Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry...

14
Computational Chemistry Methods for Nanoporous Materials Jack D. Evans, Guillaume Fraux, Romain Gaillac, ,Daniela Kohen, ,§ Fabien Trousselet, Jean-Mathieu Vanson, ,,and Franc ̧ ois-Xavier Coudert* ,Chimie ParisTech, PSL Research University, CNRS, Institut de Recherche de Chimie Paris, 75005 Paris, France Air Liquide, Centre de Recherche Paris Saclay, 78354 Jouy-en-Josas, France § Department of Chemistry, Carleton College, Northeld, Minnesota 55057, United States E ́ cole Normale Supe ́ rieure, PSL Research University, De ́ partement de Chimie, Sorbonne Universite ́ s UPMC Univ Paris 06, CNRS UMR 8640 PASTEUR, 24 rue Lhomond, 75005 Paris, France Laboratoire de Synthe ̀ se et Fonctionnalisation des Cé ramiques, UMR 3080 Saint Gobain CREE/CNRS, 550 Avenue Alphonse Jauret, 84306 Cavaillon, France * S Supporting Information ABSTRACT: We present here the computational chemistry methods our group uses to investigate the physical and chemical properties of nanoporous materials and adsorbed uids. We highlight the multiple time and length scales at which these properties can be examined and discuss the computational tools relevant to each scale. Furthermore, we include the key points to considerupsides, downsides, and possible pitfallsfor these methods. INTRODUCTION The Coudert research group uses molecular simulation, at various scales, to study a wide range of porous materials and molecular uids near interfaces or under connement. Outlined herein is a review of the computational chemistry toolbox we routinely use. Nonetheless we will rst say here a few words about the systems of interest to our research: nanoporous materials and the adsorption of molecular uids within their pore space. Nanoporous materials exhibit high internal surface area, resulting in a wide range of applications, examples of which include large-scale processes in key sectors of the chemical industry, such as gas separation and capture, 1 storage, 2 liquid separation, heterogeneous catalysis, 3 drug delivery, 4 and more. 5 Specically, nanoporous materials display a very rich range of chemical composition (inorganic, organic, metalorganic) and geometrical properties (macroporous, mesoporous, micro- porous). This oers a diverse scope of physical and chemical properties. Among nanoporous materials, our group focuses on two main families of materials, namely zeolites and metalorganic frameworks (MOFs). 6,7 Inorganic zeolites represent the current generation of porous materials widely used in industrial processes, due to their large-scale availability, reasonable cost, and high thermal, mechanical, and chemical stability. The very topical metalorganic frameworks (MOFs) are a more recent class of materials. MOFs are widely considered contenders for specic applications, due to their extraordinarily high porosity and tunability of both their framework dimensions and internal surface chemistry. Of particular interest to our group are framework materials that exhibit counterintuitive or anom- alousphysical properties, such as stimuli-responsive materials (sometimes also called multifunctional or smartmaterials). 8 These frameworks exhibit large-scale changes in their structure and physicochemical properties in response to small stimuli, such as temperature change, mechanical constraint, guest adsorption, and exposure to light or magnetic elds. Our research seeks to capture the structure, dynamics, and thermodynamics of nanoporous systems using an atomistic level description. Because the systems vary widely in length scales, and the processes we want to observe can signicantly dier in their time scales, we use dierent levels of theoryor theories with dierent assumptions. This may, for example, include explicitly treating the electrons of each atom (quantum chemistryapproaches) 9 or treating atoms or groups of atoms as a single unit (classicalapproaches). 10 In all cases, however, the goal of computational chemistry is the same: to model the macroscopic behavior of complex heterogeneous systems, constructed from individual entities and their interactions, under certain external (experimental) conditions. This is achieved by relying on statistical thermodynamics as our fundamental tool for bridging the interactions at the atomistic scale and the macroscopic properties of a condensed matter system. Though numerical methods of statistical thermody- namics can be applied to nonequilibrium processes, 11 we focus Special Issue: Methods and Protocols in Materials Chemistry Received: July 21, 2016 Revised: September 3, 2016 Published: September 6, 2016 Review pubs.acs.org/cm © 2016 American Chemical Society 199 DOI: 10.1021/acs.chemmater.6b02994 Chem. Mater. 2017, 29, 199212 This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Transcript of Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry...

