material science, namely the prediction of stable terminations of the CdS (1120) surface

6

Page 6 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 8: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

in an electrochemical medium.

2 Soft-sphere model

The electrostatic solvation energy is defined as the difference between the total energy of a

given atomic system in the presence of the dielectric environment Gel and the energy of the

same system in vacuum G0

∆Gel = Gel −G0. (2)

A full comparison with experimental solvation energies needs the inclusion of non-electrostatic

contributions to obtain the total free energy of solvation ∆Gsol. In this case the main terms

in the solute Hamiltonian, as introduced by PCM,12 are

∆Gsol = ∆Gel +Gcav +Grep +Gdis +∆Gtm + P∆V, (3)

where ∆Gel is the electrostatic contribution, Gcav the cavitation energy, i.e. the energy

necessary to build up the solute cavity inside the solvent medium. Grep is a repulsion term

representing the continuum counterpart of the short-range interactions induced by the Pauli

exclusion principle, whilst the dissociation energy Gdis reflects van der Waals interactions.

The thermal term ∆Gtm accounts for the vibrational and rotational changes and, finally,

P∆V includes volume changes in the solute Hamiltonian.

In the case of a charge density dependent cavity, it has been shown that the non-

electrostatic terms Gcav, Grep, and Gdis can be expressed as linear functions of the “quantum

surface” S and “quantum volume” V of the dielectric cavity.19,22 In particular, Scherlis et

al.19 extended the work of Fattebert and Gygi18 including a cavitation term expressed as a

product of the experimental surface tension of the solvent γ and the quantum surface S of

the solute cavity

Gcav = γS. (4)

7

Page 7 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 9: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Both the quantum surface and volume can be expressed in terms of the dielectric function

ǫ(r). Neglecting as in other models the thermal contribution ∆Gtm to the total energy and

considering that for simulations at standard pressure the term P∆V does not play any role,

the total solvation free energy can be modeled by the following expression22

∆Gsol = ∆Gel + (α + γ)S + βV, (5)

where α and β are solvent-specific parameters which can be fitted to experimental data

such as solvation energies of neutral molecules and ions. The linear relation between the

repulsion energy and the quantum surface follows the idea of traditional PCM approaches,

since in PCM models Grep is proportional to the solute electronic density lying outside the

dielectric cavity34 and this last quantity can be linearly correlated to the cavity surface. The

linear model for the non-electrostatic terms follows the same lines as the solvation models

developed by Cramer and Truhlar,13,35 i.e. the SMx family, where these contributions are

evaluated as a sum over all the atoms, each one proportional to the solvent-accessible surface

area, albeit they use more variables. Although Eq. (5) has been introduced within a charge-

dependent model,22 we will use the same expression for the total solvation free energy for the

soft-sphere model after a suitable generalization of the quantum surface and volume. These

last non-electrostatic contributions do not change during the wavefunction optimization as

they do not depend anymore on the electronic charge density (now the dielectric cavity is

a function of the atomic coordinates). As a consequence only the electrostatic contribution

Gel needs to be computed via an optimization of the electronic wavefunction at the ab-initio

level. Appendix A outlines the main equations that have to be implemented in a DFT code

for the wavefunction optimization procedure in presence of an implicit environment.

The spatially varying dielectric function ǫ[ρ](r) or ǫ(r, Ri) has to meet several condi-

tions:

• Go monotonically from a value of 1 inside the cavity (mimicking vacuum) to the bulk

8

Page 8 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 10: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

dielectric constant ǫ0;

• Be smooth in the transition region to guarantee a proper discretization on a three

dimensional grid;

• Have as many continuous derivatives as possible to avoid convergence and numerical

issues during the iterative wavefunction optimization loop.

The soft-sphere cavity, which satisfies all these requirements, will be presented in Sec. 2.1.

Definitions of the quantum surface and volume will be given in Sec. 2.2.

For a dielectric cavity where ǫ(r, Ri) is an explicit function of the atomic coordinates

Ri, additional contributions arise to the atomic forces with respect to their gas-phase for-

mulation. Formulas for the forces are presented in Appendix A.3. A test for the energy and

force accuracy is reported on Sec. 2.3.

The methodology described in this paper as well as a complementary method based on a

charge density dependent continuum solvation model, developed by Andreussi et al.,22 have

been implemented in the BigDFT software package.

2.1 Definition of the dielectric cavity

We define the soft-sphere cavity by means of atomic-centered interlocking spheres described

by continuous and differentiable h functions which smoothely switch from 0 inside the sphere

to the value 1 outside. Their product, which is still continuous and differentiable in the whole

domain, reproduces the analytic cavity

ǫ(r, Ri) = (ǫ0 − 1)

i

h(ξ; ‖r − Ri‖)

+ 1, (6)

where ǫ0 is the dielectric constant of the surrounding medium, ξ a set of parameters

describing the spheres and ‖r−Ri‖ the distance from the sphere center Ri. The derivative

9

Page 9 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 11: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

of the dielectric function with respect to Ri is

∂ǫ(r, Ri)∂Ri

=

[

ǫ(r, Ri)− 1

h(ξ; ‖r − Ri‖)

]

∂h(ξ; ‖r − Ri‖)∂Ri

(7)

for a given function h. This function can be expressed in terms of the error function

h(ri,∆; ‖r − Ri‖) =1

2

[

1 + erf

(‖r − Ri‖ − ri∆

)]

, (8)

where ri are the sphere radii which depend on the particular atomic species, and ∆ a pa-

rameter (fixed for all atoms) which controls the transition region (≈ 4∆ wide) where the

polarization charge is located. This choice for h guarantees a “soft” dielectric function in

Eq. (6) that is analytic everywhere except at the point r = Ri. With our choice of ∆

the discontinuities in the derivatives are however exponentially small and do not have any

negative effect on our calculations.

Following the lines of all PCM implementations,11,12,35,36 ri are taken to be equal to the

van der Waals radii rvdWi of the corresponding elements scaled by a constant multiplying

factor f (i.e. ri = frvdWi ). Finally, the soft-sphere cavity for a system of atoms with

coordinates Ri is uniquely defined by the set of parameters ǫ0, rvdWi ,∆, f. The solvation

model needs three additional parameters α, β and γ for the non-electrostatic terms to the

total solvation energy [Eq. (5)]. The parametrization of all parameters will be reported on

Sec. 3.

2.2 Geometrical contributions

Eq. (3) allows us to compute the total free energy of solvation. Thanks to Eq. (5) the inte-

gration at DFT level of the non-electrostatic terms becomes straightforward. The evaluation

of the free energy and its functional derivative can be expressed in term of the quantum

10

Page 10 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 12: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

surface and volume. Introducing the function

θ[ǫ](r) =ǫ0 − ǫ(r)

ǫ0 − 1, (9)

which takes a value of one inside the dielectric cavity and zero externally, the quantum

volume V of the solute can be defined as

V [ǫ] =

drθ[ǫ](r). (10)

The integral has to be evaluated in the whole simulation domain. The associated quantum

surface S can be defined from the gradient of θ[ǫ]:21

S[ǫ] =

dr|∇θ[ǫ](r)| = 1

ǫ0 − 1

dr|∇ǫ(r)|. (11)

The functional derivatives of the quantum volume and surface with respect to the dielectric

function are:

δV [ǫ]

δǫ(r)= − 1

ǫ0 − 1, (12)

δS[ǫ]

δǫ(r)=

1

ǫ0 − 1

dr′δ∂iǫ(r

′)

δǫ(r)

∂iǫ(r′)

|∇ǫ(r′)|

=1

ǫ0 − 1

(

∂iǫ(r)∂jǫ(r)∂i∂jǫ(r)

|∇ǫ(r)|3 − ∇2ǫ(r)

|∇ǫ(r)|

)

, (13)

where repeated indices are considered as summed.

Thanks to the continuous and differentiable permittivity ǫ(r), the function θ[ǫ](r) as well

as the quantum volume and surface can be analytically computed. This holds also for their

derivatives of Eq. (11-13).

11

Page 11 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 13: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

2.3 Energy and force accuracy

Accurate energies and forces are fundamental, especially for molecular dynamics and struc-

ture prediction calculations. Smooth and analytical cavities can guarantee accuracy levels

comparable to the ones in vacuum. Appendix A.3 shows the additional terms to the forces

for the soft-sphere model. These can be analytically computed, [see Eq. (48)]. In addition,

it is worth noting that all these analytic contributions to the forces can be computed nonce

once during a wave function optimization.

The extent of the transition region (≈ 4∆ wide) can affect the energy and force accuracy.

A large ∆ can lead to a not-physical superposition of large dielectric cavities, while a small

value can require a high resolution mesh which becomes computationally expensive. As

a consequence a compromise is needed, considering that the more mesh points fall in the

transition region, the more accurate is the description of solvent effects over the solute and

the polarization charge there located. In the following we report a test of the energy and

force accuracy for a water molecule in implicit water. Details of the model’s parameter

setting will be given in Sec. 3. In the present test we used the parametrization reported in

row 3 of Table 1.

Forces can be tested by means of the test-forces tool of BigDFT, an internal program

which compares the variation of the total energy ∆E with the path integral of the forces

−Fa · dra. As consequence of Eq. (46), their difference δ = ∆E +∑

a Fa · dra should be

zero. Due to the finite discretization and the numerical integration, such value depends on

the 3 dimensional spatial grid [hx, hy, hz] of the simulation domain (where the KS problem

is solved). It reaches δ ∼ 10−6 Ha in gas phase for very accurate calculations. Typical values

of hi (with i = x, y, z) are in most cases between 0.3 and 0.6 bohr. Present test runs make

use of an uniform grid spacing in all three directions, i.e. hgrid := hx = hy = hz. Fig. 1

shows the accuracy δ of the force computation as function of ∆. Solid lines represent the

value of δ for a test-forces run of the H2O molecule in vacuum for a given hgrid. The accuracy

of the forces is comparable to that of vacuum runs, for ∆ values between 0.2 and 0.8 a.u.;

12

Page 12 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 14: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

deviations start to take place at ∆ = 0.1 a.u., while smaller values affect the convergence

of the generalized Poisson solver due to the increased sharpness of the transition region. In

Fig. 1b we reported the accuracy of the solvation free energy ∆Gsol, i.e. Eq. (5), for the

same set of calculations. The accuracy is defined as the difference between values extracted

for a given hgrid and the most accurate grid we tested at hgrid = 0.2 a.u.. Varying ∆ the

same accuracy is recovered for hgrid ≥ 0.4 a.u.. An hgrid = 0.4 a.u. guarantees an accuracy

of 10−2 kcal/mol. Similar results obtained with other small organic molecules lead to similar

conclusions. Performance of the soft-sphere model is discussed in Appendix B for the same

system.