Page 1: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

Computational Chemistry Methods for Nanoporous MaterialsJack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡ Daniela Kohen,†,§ Fabien Trousselet,†

Jean-Mathieu Vanson,†,∥,⊥ and Francois-Xavier Coudert*,†

†Chimie ParisTech, PSL Research University, CNRS, Institut de Recherche de Chimie Paris, 75005 Paris, France‡Air Liquide, Centre de Recherche Paris Saclay, 78354 Jouy-en-Josas, France§Department of Chemistry, Carleton College, Northfield, Minnesota 55057, United States∥Ecole Normale Superieure, PSL Research University, Departement de Chimie, Sorbonne Universites − UPMC Univ Paris 06, CNRSUMR 8640 PASTEUR, 24 rue Lhomond, 75005 Paris, France⊥Laboratoire de Synthese et Fonctionnalisation des Ceramiques, UMR 3080 Saint Gobain CREE/CNRS, 550 Avenue AlphonseJauffret, 84306 Cavaillon, France

*S Supporting Information

ABSTRACT: We present here the computational chemistrymethods our group uses to investigate the physical and chemicalproperties of nanoporous materials and adsorbed fluids. Wehighlight the multiple time and length scales at which theseproperties can be examined and discuss the computational toolsrelevant to each scale. Furthermore, we include the key points toconsiderupsides, downsides, and possible pitfallsfor thesemethods.

■ INTRODUCTION

The Coudert research group uses molecular simulation, atvarious scales, to study a wide range of porous materials andmolecular fluids near interfaces or under confinement. Outlinedherein is a review of the computational chemistry toolbox weroutinely use. Nonetheless we will first say here a few wordsabout the systems of interest to our research: nanoporousmaterials and the adsorption of molecular fluids within theirpore space.Nanoporous materials exhibit high internal surface area,

resulting in a wide range of applications, examples of whichinclude large-scale processes in key sectors of the chemicalindustry, such as gas separation and capture,1 storage,2 liquidseparation, heterogeneous catalysis,3 drug delivery,4 and more.5

Specifically, nanoporous materials display a very rich range ofchemical composition (inorganic, organic, metal−organic) andgeometrical properties (macroporous, mesoporous, micro-porous). This offers a diverse scope of physical and chemicalproperties. Among nanoporous materials, our group focuses ontwo main families of materials, namely zeolites and metal−organic frameworks (MOFs).6,7 Inorganic zeolites represent thecurrent generation of porous materials widely used in industrialprocesses, due to their large-scale availability, reasonable cost,and high thermal, mechanical, and chemical stability. The verytopical metal−organic frameworks (MOFs) are a more recentclass of materials. MOFs are widely considered contenders forspecific applications, due to their extraordinarily high porosityand tunability of both their framework dimensions and internalsurface chemistry. Of particular interest to our group areframework materials that exhibit counterintuitive or “anom-

alous” physical properties, such as stimuli-responsive materials(sometimes also called multifunctional or “smart” materials).8

These frameworks exhibit large-scale changes in their structureand physicochemical properties in response to small stimuli,such as temperature change, mechanical constraint, guestadsorption, and exposure to light or magnetic fields.Our research seeks to capture the structure, dynamics, and

thermodynamics of nanoporous systems using an atomisticlevel description. Because the systems vary widely in lengthscales, and the processes we want to observe can significantlydiffer in their time scales, we use different “levels of theory” ortheories with different assumptions. This may, for example,include explicitly treating the electrons of each atom (“quantumchemistry” approaches)9 or treating atoms or groups of atomsas a single unit (“classical” approaches).10 In all cases, however,the goal of computational chemistry is the same: to model themacroscopic behavior of complex heterogeneous systems,constructed from individual entities and their interactions,under certain external (experimental) conditions. This isachieved by relying on statistical thermodynamics as ourfundamental tool for bridging the interactions at the atomisticscale and the macroscopic properties of a condensed mattersystem. Though numerical methods of statistical thermody-namics can be applied to nonequilibrium processes,11 we focus