3 Parametrization

The soft-sphere solvation model is defined by the parameter set ζ = ǫ0, rvdWi ,∆, f ;α, β, γ,

i.e. six parameters plus an atomic radius rvdWi for each distinct atomic species of the quan-

tum system in contact with the implicit solvent. Only the first four parameters are related

to the dielectric cavity and the electrostatic contribution ∆Gel. To parametrize the solvation

model we need to map the ζ = ǫ0, rvdWi ,∆, f ;α, β, γ parameter space, and minimize the

mean absolute error (MAE) over a given collection of experimental free solvation energies.

In the following we will report the parametrization for a water environment, both for neutral

molecules and ions. In addition, we have parametrized two non-aqueous solvents, namely

mesitylene and ethanol. Structures and experimental data have been retrieved from the

Minnesota Solvation Database, version 2012.28,37–39 Solutes used as benchmark contain at

most the following elements: H, C, N, O, F, Si, P, S, Cl, Br, and I.

Soft norm-conserving pseudopotentials including a non-linear core correction40,41 along

with PBE functional were used to describe the core electrons and exchange-correlation for

all calculations of the present section. The Libxc42 library was exploited for the calculation

of the functionals. Solvation free energies were computed via Eq. (5) using atomic structures

13

Page 13 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 15: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Figure 1: H2O molecule in implicit water described by the soft-sphere model: (a) force accuracy δ

obtained by means of the test-forces tool of BigDFT (see text) as function of the cavity parameter∆ for different grid spacing hgrid (colored solid lines represent the accuracy in vacuum runs for eachhgrid); (b) accuracy of the solvation free energy ∆Gsol with respect to the spacial grid hgrid fordifferent ∆.

14

Page 14 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 16: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

optimized both in vacuum and in the presence of the implicit solvent. BigDFT represents

the wavefunction in a wavelet basis set. These basis functions are centered on a cartesian

mesh of resolution [hx, hy, hz].25 The generalized Poisson equation, i.e. Eq. (1), has been

solved via the recently developed generalized Poisson solver based on Interpolating Scaling

Functions.9 Its preconditioner, based on the BigDFT vacuum Poisson solver, allows to solve

the electrostatic problem in some ten iterations with an exact treatment of free, surface and

periodic boundary conditions.27,43 A new version of the algorithm which includes a proper

use of an input guess is reported in Appendix B. All simulations reported in the present

study use an uniform grid of hx = hy = hz = 0.40 a.u. and free boundary conditions.

3.1 Neutral molecules

The soft-sphere model has been benchmarked on a collection of 274 aqueous solvation free

energies to reproduce a water environment. At first, to speed up the exploration of the

parameter space, we trained the model on a representative subset, which consists of 13

molecules spanning the main functional groups of the entire set.44

3.1.1 Cavity parameters

In order to reduce the size of the parameter space (in principle a MAE should be extracted

for each ζ setup), we proceeded fixing the largest possible number of parameters. Since

we are benchmarking the model on aqueous solvation free energies, the experimental value

at low frequency and ambient conditions of ǫ0 = 78.36 has been used. A critical point in

PCM approaches is the choice of shape and size of the cavity, which should reflect the charge

distribution of the solute. In the first formulation of this model,45 PCM atomic radii ri were

chosen to be proportional to the van der Waals radii.46,47 Importing a consistent set of van

der Waals radii rvdWi which spans the whole periodic table is more meaningful than fitting

radii for all the elements which would presumably lead to some overfitting. We decided to

import rvdWi from the Unified Force Field (UFF) parameterization48 for all elements of the

15

Page 15 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 17: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

periodic table. Hydrogen atoms have been considered explicit in the present work. ∆ has

been fixed to 0.5 a.u. for all atoms, generating a transition region of ≈ 2 a.u.. Different

choices of ∆ will be also investigated and discussed at the end of the paragraph. Finally, only

the multiplying factor f for the atomic radii are left free to vary, reducing the four-parameter

problem for the cavity definition to a single-parameter optimization.

The flexibility to locally change atomic radii has also another advantage. Considering

a wet surface, the radii for atoms which are in the bulk region may be further enlarged to

ensure that the dielectric permittivity is exactly 1 for all the domain which is inaccessible

by the solvent. This condition is not always satisfied by an implicit solvation model and has

to be checked.

Since we imported the non-electrostatic contributions from the sccs approach, we started

the present parametrization from the same non-electrostatic proportional factors (α + γ =

50.0 dyn/cm, β = −0.35 GPa). Finally, we ended up with a single-parameter optimization

for f to define the model. A value of f = 1.12 minimizes the prediction error of the

experimental aqueous solvation energies over the representative set of 13 molecules, giving a

MAE of 1.39 kcal/mol. Exporting the same ζ setup (with the optimized f) to the whole

set of 274 neutral molecules, we obtained a MAE of 1.10 kcal/mol. This demonstrates that

the soft-sphere model gives about the same accuracy as the charge density dependent sccs

model (MAE of 1.14 kcal/mol on the same set of 274 neutrals, see Table 1) and that it

is possible to export the linear approach for the non-electrostatic terms (implemented and

parametrized for the sccs model).

A careful analysis of the error distribution over all functional groups, allowed us to fine

tuning some van der Waals atomic radii rvdWi . We found that a slight variation of the

Nitrogen radius from the UFF48 value 1.83 Å to the one of Bondi46 1.55 Å improves the

whole parametrization. This trend reflects the interaction of the Nitrogen lone pair with

Hydrogen atoms of water, making their reciprocal distances smaller. The parametrization

curve varying f over the 13 molecules has been reported in Fig. (2) (black dots, full line).

16

Page 16 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 18: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Table 1: Mean Absolute Errors in aqueous solvation free energies (kcal/mol) for a set of 274neutral molecules obtained with the soft-sphere solvation model.

atomic radii rvdWi cavity multiplying factor f α+ γ [dyn/cm] β [GPa] MAE [kcal/mol]

UFF48 1.12 50.0 -0.35 1.10

UFFa 1.16 11.5 0.00 1.12

UFFa 1.12 50.0 -0.35 1.03

Pauling47 1.36 50.0 -0.35 1.17

Bondi46 1.32 50.0 -0.35 1.06

a Van der Waals atomic radii from the Unified Force Field setup48 with Bondi’s radius46 for Nitrogen.

The MAE minimum is located again at f = 1.12 with a MAE of 1.15 kcal/mol for the 13

molecules. Applying the new setup to the full set of 274 molecules gives a MAE of 1.03

kcal/mol (see Table 1). This remarkable result has been achieved with the crude import of

the linear non-electrostatic model from sccs. The inset of Fig. 2 reports a global comparison

between experimental and ab-initio aqueous solvation free energies for the full set of 274

neutrals. All dots stay close to the diagonal, reflecting the accuracy of the soft-sphere model

over all functional groups.

PCM formulations in the literature mainly use van der Waals radii from Pauling47 or

Bondi46’s work. Even though these sets work well for organic molecules, values for the full

periodic table are not available, making the UFF setup more tempting for general ab-initio

calculations. Nevertheles we explored the performance of our solvation model with the van

der Wall radii of Pauling and Bondi’s. Following the same procediure as in our the previous

optimization, we first performed an optimization of the cavity multiplying factor f over the

representative set of 13 molecules, and only afterwards for the entire ζ setup of 274 neutral

molecules. Both f and MAEs for the full set are reported in Table 1. The optimized f for

both sets is larger than the one obtained with the UFF implementation. This is related to

Hydrogen radius, being 1.20 Å in the Pauling and Bondi’s work, and 1.443 Å in the UFF. The

overall accuracy does not improve with respect to the UFF parametrization, where MAEs

of 1.17 and 1.06 kcal/mol have been found for Pauling and Bondi’s radii, respectively.

17

Page 17 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 19: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Figure 2: Mean Absolute Error in aqueous solvation free energies for the representative set of 13neutral molecules as function of the cavity multiplying factor f (UFF radii rvdWi

48 with Bondi’sradius46 for Nitrogen, α + γ = 50.0 dyn/cm and β = −0.35 GPa). The inset reports experimentaland ab-initio aqueous solvation free energies for the full set of 274 neutral molecules with the sameparameters setup (optimized f = 1.12).

So far, all parameters related to the dielectric cavity have been optimized, except ∆. A

window of 0.5 < ∆ < 0.65 does not change the results in term of accuracy (increasing ∆

request a proportional increase of f to recover the same MAE).

3.1.2 Non-electrostatic parameters

The last optimization step treats the non-electrostatic parameters α, β and γ. Assuming a

surface tension of γ = 72 dyn/cm for water at room temperature, only α and β need to be

optimized. The sum α + γ can be considered as a single parameter since both variables are

multiplying factors of the quantum surface S in Eq. (5). We extrapolated all the β and α+γ

couples which minimize the MAE over the representative set of 13 molecules (determining

also the optimized f for each fixed couple). As noticed in Ref. [ 22], we found an almost

perfectly linear behavior between the error, i.e. the MAE, as function of β and α+ γ. This

was probably due to the small variation in size of the molecules considered in the training

18

Page 18 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 20: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

set, which makes the volume and the surface terms to be linearly related. All MAE minima

lie approximately on the same line β = c(α + γ) + d, where c = −9.1 nm−1 and d = 0.1045

GPa. The best parametrization still correspond to the sccs optimal setup with α+ γ = 50.0

dyn/cm and β = −0.35 GPa.

However, given the strong correlation between the two tunable parameters, it is important

to stress that the optimal parametrization reported above may be strongly linked to the type

of systems (small isolated molecules) under study. Whether these parameters could be used

in systems with very different sizes, e.g. large molecular and supramolecular systems or

solids described using a slab geometry, is not obvious. For this reason, unless a more refined

optimization that takes into account experimental data on large systems is performed, a

parametrization only relying on the surface term, thus putting β = 0 GPa, should be more

transferable and preferable for the above applications. For surface calculations, β = 0 GPa

is mandatory to avoid an unphysical dependence of the energy on the slab volume.

3.1.3 Results

The final MAEs for both surface (row 2) and small cluster (row 3) parametrizations over

the full set of 274 experimental solvation energies are 1.12 and 1.03 kcal/mol, respectively

(see Table 1). Aqueous solvents are widely used and of great relevance in wet chemistry.

Investigations may require a mixed explicit/implicit treatment of a water environment. The

soft-sphere model predicts a solvation free energy for H2O and (H2O)2 in implicit water of

−5.74 and −10.19 kcal/mol for the surface parametrization, and −5.60 and −10.34 kcal/mol

for the small clusters one. Experimental values are −6.31 and −11.27 kcal/mol, respec-

tively.28 Predictions are within 1.1 kcal/mol and both parametrizations perform pretty well.

Marenich et al. in Ref. [ 38] explored performances of several implicit solvation models.

These approaches hold as common feature the use of dozen of parameters to model all

contributions to the solvation free energy. Among others, they tested the SM838 model of the

SMx family, the Integral Equation Formalism Polarizable Continuum Model45 of Gaussian

19

Page 19 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 21: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

03 (IEF-PCM),49,50 the Conductor-like PCM model51 in GAMESS52 (C-PCM/GAMESS),

Jaguar’s Poisson-Boltzmann self-consistent reaction field solver53,54 (PB/Jaguar) and the

Generalized Conductor-like Screening Model in NWChem55 (GCOSMO/NWChem). MAEs

of these methods for aqueous solvation free energies are been reported in Table 2. for the

set of 274 neutrals Data from the recent SM12 implementation39 have been also included.

In addition we tested the charge-dependent sccs model implemented within the BigDFT

package. Using only two parameters (surface parametrization), the soft-sphere model applied

to the same dataset lies in the same range of accuracy. A carefully tuning of atomic radii

and/or a functional dependence of delta over the periodic table could further lower the MAE

and improve the whole accuracy.

Table 2: Mean Absolute Errors in aqueous solvation free energies (kcal/mol) for several solvationmodels (MAEs from Ref. [ 38]). Model benchmarks refer to same set of 274 neutrals, 60 anionsand 52 cations of the Minnesota Solvation Database, version 2012.28

Method neutrals cations anions

soft-spherea 1.12 2.13 2.96

sccs23 1.14b 2.27c 5.54c

SM838 0.55 2.70 3.70

SM1239 0.59 2.90 2.90

PB/Jaguar38 0.86 3.10 4.80

IEF-PCM38 1.18 3.70 5.50

C-PCM/GAMESS38 1.57 7.70 8.90

GCOSMO/NWChem38 8.17 11.00 7.00

a parametrization of row 2 Table 1; b sccs implemented in BigDFT; c sccs for ions corresponds to a

reduced set of 55 anions and 51 cations23 of the same Minnesota data set.

3.2 Anions and cations

In order to strictly validate the soft-sphere model, we benchmarked it on a set of aqueous free

energies of solvation for 112 singly-charged ions (60 anions and 52 cations). The Minnesota

Solvation Database, version 2012,28,37,39 contains two sets of them, i.e. the unclustered and

20

Page 20 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 22: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

the selectively clustered data set. The last is identical to the first, except for 31 ions which are

clustered with an explicit water molecule. More details can be found in the article discussing

the SM6 parametrization.37 As done in the last articles of the SMx family,38,39 and in order

to make comparisons with others solvation models, we benchmarked the soft-sphere model on

the selectively clustered data set. The two main model parametrizations have been explored

in row 2 and 3 of Table 1, except for the radii multiplying factor f which was left free to be

optimized.

Due to the nonlinear effects in the polarization of the surrounding dielectric with high

fields, the factor f , or alternatively all the radii of atoms bearing the ionic charge, need

to be reconsidered.11,12 Furthermore, ab-initio gas-phase optimizations and Monte Carlo

simulations of systems in aqueous solution suggested that solvent molecules of the first

solvation shell come closer to a charged molecule than to a neutral.56 Although most PCM

models implement a charge dependency on the atomic radii,11,57 we decided to reproduce

experimental solvation energies optimizing the global cavity factor f both for anions and

cations. A similar approach has been also investigated by others authors.56 This procedure

is less accurate than modifying single atomic radii, especially when the solute molecule sizes

become larger or for extended interfaces. However, it is still a good approximation for small

size ions. A more detailed approach with a fine tuning of the individual van der Waals radii

and their dependence on the local atomic charges is beyond the scope of the present paper.

Table 3 reports MAEs both for anions and cations. Fixing the non-electrostatic parame-

ters to α+γ = 50.0 dyn/cm and β = −0.35 GPa, the optimization procedure on f suggested

an f = 0.98 as optimal multiplying factor for anions and f = 1.10 for cations, with MAEs

of 3.05 and 2.07 kcal/mol, respectively. These results are in perfect agreement with similar

findings obtained in the parametrization of self-consistent cavities.23 Fig. 3 reports experi-

mental and ab-initio aqueous solvation free energies both for anions (empty blue circles) and

cations (full black circles).

Concerning cations, dots belonging to the diagonal are mainly cations containing Nitro-

21

Page 21 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 23: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

gen. On the other side, the model tends to underestimate free solvation energies for four

different types of alcohols, two types of ketons and the water cation. This trend suggests

that smaller cavities are needed for these cations due to the stronger electrostatic interaction

between the solvent and the solute. Excluding these last seven solutes, the overall MAE for

the remaining 45 cations decreases from 2.07 to 1.33 kcal/mol. A factor of f = 1.04 allows

to improve the prediction for the 7 units (alcohols, ketons and the water cation) where their

MAE is 2.07 kcal/mol. This fine analysis has been conducted to identify margins for im-

provements. Similar arguments are valid for anions, but a systematic study of functional

groups is less straightforward since oxygen comes into play in the majority of the units. An

optimization of f keeping fixed α + γ = 11.5 dyn/cm and β = 0 GPa, leads to a similar

accuracy for the prediction of solvation free energies (see Table 3). MAEs are 2.96 and 2.13

kcal/mol for anions and cations, respectively.

We see that the soft-sphere model performs very well for ions in water, considering that

the estimated experimental average uncertainty for solvation free energies of ionic solutes

is 3 kcal/mol.37 It is worth noting the high accuracy has been reached, by only optimizing

one parameter, i.e. f , with respect to the parametrization for neutral solutes. Table 2

reports MAEs for the same set of ions in water obtained with others solvation models. All

data refer to the same selectively clustered data set of 60 anions and 52 cations of the

Minnesota Solvation Database, version 2012, except sccs which corresponds to a reduced set

of 55 anions and 51 cations23 of the same database. The soft-sphere model outperforms all

current implicit solvation models, with a MAE of 2.57 kcal/mol over the full ion dataset. The

possibility to use the same α, β and γ and to change only the radii upon ionization allows

to tackle systems which contains simultaneously neutral and charged complexes. This is not

possible in charge density dependent approaches which can not distinguish between atoms

and need different non-electrostatic parametrizations (and then different Hamiltonians) for

neutrals and ions.

Since both parametrizations of Table 1, i.e. row 2 and 3, are able to accurately reproduce

22

Page 22 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 24: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

solvation energies of neutrals and ions, and taking into account that surface calculations

require β = 0 GPa and that reducing the parameter space is always advantageous, we can

just use the quantum surface S in the model of non-electrostatic contributions in Eq. (5).

The whole solvation model is thus uniquely defined by two parameters, i.e. f and α.

Figure 3: Experimental and ab-initio aqueous solvation free energies for a set of 112 single-chargedions, of which 60 anions (empty blue circles) and 52 cations (full black circles), obtained with thesoft-sphere model (UFF radii rvdWi

48 with Bondi’s radius46 for Nitrogen, α+ γ = 50.0 dyn/cm andβ = −0.35 GPa, f = 0.98 for anions and 1.10 for cations).

Table 3: Mean Absolute Errors in aqueous solvation free energies (kcal/mol) for a set of 112 ions(60 anions and 52 cations) obtained with the soft-sphere modela .

cavity multiplying factor f α+ γ [dyn/cm] β [GPa] MAE [kcal/mol]

Anions 0.98 50.0 -0.35 3.051.00 11.5 0.00 2.96

Alcohols and ketons cations 1.04 50.0 -0.35 2.071.04 11.5 0.00 1.88

Cations without alcohols and ketons 1.10 50.0 -0.35 1.331.11 11.5 0.00 1.46

All Cations 1.10 50.0 -0.35 2.071.10 11.5 0.00 2.13

Total ions - 50.0 -0.35 2.60- 11.5 0.00 2.57

a

UFF radii rvdWi

48 with Bondi’s radius for Nitrogen.

23

Page 23 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 25: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

3.3 Non-aqueous solvents

The study of chemical reactions within non-aqueous solvents is as also important. Therefore

it is worth investigating how the parametrization has to be modified for a non-aqueous

solvent. As test cases, we took two solvents, namely mesitylene and ethanol, which have low

dielectric constants ǫ0 of 2.2658 and 24.85, respectively at 20 C.

To optimize the soft-sphere parameters ζ = ǫ0, rvdWi ,∆, f ;α, β, γ, experimental sol-

vation free energies for a set of seven (mesitylene) and eight (ethanol) organic molecules

have been taken from the Minnesota Solvation Database.28 Changing only ǫ0 to the exper-

imental value of the non-aqueous solvent and taking all others parameters from the water

parametrization (row 2 or 3 of Table 1) does not reproduce the experimental data, underesti-

mating all solvation free energies with an overall MAE of 3.42 and 2.84 kcal/mol for mesity-

lene and ethanol, respectively. Turning off the non-electrostatic contributions (α + γ = 0

dyn/cm, β = 0 GPa), and using the proper ǫ0 has the same effect with MAEs of 3.42 and 1.74

kcal/mol. As a consequence each solvent needs its own parametrization. PCM approaches

usually have solvent-dependent parameters, like the multiplying factor for the atomic radii

or several collections of parameters for the non-electrostatic energetic.

We therefore set in addition the values rvdWi for the radii to the UFF values48 with the

exception of Nitrogen where we use the Bondi’s radius and used again ∆ = 0.5 as we did for

water. The surface tension at 20 C of mesitylene is γ = 28.80 dyn/cm,59 whilst for ethanol

is γ = 22.10 dyn/cm. Since solvation energies are well reproduced with only the surface

term in Eq. (5), we end up with a two-parameters optimization, i.e. f and α. First we

optimized α, keeping f to the water value of f = 1.16. This provides the major correction

to the overall MAE. Then f and α have been fine tuned concurrently. A value of f = 1.22

and α+ γ = −12.0 dyn/cm provide a MAE of 0.71 kcal/mol for mesitylene. Doing the same

for ethanol, a f = 1.22 and α+ γ = −4.0 dyn/cm allow to reproduce solvation free energies

with a MAE of 1.28 kcal/mol. Accuracies are comparable to the SM8 model, which gives a

MAE of 0.40 (mesitylene) and 1.53 (ethanol) kcal/mol for the same data set of neutrals.38

24

Page 24 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 26: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Hence our approach has an accuracy comparable to similar state-of-the-art methods, despite

the reduced number of parameters. Nonetheless, it is important to stress that the reported

parameterizations may be affected by the reduced size of the fitting set: the limited numbers

and the narrow ranges of experimental solvation free energies used for the fit may give rise

to a parameterization which is not transferable to a broader set of compounds.

This is particularly striking in a model which only relies on two tunable parameters, for

which the optimized non-electrostatic terms end up being always negative. Thus, similarly

to the electrostatic term, these terms contribute to a stabilization of the solute, excluding

the possibility of modeling insoluble compounds with positive solvation free energies. On

the other hand, the optimization procedure gave a prefactor for the atomic radii f greater

than the corresponding value in water. This trend is in agreement with the larger sizes

of mesitylene and ethanol molecules with respect to H2O, and meets previous studies of

solvated molecules where the cavity is defined as a convolution of both solute and solvent

electronic densities.60 Including an optimization on β or additional dependences for the non-

electrostatic terms on specific solvent descriptors like the refractive index and the Abraham’s

hydrogen bond acidity and basicity parameters,38 can further improve the prediction power

on non-aqueous solvents and make the soft-sphere model universal, where “universal” means

applicable to all solvents.

4 Solid-liquid interfaces and contact angle

As a further benchmark system for implicit solvation models we consider the contact angle

θC that a liquid drop on a surface forms with the surface plane. It is experimentally accessi-

ble quantity and a collection on several substrates can be used as test set. Sessile-drop and

captive-bubble techniques are common experimental reference methods.61 This benchmark-

ing approach is alos useful for the parametrization of elements of the periodic table that are

not represented in standard databases of experimental solvation energies.

25

Page 25 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 27: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Although the idea of a direct comparison of predicted and experimental contact angles

seems straightforward, its is not technically challenging. Clean surfaces do not exist in na-

ture where reconstructions, adsorbed molecules or radicals, and defects usually take place.

An implicit solvation approach gives results for ideal surfaces whereas experimental measure-

ments may be influenced by all these deviations from the ideal behaviour. Other simulation

techniques could be applied where both the surface and the liquid are treated explicitly (i.e.

molecular dynamics, Monte Carlo methods or others62). The proposed benchmark provides

necessary conditions that an implicit scheme has to fulfill if applied to a solid-liquid interface.

Simple conditions have to be verified like partial or complete wetting, and the degree of the

solid-liquid interaction by approximately comparing θC.

The contact angle quantifies the wettability of a surface and reflects the equilibrium

between liquid (drop), solid (surface) and vapor-phase interactions. Looking to the contact

line at the surface and imposing the thermodynamic equilibrium between surface tensions of

the three phases, a relation can be deduced for the contact angle in terms of the interfacial

energies solid-vapor γSG, solid-liquid γSL, and liquid-vapor γLG (i.e. the surface tension)

cos θC =γSG − γSL

γLG

. (14)

This relation is known as the Young’s equation.63,64 Since the three surface energies at equi-

librium form the side of a triangle, partial wetting occurs only when the triangle inequalities

γij < γjk + γik is satisfied.

θC can easily be related to the work of adhesion of the solid and liquid phases in contact,

i.e. WLS = γSG + γLG − γSL, leading to the Young-Dupré equation65 WLS = γLG(cos θC + 1).

The work of adhesion is the work which must be done to separate two adjacent materials at

their phase boundary (liquid-liquid or solid-liquid) from one another. Conversely, it is the

energy which is released in the process of wetting and is a measure of the strength of the

contact between two phases. γSG and γLG are the energies required to create a new surface

26

Page 26 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 28: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

whilst γSL represents the work that has to be done in order to form the interface. These

quantities are difficult to determine experimentally and only θC and γLG can be directly

determined. Contact angle measurements can give access to γSG and γLG indirectly via the

Young equation. In this case γSL is assumed to be a function of γSG and γLG. Several

functional relations are present in the literature,66–69 taking into account the polar or non-

polar character of the solvent as well as its acid or basic nature.70 In order to classify

hydrophobic and hydrophilic surfaces, van der Oss exploited the concept of free energy of

solvation ∆GSL both for molecules and condensed matter systems.71 It can be related to the

work of adhesion being ∆GSL = −WLS

Another parameter, especially useful to characterize hydrophilic surfaces, is the spreading

coefficient S defined as S = γSG − γLG − γSL. It represents the work performed to spread

a liquid over a unit surface area of a clean and non-reactive solid (or another liquid) at

constant temperature and pressure and in equilibrium with liquid vapor. Coupled with the

Young’s equation it leads to S = γLG(cos θC − 1), which means that a liquid drop partial

wets the surface only for S < 0, while a total wetting occurs for S > 0. Zero wetting takes

place when WLS < 0. These relations can be made explicit by rewriting cos θC as

cos θC =WLS

γLG

− 1 =S

γLG

+ 1. (15)

It is also clear that WLS > γLG implies that θC < 90, i.e. the surface has an hydrophilic

character, whereas it is hydrophobic in the opposite case. Furthermore, both WLS and

S should be similar in magnitude to γLG to achieve partial-wetting conditions (i.e. 0 ≤

WLS/γLG ≤ 2, −2 ≤ S/γLG ≤ 0). Once θC is experimentally determined, these conditions

need to be fulfilled, providing useful fixed points to benchmark a solvation model.

For a slab, the DFT surface energy γSG is calculated according to

γSG =1

2Alim

N→+∞

(

ENslab-SG −NEbulk

)

(16)

27

Page 27 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 29: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

where N is the number of atoms within the slab, A is the slab’s cross section area, ENslab-SG

is the total energy of the N -atoms relaxed slab, and Ebulk the energy of the bulk system per

atom. The solid-liquid interface energy γSL can be easily computed by means of a similar

equation

γSL =1

2Alim

N→+∞

(

ENslab-SL −NEbulk

)

, (17)

considering ENslab-SL as the total energy of the slab in contact with the implicit solvent. This

approach gives direct access to γSL and it is not necessary to specify an analytical functional

dependence of γSL in terms of γSG and γLG in order to estimate the contact angle via Eq.

(14).

We determined spreading the coefficient S, the work of adhesion WLS and the contact

angles θC for a collection of surfaces in contact with water. Table 4 reports computed data for

a silver (001) surface, a cleaved and reconstructed (001) surface of SiO2 α-quartz72 and two

carbon-based surfaces, i.e. diamond (001) and graphene. Thanks to the implemented UFF

tabulation of van der Waals radii, our soft-sphere model is able to handle such materials. We

used the parametrization reported in the second row of Table 1, valid for surface calculations

(β = 0 GPa). It is worth noting that additional benchmark procedures have to be carried out

for the cavity parameters ζ = ǫ0, rvdWi ,∆, f ;α, β, γ to guarantee a consistent physical

description of the solid-liquid interface. In particular van der Waals radii rvdWi for elements

not present in the regular set of organic molecules have to be monitored. Small values can

lead to no-physical dissolution of surface atoms.

The computation of Ebulk is not necessary because θC, WLS and S are all functions of

the difference γSG − γSL, and Ebulk of Eq. (16,17) is cleared. From this side, cos θC can be

seen as the solvation energy per unit area ∆Gsolslab = (EN

slab-SL −ENslab-SG)/2A divided by γLG,

namely

cos θC = −Gsolslab

γLG

. (18)

For water at room temperature γLG has the value 72 dyn/cm. The parameters for the DFT

28

Page 28 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 30: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

calculations (grid spacings, Monkhorst-Pack k -point mesh) have been chosen in BigDFT

to reach a surface energy accuracy of 0.01 J/m2. All systems have been relaxed both in

vacuum and in the presence of the implicit aqueous environment until all the forces were less

than 5 meV/Å. BigDFT allows to use exact surface boundary conditions, avoiding spurious

interactions in the direction orthogonal to the surface. As an additional check, we compared

vacuum surface energies [Eq. (16)] with DFT literature data, recovering a good agreement

for all systems.

Table 4: Simulated work of adhesion WLS, spreading coefficient S and contact angle θC for severalsurfaces in contact with implicit water described by the soft-sphere model.

slab layers WLS S θC

[mJ/m2] [mJ/m2] [degree]

silver (001) 8 479 33 -

SiO2 α-quartz cleaved (001) 18 181 35 -

SiO2 α-quartz reconstructed (001) 27 76 -69 87

Diamond (001) 12 69 -76 92

Graphene 1 62 -82 97

Unless an oxide layer is present, clean metal surfaces are hydrophilic meaning that wa-

ter completely or nearly completely wets them (gold, silver, copper and others),61,73 The

positive value of S for clean (001) silver surface (see Table 4) predicted by the soft-sphere

model agrees with its hydrophilic character, allowing for a complete wetting of its surface.

A similar validation can be extended to other metallic materials and represents a prelimi-

nary checkpoint if implicit solvation approaches need to be integrated in such solid-liquid

investigations.

Furthermore, as a consequence of their bulk bonding strength, surfaces can also be divided

into high- and low-energy surfaces, The stronger the chemical bonds are, i.e. the higher the

surface energy is, the more easily complete wetting is reached. In our set of surfaces, the

cleaved SiO2 α-quartz surface falls in this last class. Among the different SiO2 highly-reactive

surfaces, the (001) was chosen because it is the most stable.72 Applying Eq. (16) we recovered

29

Page 29 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 31: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

a surface energy of γSG = 2.22 J/m2. This value is consistent to the one reported in Ref. [

72], from where the cleaved and reconstructed SiO2 structures come from. Implicit solvation

correctly predicts its hydrophilic character with a spreading coefficient S > 0 as reported in

table 4. Conversely, the reconstructed SiO2 α-quartz (001) has a γSG o 0.39 J/m2. Because of

its negative spreading coefficient, implicit solvation predicts a partial wetting for this surface

with a contact angle of θC = 87 and a work of adhesion of WLS = 76 mJ/m2. Therefore

the implicit procedure is able to predict different θC for high and low energy surfaces of

the same material. A direct comparison with experimental contact angles is difficult since

real surfaces undergo reconstruction as well as hydroxylation, contamination and patterning

concurrently.74

Similar considerations hold for the clean (001) diamond surface. Molecular dynamics and

Monte Carlo simulations predicted an hydrophobic character for an hydrogen-terminated

diamond slab.62,75 In our case a work of adhesion of 69 mJ/m2 has been obtained, together

with a θC = 92 degree. Remaining within the carbon-material class, we applied the implicit

solvation scheme to a graphene sheet. Graphene is a two dimensional arrangement of carbon

atoms, bonded covalently to form an honeycomb lattice due to the sp2 orbital hybridization.

Weak van der Waals interactions take place with water molecules once graphene is in contact

with an aqueous solution, allowing for a partial wetting of the surface. Several experimental

and simulation studies investigated the wetting features of such carbon allotropes. For

graphene the measured a contact angle is 127,76 although a theoretical study demonstrated

a discrepency with the contact angle of graphite,77,78 suggesting a θC of the order of 95-100.79

For the partial wetting behavior, data extracted with our soft-sphere solvation scheme is in

agreement with these experimental and theoretical studies, predicting a negative spreading

coefficient, a contact angle of 97 and a work of adhesion of 62 mJ/m2. A lowering of the

cavity prefactor to f = 1.12 does not affect the whole wetting predictions for all studied

solid-liquid interfaces in terms of the hydrophobic/hydrophilic character.

30

Page 30 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 32: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

5 Structure of CdS (1120) surface in electrochemical me-

dia

As an application, we investigated the structure of the CdS (1120) surface in contact with

water at different pH and bias potentials, mimicking electrochemical conditions. CdS is

among the few promising photocatalytic materials for the water splitting reaction, due to

its appropriate band gap and band positions.80–82 Recently, it was reported that nanorods

and other nanoporous materials based on CdS show very high photon-to-hydrogen conversion

efficiency when used together with hole-scavenger molecules.83–85 It was observed experimen-

tally that a high pH is necessary for the efficient photo-catalytic activity of CdS.84 However,

photocorrosion was seen after prolonged photo-exposure at such pH levels.84,85 Understand-

ing the structure and reactivity of the CdS surfaces at electrochemical conditions can provide

valuable informations on improving the CdS-based photocatalytic systems.

We employed here the ab-initio thermodynamics based “computational hydrogen elec-

trode” (CHE) approach86,87 to study the structure of the most exposed CdS (1120) surface

at different pH and bias potentials. Solvent effects have been included using the implicit

soft-sphere solvation model. Most of the CHE computations in literature were carried out

under in vacuo conditions,87–92 and a qualitative understanding of solvent effects on the

results of such computations remain unclear.93–96 Recently the CHE approach was used to

study the structure of CdS surfaces in contact with vacuum.97

Here we have computed the formation free energies of various surface terminations (O,

OH, OOH), varying pH and bias potentials. The following reactions were considered for the

formation of these terminations:

2H2O∗ → 2O∗ + 4H+ + 4e−, (19)

2H2O∗ → 2HO∗ + 2H+ + 2e−, (20)

31

Page 31 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 33: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

2H2O∗ + 2H2O → 2HOO∗ + 6H+ + 6e−. (21)

The above equations include two moles of H2O∗ as there are two adsorption sites in the

surface model that we have considered. Here O∗, HO∗, HOO∗ and H2O∗ represent O, HO,

HOO and H2O terminated surfaces, respectively. Following Ref. [ 86], the free energy

change ∆G of the above reactions with reference to a standard hydrogen electrode (SHE)

at pH2= 1 bar and T = 298 K as a function of bias potential U and pH can be calculated,

respectively, as

∆G2O∗ = ∆E2O∗ +∆EZPE2O∗ − T∆S2O∗

+4∆G(pH)− 4eU, (22)

∆G2HO∗ = ∆E2HO∗ +∆EZPE2HO∗ − T∆S2HO∗

+2∆G(pH)− 2eU, (23)

∆G2HOO∗ = ∆E2HOO∗ +∆EZPE2HOO∗ − T∆S2HOO∗

+6∆G(pH)− 6eU. (24)

Here ∆EZPE and T∆S are the zero-point energy correction and the vibrational entropic

contribution to the free energy, respectively, and ∆G(pH) = −kBTpH ln 10. ∆E2O∗ , ∆E2HO∗

and ∆E2HOO∗ are calculated from DFT calculations as follows:

∆E2O∗ = E2O∗ + 2EH2− E2H2O∗ , (25)

∆E2HO∗ = E2HO∗ + EH2− E2H2O∗ , (26)

∆E2HOO∗ = E2HOO∗ + 3EH2− E2H2O∗ − 2EH2O. (27)

32

Page 32 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 34: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

The energies of the surface-adsorbed structures were calculated with and without the implicit

solvation model.

Preliminary calculations showed that a slab built up from an 1×1 supercell with 8 atomic

layers, i.e. Cd16S16, and a k-point sampling 5×1×5 adequately converged the energy (see

the Supporting Information). Zero-point energy and vibrational entropic contributions were

computed from the harmonic frequencies, extracted by means of a finite difference method.

For these last two contributions, a 2×1×2 k-points mesh guarantees reasonable convergence.

Further, we found that both ∆EZPE and T∆S were nearly the same with and without con-

tinuum solvent (see the Supporting Information). Thus, we have not included solvent effects

while computing these quantities. All computations were performed with spin-polarized

DFT in a wavelet basis uising the PBE98 exchange-correlation functional as implemented

in the BigDFT program.25,26,42 Surface boundary conditions were employed for solving the

Poisson equations.43 The real space grids along all the three directions were set to 0.45 Bohr.

Soft norm-conserving pseudopotentials including non-linear core correction40,41 were used to

describe the core electrons. For the dielectric cavity, a van der Waals radius for Cd of 1.5846

Å has been used, since the UFF value 1.424 Å leads to the surface dissolution. The implicit

model predicts a spreading coefficient S of 28 mJ/m2 for the CdS (1120) surface in contact

with water, confirming its hydrophilic character (S > 0).

Different contributions to ∆G at standard conditions (with reference to SHE) are given in

Table 5. The zero-point energy and the entropy of H2(g) and H2O(g) were taken from an ex-

perimental database.99 GH2O(l) was computed asGH2O(g)(298 K, 1 bar)−∆Gv(298 K, 1 bar),100

where ∆Gv is the free energy of vaporization, equal to 0.09 eV computed based on an exper-

imental database99 (see also the Supporting Information). We have considered adsorption

at both surface Cd and S sites for all the adsorbates. It was found that except for oxygen,

all the other adsorbates prefer to bind on Cd sites, consistent with previous works.97

In Fig. 4 we show the free energy change ∆G for different surface terminations as a

function of bias potential U at pH = 0 and pH = 14. We report results both for pure gas-

33

Page 33 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 35: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

phase conditions (solid lines) and in the presence of an implicit water environment (dotted

lines). For small bias potentials, the most stable surface is H2O terminated, where molecular

water is adsorbed on surface Cd atoms at both pH values. The HOO terminated surface

is stable above 3.12 V (2.53 V) bias potential with (without) implicit water at pH = 0,

whilst in the range from 0.93 V (1.24 V) to 3.12 V (2.53 V) the O terminated surface is

thermodynamically preferred under these conditions. Similar conclusions hold for pH = 14.

The inclusion of the solvent interaction results in a substantial shift of the relative stability

with respect to the bias potentials. For e.g., the threshold bias for observing HOO terminated

surface has increased by nearly 0.6 V by including the implicit solvent in the calculations.

Table 5: Contributions to the free energy change ∆G computed at 298 K and 1 bar with referenceto the standard hydrogen electrode for O, HO and HOO terminated CdS (1120) surfaces. ∆ESolvent

and ∆E are the change in energy calculated with and without implicit solvent, respectively [Eqs.(25-27)].

∆E [eV] ∆ESolvent [eV] ∆EZPE [eV] T∆S [eV]

2O∗ (S) 6.39 5.14 -0.63 0.79

2HO∗ (Cd) 5.38 4.73 -0.40 0.56

2HOO∗ (Cd) 10.96 10.87 -0.82 0.10

6 Conclusions

A soft-sphere solvation model for ab-initio electronic-structure calculations has been devel-

oped, parametrized and tested. The transition from the dielectric cavity to outside is contin-

uous and differentiable, allowing for the analytical calculation of the additional terms to the

forces as well as the cavity-dependent non-electrostatic contributions to the total energy that

are described in terms of the quantum surface. As consequence we could obtain very high

accuracy on energies and forces. The computational cost does not increase significantly with

respect to standard gas-phase calculations. We expect that these features make our methods

useful for numerous applications in material science. Fixed the atomic radii, only two fitting

34

Page 34 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 36: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Figure 4: Free energy change ∆G as a function of the bias potential U for O, HO, HOO andH2O terminated CdS (1120) surfaces at pH = 0 and pH = 14. Results obtained in pure gas-phaseconditions (solid lines) and with the presence of an implicit water environment (dotted lines) areshown. Various threshold bias potentials are marked.

35

Page 35 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 37: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

parameters are needed to define the model, independently of the cluster or slab geometry.

Mean absolute errors with respect to experimental solvation energies for neutral molecules

and ions are close to the performances of others solvation approaches with a larger number

of parameters. In particular for ions the implemented approach outperforms in accuracy all

existing implicit solvation models. The flexibility of the model to locally change atomic radii

allows to simultaneously simulate neutrals and ions, conditions typical of a double layer of

an electrolyte in contact with a surface. A new benchmark protocol for solid-liquid inter-

faces is proposed in terms of wettability and contact angles. This allows to find parameters

for elements not represented in current solvation databases. Results are shown for several

metallic and insulating surfaces in contact with water. As a showcase application we studied

the terminations of CdS (1120) surface in an electrochemical medium as a function of bias

potential and pH. We observed that the inclusion of solvent effects introduces a substantial

shift in the bias potentials at which surface terminations change. The relative stabilities

of several adsorbed species are affected by the presence of solvent. As a consequence, its

inclusion in computations which are based on the CHE approach is crucial.

Acknowledgement

G. Fisicaro thanks Dr. D. Tomerini for continuous useful discussions. This work was done

within the PASC and NCCR MARVEL projects. Computer resources were provided by

the Swiss National Supercomputing Centre (CSCS) under Project ID s707. L. Genovese

acknowledges also support from the EXTMOS EU project.

Supporting Information

Convergence tests are reported varying the k -point sampling for the total energy as well as

the zero-point energy EZPE and the vibrational entropic contribution TS of the clean CdS

surface. Formulas are presented for the free energy of liquid water GH2O(l)(298 K, 1 bar).

This information is available free of charge via the Internet at http://pubs.acs.org.

36

Page 36 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 38: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

A Set up of KS Self-Consistent cycle under implicit en-

vironments

In this section we will outline the main equations that have to be implemented in a DFT code

for the wavefunction optimization procedure in presence of implicit environments. Formulas

reported are general and do not depend on the particular functional form of the dependence

on the atomic coordinates Ri or on the electronic charge density ρ(r) for the cavity.

A.1 Electrostatic Energy

For an implicit environment the Hartree Potential might be seen as a functional of both

the electronic density and the dielectric function, i.e. φ = φ[ρ, ǫ; Ω]. Here Ω represents

the real-space domain (together with the appropriate boundary conditions) choosen to solve

Eq. (1). Clearly, linearity is preserved with respect to ρ, namely φ[ρ1+ρ2, ǫ; Ω] = φ[ρ1, ǫ; Ω]+

φ[ρ2, ǫ; Ω]. A given electrostatic potential is related to an electrostatic energy

Eel[ρ, ǫ; Ω] ≡1

2

Ω

drρ(r)φ[ρ, ǫ; Ω](r) , (28)

that by construction is quadratic in the charge density ρ, which implies

Eel[ρ1 + ρ2, ǫ; Ω] = Eel[ρ1, ǫ; Ω] + Eel[ρ2, ǫ; Ω] +

Ω

drφ[ρ1, ǫ; Ω](r)ρ2(r) . (29)

It can be easily demonstrated that the functional derivative of the electrostatic energy with

respect to the dielectric function is:

δEel[ρ, ǫ; Ω]

δǫ(r)= − 1

8π|∇φ[ρ, ǫ; Ω](r)|2 . (30)

As the electrostatic environment is modified, also the terms describing the ionic energy

and potential must be defined accordingly. In the definition of the external potential con-

37

Page 37 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 39: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

tribution, we have to treat consistently the contribution of v(r) coming from the solution of

the generalized Poisson equation where ionic charges ρaion can be given as input. In other

terms, we can define

v(r) = vion(r) + v∆(r) (31)

where vion = φ[ρion, ǫ; Ω], and ρion =∑

a ρaion. Here v∆ contains extra terms, localised around

the atoms, coming for instance from the definition of the pseudopotentials. In this framework,

the ionic energy of the system can equivalently be defined as the electrostatic energy of the

above generalized Poisson equation minus the self-energy contribution of each of the ions in

vacuum.

Eion[ǫ] = Eel[ρion, ǫ; Ω]− Eself , (32)

where

Eself =∑

a

Eel[ρaion, 1;R3] (33)

A.2 Calculating the Ground-State energy in the Kohn-Sham for-

mulation

The total energy within the Kohn-Sham formalism and the soft-sphere solvation model might

be defined as:

E[ρ, ǫ] = Eion[ǫ] + Ts[ρ] +

v(r)ρ(r)dr + Eel[ρ, ǫ; Ω] + Exc[ρ] + (α + γ)S[ǫ] + βV [ǫ]

= Ts[ρ] + E∆[ρ] + Exc[ρ] + Eel[ρ+ ρion, ǫ; Ω]− Eself + (α + γ)S[ǫ] + βV [ǫ] . (34)

where Ts[ρ] is the electron kinetic energy, Eel[ρ+ ρion, ǫ; Ω] the total electrostatic energy re-

lated to the ionic charges ρion and electronic charge density ρ, Exc[ρ] the exchange-correlation

term. The additional terms E∆[ρ] and Eself are both related to the ionic charges. In par-

ticular E∆ =∫

drv∆(r)ρ(r). The non-electrostatic terms to the total energy come from Eq.

(5).

38

Page 38 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 40: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Within this formula, the Kohn-Sham hamiltonian can be defined asHKS = −1/2∇2+VKS,

where

VKS[ρ](r) = Vxc(r) + v∆(r) + φtot(r) + Vextra(r) , (35)

Vextra(r) = − 1

dr′δǫ(r′)

δρ(r)

|∇φtot(r′)|2 + (α + γ)

δS[ǫ]

δǫ(r′)

− β

ǫ0 − 1

δǫ(r)

δρ(r). (36)

Here we called the total solute charge ρtot = ρ+ ρion and the corresponding potential φtot =

φ[ρtot, ǫ; Ω]. In Eqs. (35) and (36) an explicit dependence of the dielectric function from

the charge density has been considered, namely ǫ = ǫ[ρ], and δS[ǫ]/δǫ(r′) is explicited in

Eq. (13). When ǫ is a local function of the density,18,22 we have δǫ(r′)/δρ(r) = ǫ(r)δ(r− r′)

and ∂iǫ = ǫ∂iρ. Therefore

δS

δρ=

ǫ

ǫ0 − 1

ǫ

|ǫ|

(

∂iρ∂jρ∂i∂jρ

|∇ρ|3 − ∇2ρ

|∇ρ|

)

(37)

= − |ǫ|ǫ0 − 1

∂i∂iρ

|∇ρ| (38)

Vextra(r) = −ǫ(r)(

1

8π|∇φtot(r)|2 +

α + γ

ǫ0 − 1(39)

[

∂iρ(r)∂jρ(r)∂i∂jρ(r)

|∇ρ(r)|3 − ∇2ρ(r)

|∇ρ(r)|

]

ǫ0 − 1

)

As the dielectric function ǫ might evolve with the charge density, it is important to re-

evaluate the ionic contribution to the total energy for each of the density optimisation steps.

In the above equation we have also assumed that ǫ/|ǫ| = −1 as the dielectric function should

increase with the decreasing of the density.

The DFT energy can be alternatively obtained from the band structure term of the KS

Hamiltonian:

EBS =∑

j

fj〈ψj|HKS|ψj〉 = Ts[ρ] + Epot (40)

39

Page 39 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 41: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

where the potential energy is defined as the expectation value of the local potential:

Epot =

drVKS(r)ρ(r) = Etotel − Eion − Eself + Eel + E∆ + EVxc + EVex . (41)

Here we called Etotel = Eel[ρtot, ǫ; Ω], and

EVex =

drVextra(r)ρ(r) , (42)

EVxc =

drVxc(r)ρ(r) , (43)

By defining the Hartree energy of the cavity as

EcH ≡ Epot − Etot

el − E∆ − EVxc , (44)

we can recover the formal expression for the KS energy:

E = EBS − EcH − EVxc − Eself . (45)

The above expression might be interesting as it does not require an explicit evaluation of

the double-counting term EVex.

A.3 Forces for the soft-sphere model

Forces are defined as minus the derivative of the total energy E[ρ, ǫ] [Eq. (34)] with respect

to the atomic positions:

Fa = − dE

dRa

. (46)

Here the subscript a points to a particular atom.

The Hellmann-Feynman theorem states that, at self-consistency, the total derivative be-

comes the derivative with respect to the explicit dependence on the atomic positions. Since

the electronic kinetic energy Ts[ρ] and the exchange-correlation contribution Exc[ρ] do not

40

Page 40 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 42: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

depend explicitly on the atomic coordinates, the only terms contributing to the forces are

related to the total electrostatic energy and to the non-electrostatic contributions of the

surrounding dielectric related to the quantum surface and volume. Contributions also come

from the additional terms E∆[ρ] and Eself.

Being φtot = φ[ρtot, ǫ; Ω] the electrostatic potential generated by ρ+ρion, the contribution

coming from the total electrostatic energy Eel[ρ+ ρion, ǫ; Ω] can be split in different parts:

∂Eel

∂Ra

=

drφtot(r)∂ρion(r)

∂Ra

− 1

dr∂ǫ(r)

∂Ra

|∇φtot(r)|2 , (47)

which gives rise to the following expression

Fa = −∂E∆

∂Ra

−∫

drφtot(r)∂ρion(r)

∂Ra

−∫

dr∂ǫ(r)

∂Ra

(α + γ)δS[ǫ]

δǫ(r)+ β

δV [ǫ]

δǫ(r)− 1

8π|∇φtot(r)|2

. (48)

By symmetry ∂Eself/∂Ra = 0. The first two terms can be calculated as in vacuum. Consid-

ering the particular case of our soft-sphere cavity defined by Eq. (6), all non-electrostatic

additional terms can be analytically evaluated. To compute the nabla of φtot, central, for-

ward and backward finite difference filters of order 16 have been used, which match the

accuracy of the underlying vacuum Poisson solver.27

B Input guess implementation for the PCG algorithm

In our previous work9 we reported a preconditioned conjugate gradient (PCG) algorithm to

solve the generalized Poisson equation, i.e. Eq. (1), to compute the electrostatic potential

φ(r) generated by a given charge density ρ(r). We demonstrated that it is able to solve the

electrostatic problem with some ten iterations reaching an accuracy of ∼ 10−10 arb. units

with respect to analytic functions.

During a self-consistent field (SCF) loop, the generalized Poisson equation has to be

41

Page 41 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 43: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

solved several times as much as the number of iterations needed to reach the wavefunction

convergence. Solving the problem repeatedly from scratch is computationally inefficient,

especially when the wavefunction is approaching its final configuration. A proper use of

an input guess for the potential φ(r) from previous calculations can considerably speed up

the solution of the electrostatic problem and, by consequence, the whole SCF loop making

implicit solvation and gas phase runtime closer. In our previous paper we implemented a

standard input guess approach for the electrostatic potential, where the generalized Poisson

operator

A = − 1

4π∇ · ǫ(r)∇ (49)

was applied to the potential input guess φ0 and the minimization procedure started from the

initial gradient r0 = ρ − Aφ0. Here ǫ(r) is the continuum dielectric function describing the

dielectric cavity. This approach makes use of a finite difference filter for the nabla operator

∇, which could affect the overall converge especially when the mesh of the simulation box

is too coarse or numbers become small with respect to the numerical noise. Such an effect

could definitely corrupt the overall PCG performances.

In order to prevent that, a solution has been found and reported in Algorithm 1. The

implemented idea follows the same mathematical argument which allowed us to avoid the use

of finite difference filters within the PCG scheme (see Ref. [ 9] for details). The minimization

procedure starts from the initial gradient r0. P is the preconditioner which inverse has to

be applied to the residual vector rk returning the preconditioned residual vk, and, finally, φk

is the solution of the generalized Poisson equation. The convergence criterion is imposed on

the Euclidean norm of the residual vector rk.

Both performances and accuracy of the suggested PCG scheme (with input guess) has

been tested for a single-point DFT energy evaluation of a water molecule in implicit water

described by the soft-sphere solvation model. Fig. 5 report the gradient norm during the

wave function optimization. The SCF loop converged in 15 iterations. Red vertical dots

represent the number of PCG iterations needed to solve the generalized Poisson equation on

42

Page 42 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 44: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Algorithm 1 Preconditioned conjugate gradient (PCG) algorithm with input guess

1: compute q = 14π

√ǫ ∇2√ǫ

2: r0 = ρ− φ0q

3: v0 = P−1r0

4: r1 = (φ0 − v0)q

5: set p0 = s0 = 0, φ1 = v0

6: for k = 1, ... do

7: if ‖rk‖2 ≤ τ exit

8: vk = P−1rk

9: pk = vk + βkpk−1 (where βk = (vk,rk)(vk−1,rk−1)

, k 6= 0 )

10: sk = Apk = vkq + rk + βksk−1 (since Avk = vkq + rk)

11: αk = (vk,rk)(pk,sk)

12: φk+1 = φk + αkpk

13: rk+1 = rk − αksk

14: end for

43

Page 43 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 45: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

each SCF step. Beside the input guess, a dynamic exit from the PCG loop has been also used

as accurate potentials are not needed during the first stages of SCF optimization. Hence the

exit threshold τ from the PCG loop on the residual norm ‖rk‖2 is made dynamic, set equal

to the norm of the Kohn-Sham energy gradient from the current wave function optimization

and allowed to vary between τmin and τmax (setting τmin = 10−4 and τmin = 10−6 guarantees

a correct SCF convergence while lowering the overall electrostatic calculation time for the

water molecule in implicit water). Fig. 5 shows that a mean of 4 PCG iterations are needed

for each SCF iteration, which ultimately consists of a standard Poisson application and a

mere vector multiplication. Tests on a larger system have been reported in our previous

work9 for a protein PDB ID: 1y49 (122 atoms) and compared to an optimized version of

the COSMO51 Poisson Solver for dielectric environments.101 Applying the soft-sphere cavity

with the new input guess implementation (Algorithm 1), we recovered a ratio of the wave

function optimization runtime in solvent and gas phase of 1.16.

Figure 5: SCF convergence during a single-point DFT energy evaluation for water molecules inimplicit water. Red vertical dots represent the number of PCG iterations to solve the generalizedPoisson equation. Input guess and dynamic exit from the PCG loop have been used.

44

Page 44 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 46: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

References

(1) Dabo, I.; Li, Y.; Bonnet, N.; Marzari, N. In Fuel Cell Science: Theory, Fundamentals

and Bio-Catalysis; Wieckowski, A., Nørskov, J., Eds.; John Wiley & Sons, Inc., 2010;

pp 415–431.

(2) White, J. A.; Schwegler, E.; Galli, G.; Gygi, F. The solvation of Na+ in water: First-

principles simulations. J. Chem. Phys. 2000, 113, 4668–4673.

(3) Goedecker, S. Minima hopping: An efficient search method for the global minimum

of the potential energy surface of complex molecular systems. J. Chem. Phys. 2004,

120, 9911–9917.

(4) Schaefer, B.; Mohr, S.; Amsler, M.; Goedecker, S. Minima hopping guided path search:

An efficient method for finding complex chemical reaction pathways. J. Chem. Phys.

2014, 140, –.

(5) 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.

(6) Grossman, J. C.; Schwegler, E.; Draeger, E. W.; Gygi, F.; Galli, G. Towards an

assessment of the accuracy of density functional theory for first principles simulations

of water. J. Chem. Phys. 2004, 120, 300–311.

(7) VandeVondele, J.; Mohamed, F.; Krack, M.; Hutter, J.; Sprik, M.; Parrinello, M. The

influence of temperature and density functional models in ab initio molecular dynamics

simulation of liquid water. J. Chem. Phys. 2005, 122, 014515.

(8) Sit, P. H.-L.; Marzari, N. Static and dynamical properties of heavy water at ambi-

ent conditions from first-principles molecular dynamics. J. Chem. Phys. 2005, 122,

204510.

45

Page 45 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 47: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(9) Fisicaro, G.; Genovese, L.; Andreussi, O.; Marzari, N.; Goedecker, S. A generalized

Poisson and Poisson-Boltzmann solver for electrostatic environments. J. Chem. Phys.

2016, 144 .

(10) Onsager, L. Electric Moments of Molecules in Liquids. J. Am. Chem. Soc. 1936, 58,

1486–1493.

(11) 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.

(12) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation

Models. Chem. Rev. 2005, 105, 2999–3094.

(13) Cramer, C. J.; Truhlar, D. G. Implicit Solvation Models: Equilibria, Structure, Spec-

tra, and Dynamics. Chem. Rev. 1999, 99, 2161–2200.

(14) Senn, H. M.; Margl, P. M.; Schmid, R.; Ziegler, T.; Blöchl, P. E. Ab initio molecular

dynamics with a continuum solvation model. J. Chem. Phys. 2003, 118 .

(15) Lange, A. W.; Herbert, J. M. A smooth, nonsingular, and faithful discretization scheme

for polarizable continuum models: The switching/Gaussian approach. J. Chem. Phys.

2010, 133 .

(16) Foresman, J. B.; Keith, T. A.; Wiberg, K. B.; Snoonian, J.; Frisch, M. J. Solvent

Effects. 5. Influence of Cavity Shape, Truncation of Electrostatics, and Electron Cor-

relation on ab Initio Reaction Field Calculations. J. Phys. Chem. 1996, 100, 16098–

16104.

(17) Wiberg, K. B.; Keith, T. A.; Frisch, M. J.; Murcko, M. Solvent Effects on 1,2-

Dihaloethane Gauche/Trans Ratios. J. Phys. Chem. 1995, 99, 9072–9079.

(18) Fattebert, J.-L.; Gygi, F. Density functional theory for efficient ab initio molecular

dynamics simulations in solution. J. Comput. Chem. 2002, 23, 662–666.

46

Page 46 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 48: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(19) Scherlis, D. A.; Fattebert, J.-L.; Gygi, F.; Cococcioni, M.; Marzari, N. A unified

electrostatic and cavitation model for first-principles molecular dynamics in solution.

J. Chem. Phys. 2006, 124, –.

(20) Dziedzic, J.; Helal, H. H.; Skylaris, C.-K.; Mostofi, A. A.; Payne, M. C. Minimal

parameter implicit solvent model for ab initio electronic-structure calculations. EPL

(Europhys. Lett.) 2011, 95, 43001.

(21) Cococcioni, M.; Mauri, F.; Ceder, G.; Marzari, N. Electronic-Enthalpy Functional for

Finite Systems Under Pressure. Phys. Rev. Lett. 2005, 94, 145501.

(22) Andreussi, O.; Dabo, I.; Marzari, N. Revised self-consistent continuum solvation in

electronic-structure calculations. J. Chem. Phys. 2012, 136, –.

(23) Dupont, C.; Andreussi, O.; Marzari, N. Self-consistent continuum solvation (SCCS):

The case of charged systems. J. Chem. Phys. 2013, 139 .

(24) Sánchez, V. M.; Sued, M.; Scherlis, D. A. First-principles molecular dynamics simula-

tions at solid-liquid interfaces with a continuum solvent. J. Chem. Phys. 2009, 131,

174108.

(25) Genovese, L.; Neelov, A.; Goedecker, S.; Deutsch, T.; Ghasemi, S. A.; Willand, A.; Cal-

iste, D.; Zilberberg, O.; Rayson, M.; Bergman, A.; Schneider, R. Daubechies wavelets

as a basis set for density functional pseudopotential calculations. J. Chem. Phys. 2008,

129 .

(26) http://www.bigdft.org

(27) Genovese, L.; Deutsch, T.; Neelov, A.; Goedecker, S.; Beylkin, G. Efficient solution of

Poisson’s equation with free boundary conditions. J. Chem. Phys. 2006, 125, –.

(28) Marenich, A. V.; Kelly, C. P.; Thompson, J. D.; Hawkins, G. D.; Chambers, C. C.;

47

Page 47 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 49: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Giesen, D. J.; Winget, P.; Cramer, C. J.; Truhlar, D. G. Minnesota Solvation Database

- version 2012 ; University of Minnesota, Minneapolis, 2012; p 3.

(29) O’Regan, B.; Gratzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized

colloidal TiO2 films. Nature 1991, 353, 737–740.

(30) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites

as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131,

6050–6051.

(31) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Efficient

Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites.

Science 2012, 338, 643–647.

(32) Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem.

Soc. Rev. 2009, 38, 253–278.

(33) Nocera, D. G. The Artificial Leaf. Acc. Chem. Res. 2012, 45, 767–776.

(34) Amovilli, C.; Mennucci, B. Self-Consistent-Field Calculation of Pauli Repulsion and

Dispersion Contributions to the Solvation Free Energy in the Polarizable Continuum

Model. J. Phys. Chem. B 1997, 101, 1051–1057.

(35) Cramer, C. J.; Truhlar, D. G. A Universal Approach to Solvation Modeling. Acc.

Chem. Res. 2008, 41, 760–768.

(36) Barone, V.; Cossi, M.; Tomasi, J. Geometry optimization of molecular structures in

solution by the polarizable continuum model. J. Comput. Chem. 1998, 19, 404–417.

(37) Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. SM6:âĂĽ A Density Functional Theory

Continuum Solvation Model for Calculating Aqueous Solvation Free Energies of Neu-

trals, Ions, and Solute-Water Clusters. J. Chem. Theory Comput. 2005, 1, 1133–1152.

48

Page 48 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 50: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(38) Marenich, A. V.; Olson, R. M.; Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Self-

Consistent Reaction Field Model for Aqueous and Nonaqueous Solutions Based on

Accurate Polarized Partial Charges. J. Chem. Theory Comput. 2007, 3, 2011–2033.

(39) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Generalized Born Solvation Model

SM12. J. Chem. Theory Comput. 2013, 9, 609–620.

(40) Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials.

Phys. Rev. B 1996, 54, 1703–1710.

(41) Willand, A.; Kvashnin, Y. O.; Genovese, L.; Vázquez-Mayagoitia, A.; Deb, A. K.;

Sadeghi, A.; Deutsch, T.; Goedecker, S. Norm-conserving pseudopotentials with chem-

ical accuracy compared to all-electron calculations. J. Chem. Phys. 2013, 138, –.

(42) Marques, M. A.; Oliveira, M. J.; Burnus, T. Libxc: A library of exchange and correla-

tion functionals for density functional theory. Comput. Phys. Comm. 2012, 183, 2272

– 2281.

(43) Genovese, L.; Deutsch, T.; Goedecker, S. Efficient and accurate three-dimensional

Poisson solver for surface problems. J. Chem. Phys. 2007, 127, –.

(44) Shivakumar, D.; Williams, J.; Wu, Y.; Damm, W.; Shelley, J.; Sherman, W. Prediction

of Absolute Solvation Free Energies using Molecular Dynamics Free Energy Perturba-

tion and the OPLS Force Field. J. Chem. Theory Comput. 2010, 6, 1509–1519.

(45) Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a con-

tinuum. A direct utilization of AB initio molecular potentials for the prevision of

solvent effects. Chem. Phys. 1981, 55, 117 – 129.

(46) Bondi, A. van der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451.

(47) Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press:Ithaca,

NY, 1960; p 257.

49

Page 49 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 51: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(48) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard III, W. A.; Skiff, W. M.

UFF, a full periodic table force field for molecular mechanics and molecular dynamics

simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035.

(49) Cancès, E.; Mennucci, B.; Tomasi, J. A new integral equation formalism for the po-

larizable continuum model: Theoretical background and applications to isotropic and

anisotropic dielectrics. J. Chem. Phys. 1997, 107, 3032–3041.

(50) Mennucci, B.; Cancès, E.; Tomasi, J. Evaluation of Solvent Effects in Isotropic

and Anisotropic Dielectrics and in Ionic Solutions with a Unified Integral Equation

Method:âĂĽ Theoretical Bases, Computational Implementation, and Numerical Ap-

plications. J. Phys. Chem. B 1997, 101, 10506–10517.

(51) Klamt, A.; Schuurmann, 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. 2 1993, 799–805.

(52) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.;

Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S.; Windus, T. L.;

Dupuis, M.; Montgomery, J. A. General atomic and molecular electronic structure

system. J. Comput. Chem. 1993, 14, 1347–1363.

(53) Tannor, D. J.; Marten, B.; Murphy, R.; Friesner, R. A.; Sitkoff, D.; Nicholls, A.;

Honig, B.; Ringnalda, M.; Goddard, W. A. Accurate First Principles Calculation

of Molecular Charge Distributions and Solvation Energies from Ab Initio Quantum

Mechanics and Continuum Dielectric Theory. J. Am. Chem. Soc. 1994, 116, 11875–

11882.

(54) Marten, B.; Kim, K.; Cortis, C.; Friesner, R. A.; Murphy, R. B.; Ringnalda, M. N.;

Sitkoff, D.; Honig, B. New Model for Calculation of Solvation Free Energies: Correc-

50

Page 50 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 52: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

tion of Self-Consistent Reaction Field Continuum Dielectric Theory for Short-Range

Hydrogen-Bonding Effects. J. Phys. Chem. 1996, 100, 11775–11788.

(55) Valiev, M.; Bylaska, E.; Govind, N.; Kowalski, K.; Straatsma, T.; Dam, H. V.;

Wang, D.; Nieplocha, J.; Apra, E.; Windus, T.; de Jong, W. NWChem: A com-

prehensive and scalable open-source solution for large scale molecular simulations.

Comput. Phys. Commun. 2010, 181, 1477 – 1489.

(56) Orozco, M.; Luque, F. Optimization of the cavity size for ab initio MST-SCRF calcu-

lations of monovalent ions. Chem. Phys. 1994, 182, 237 – 248.

(57) Miertuš, S.; Bartoš, J.; Trebatická, M. Dependance of atomic radii and volumes on

the electron distribution in solute molecule and on solute-solvent interaction. J. Mol.

Liq. 1987, 33, 139 – 156.

(58) Haynes, W. M. CRC Handbook of Chemistry and Physics; CPC Press, Boca Raton,

2012; Chapter 6, p 187.

(59) Earhart, H. W.; Komin, A. P. Kirk-Othmer Encyclopedia of Chemical Technology ;

John Wiley & Sons, Inc., 2000; p 1.

(60) Sundararaman, R.; Schwarz, K. A.; Letchworth-Weaver, K.; Arias, T. A. Spicing up

continuum solvation models with SaLSA: The spherically averaged liquid susceptibility

ansatz. J. Chem. Phys. 2015, 142, 054102.

(61) Drelich, J.; Chibowski, E.; Meng, D. D.; Terpilowski, K. Hydrophilic and superhy-

drophilic surfaces and materials. Soft Matter 2011, 7, 9804–9828.

(62) Sedlmeier, F.; Janecek, J.; Sendner, C.; Bocquet, L.; Netz, R. R.; Horinek, D. Water

at polar and nonpolar solid walls (Review). Biointerphases 2008, 3 .

(63) Young, T. An Essay on the Cohesion of Fluids. Phil. Trans. R. Soc. Lond. 1805, 95,

65–87.

51

Page 51 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 53: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(64) Tadmor, R. Line Energy and the Relation between Advancing, Receding, and Young

Contact Angles. Langmuir 2004, 20, 7659–7664.

(65) Dupré, A. M.; Dupré, P. Théorie mécanique de la chaleur ; Gauthier-Villars Paris,

1869; p 369.

(66) Girifalco, L.; Good, R. A theory for the estimation of surface and interfacial energies.

I. Derivation and application to interfacial tension. J. Phys. Chem. 1957, 61, 904–909.

(67) Good, R.; Girifalco, L. A theory for estimation of surface and interfacial energies. III.

Estimation of surface energies of solids from contact angle data. J. Phys. Chem. 1960,

64, 561–565, cited By 356.

(68) Fowkes, F. Additivity of intermolecular forces at interfaces. I. Determination of the

contribution to surface and interfacial tensions of dispersion forces in various liquids.

J. Phys. Chem. 1963, 67, 2538–2541.

(69) Li, D.; Neumann, A. Equation of state for interfacial tensions of solid-liquid systems.

Adv. Colloid Interface Sci. 1992, 39, 299 – 345.

(70) van Oss, C. J. Acid-base interfacial interactions in aqueous media. Colloids Surf. A:

Physicochem. Eng. Aspects 1993, 78, 1–49.

(71) C. J. van Oss, Interfacial Forces in Aqueous Media, Second Edition; CRC Press, 2006;

pp 131–155.

(72) Goumans, T. P. M.; Wander, A.; Brown, W. A.; Catlow, C. R. A. Structure and

stability of the (001) α-quartz surface. Phys. Chem. Chem. Phys. 2007, 9, 2146–2152.

(73) Schrader, M. E. Wettability of clean metal surfaces. J. Colloid Interface Sci. 1984,

100, 372 – 380.

52

Page 52 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 54: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(74) Kim, E.-K.; Kim, J.-Y.; Kim, S. S. Significant change in water contact angle of

electrospray-synthesized SiO2 films depending on their surface morphology. Surf. In-

terface Anal. 2013, 45, 656–660.

(75) Sendner, C.; Horinek, D.; Bocquet, L.; Netz, R. R. Interfacial Water at Hydrophobic

and Hydrophilic Surfaces: Slip, Viscosity, and Diffusion. Langmuir 2009, 25, 10768–

10781.

(76) Wang, S.; Zhang, Y.; Abidi, N.; Cabrales, L. Wettability and Surface Free Energy of

Graphene Films. Langmuir 2009, 25, 11078–11081.

(77) Fowkes, F. M.; Harkins, W. D. The State of Monolayers Adsorbed at the Interface

Solid-Aqueous Solution. J. Am. Chem. Soc. 1940, 62, 3377–3386.

(78) Werder, T.; Walther, J. H.; Jaffe, R. L.; Halicioglu, T.; Koumoutsakos, P. On the

Water-Carbon Interaction for Use in Molecular Dynamics Simulations of Graphite

and Carbon Nanotubes. J. Phys. Chem. B 2003, 107, 1345–1352.

(79) Taherian, F.; Marcon, V.; van der Vegt, N. F. A.; Leroy, F. What Is the Contact Angle

of Water on Graphene? Langmuir 2013, 29, 1457–1465.

(80) Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem.

Soc. Rev. 2009, 38, 253–278.

(81) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. A.;

Lewis, N. S. Solar Water Splitting Cells. Chem. Rev. 2010, 110, 6446–6473.

(82) Ashokkumar, M. An overview on semiconductor particulate systems for photoproduc-

tion of hydrogen. Int. J. Hydrogen Energy 1998, 23, 427 – 438.

(83) Bao, N.; Shen, L.; Takata, T.; Domen, K. Self-Templated Synthesis of Nanoporous

CdS Nanostructures for Highly Efficient Photocatalytic Hydrogen Production under

Visible Light. Chem. Mater. 2008, 20, 110–117.

53

Page 53 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 55: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(84) Simon, T.; Bouchonville, N.; Berr, M. J.; Vaneski, A.; Adrović, A.; Volbers, D.; Wyr-

wich, R.; Döblinger, M.; Susha, A. S.; Rogach, A. L.; Jäckel, F.; Stolarczyk, J. K.;

Feldmann, J. Redox shuttle mechanism enhances photocatalytic H2 generation on

Ni-decorated CdS nanorods. Nat. Mater. 2014, 13, 1013–1018.

(85) Kalisman, P.; Nakibli, Y.; Amirav, L. Perfect Photon-to-Hydrogen Conversion Effi-

ciency. Nano Lett. 2016, 16, 1776–1781.

(86) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J. R.; Bli-

gaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell

Cathode. J. Phys. Chem. B 2004, 108, 17886–17892.

(87) Valdés, Á.; Qu, Z.-W.; Kroes, G.-J.; Rossmeisl, J.; Nørskov, J. K. Oxidation and

Photo-Oxidation of Water on TiO2 Surface. J. Phys. Chem. C 2008, 112, 9872–9879.

(88) Bajdich, M.; García-Mota, M.; Vojvodic, A.; Nørskov, J. K.; Bell, A. T. Theoretical

Investigation of the Activity of Cobalt Oxides for the Electrochemical Oxidation of

Water. J. Am. Chem. Soc. 2013, 135, 13521–13530.

(89) Neufeld, O.; Toroker, M. C. Platinum-Doped α-Fe2O3 for Enhanced Water Splitting

Efficiency: A DFT+U Study. J. Phys. Chem. C 2015, 119, 5836–5847.

(90) Zhang, X.; Klaver, P.; van Santen, R.; van de Sanden, M. C. M.; Bieberle-Hütter, A.

Oxygen Evolution at Hematite Surfaces: The Impact of Structure and Oxygen Va-

cancies on Lowering the Overpotential. J. Phys. Chem. C 2016, 120, 18201–18208.

(91) Nguyen, M.-T.; Piccinin, S.; Seriani, N.; Gebauer, R. Photo-Oxidation of Water on

Defective Hematite(0001). ACS Catal. 2015, 5, 715–721.

(92) Hellman, A.; Iandolo, B.; Wickman, B.; Grönbeck, H.; Baltrusaitis, J. Electro-

oxidation of water on hematite: Effects of surface termination and oxygen vacancies

investigated by first-principles. Surf. Sci. 2015, 640, 45 – 49.

54

Page 54 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 56: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

(93) Sakong, S.; Naderian, M.; Mathew, K.; Hennig, R. G.; Groß, A. Density functional

theory study of the electrochemical interface between a Pt electrode and an aqueous

electrolyte using an implicit solvent method. J. Chem. Phys. 2015, 142, –.

(94) Sakong, S.; Groß, A. The Importance of the Electrochemical Environment in the

Electro-Oxidation of Methanol on Pt(111). ACS Catal. 2016, 6, 5575–5586.

(95) Lespes, N.; Filhol, J.-S. Using Implicit Solvent in Ab Initio Electrochemical Modeling:

Investigating Li+/Li Electrochemistry at a Li/Solvent Interface. J. Chem. Theory

Comput. 2015, 11, 3375–3382.

(96) Letchworth-Weaver, K.; Arias, T. A. Joint density functional theory of the electrode-

electrolyte interface: Application to fixed electrode potentials, interfacial capacitances,

and potentials of zero charge. Phys. Rev. B 2012, 86, 075140.

(97) Zhou, Z.; Han, F.; Guo, L.; Prezhdo, O. V. Understanding divergent behaviors in

the photocatalytic hydrogen evolution reaction on CdS and ZnS: a DFT based study.

Phys. Chem. Chem. Phys. 2016, 18, 16862–16869.

(98) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made

Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

(99) NIST Computational Chemistry Comparison and Benchmark Database; NIST Stan-

dard Reference Database Number 101, Release 18, October 2016, Editor: Russell D.

Johnson III.

(100) Klaumünzer, M.; Mačković, M.; Ferstl, P.; Voigt, M.; Spiecker, E.; Meyer, B.; Peuk-

ert, W. Phase Transition Behavior and Oriented Aggregation During Precipitation of

In(OH)3 and InOOH Nanocrystals. J. Phys. Chem. C 2012, 116, 24529–24537.

(101) Liu, F.; Luehr, N.; Kulik, H. J.; MartÃŋnez, T. J. Quantum Chemistry for Solvated

55

Page 55 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 57: Soft-sphere continuum solvation in electronic-structure ... · The self-consistent continuum solvation (sccs) model developed by Andreussi et al.22 goes in this direction, satisfying

Page 56 of 56

ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960