Special Issue: Methods and Protocols in Materials Chemistry

Received: July 21, 2016Revised: September 3, 2016Published: September 6, 2016

Review

pubs.acs.org/cm

© 2016 American Chemical Society 199 DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

This is an open access article published under an ACS AuthorChoice License, which permitscopying and redistribution of the article or any adaptations for non-commercial purposes.

Page 2: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

here on computational methods to study systems in thermal,mechanical, or chemical equilibrium.For a system in thermodynamic equilibrium in a given

experimental condition, its state is determined by a number ofthermodynamic variables forming a thermodynamic ensemble.An isochoric−isothermal process, for example, takes place inthe commonly used (N, V, T) canonical ensemble, in which thequantity of matter (N), system volume (V), and absolutetemperature (T) are fixed. In a given ensemble, thethermodynamics of the system and its macroscopic propertiesare then entirely determined by the statistical distribution of allpossible microscopic states (or microstates) of the system. Thisis formalized through the ensemble’s partition function, whosesimplest expression is, for a system with quantized states in thecanonical ensemble and with classical Maxwell−Boltzmannstatistics:

∑=−⎛

⎝⎜⎞⎠⎟Z N V T

Ek T

( , , ) expi

i

B (1)

where i is the microstate index of the system, kB is Boltzmann’sconstant, and Ei is the total energy of a given microstate.In practice, all molecular simulation methods have the same

goal: to evaluate the macroscopic properties of the system byaveraging microscopic properties over a collection of micro-states i (this is the ∑i part of eq 1), weighted by the respectiveprobability of occurrence (Boltzmann probability), whichdepends on their energy (the individual Ei’s involved in thesummation). The differences between the various computa-tional approaches available are due to different methods ofevaluating the energy of states and of sampling the microstatesof the system.Methods of evaluating energy in these systems can use

complex quantum mechanical descriptions including Hartree−Fock, post-Hartree−Fock, and the ubiquitous density func-tional theory (DFT). These highly accurate methods canelucidate structural and mechanical properties resulting fromelectronic processes. Alternatively, energies can be calculatedusing classical potentials, which are less computationallydemanding than quantum mechanical approaches. However,classical simulations require accurate potential parameters(force fields) and cannot straightforwardly model quantumphenomena such as changes in chemical bonding.Additionally, numerical simulation is used in the exploration

of states as illustrated in Figure 1b. A number of differentmethods can be employed to result in an energy or dynamicmotion of the structure. Most simply, a single state can beevaluated to obtain a single-point energy of the system;however, more commonly, we seek to obtain an optimalstructure (local energy minimum) or transition states. Minima,and other states, are realized by iterative single-point energyevaluation corresponding to changes of the structure. More-over, through molecular dynamics, the motion of structures isgenerated by an exploration of states using iterative integrationof Newton’s laws of motion. Finally, states can be explored bystochastic displacementsthis is the basis for Monte Carlosimulations, a useful method for simulating gas adsorption.A number of properties, or time and length scales, are not

available by direct molecular simulation as described above.Here we use several geometric and statistical physics basedapproaches at the mesoscale. The combination of all thesemethods is a rather powerful computational chemistry toolboxfor nanoporous materials, that we highlight in this review.

■ QUANTUM MECHANICAL SIMULATIONIn order to study the structure, energetics, electronic states, andother physicochemical properties of porous crystals with a highlevel of accuracy, one of the tools we routinely use is quantumchemistry calculations based on density functional theory(DFT).9

Fundamentally, in quantum simulations of electronic ground-state properties, the wave function has to be solved (ψ) tosatisfy the Schrodinger equation to give the energy (E) of thesystem. An example of the Schrodinger equation for multipleelectrons interacting with multiple fixed nuclei is given in eqs 2,3, and 4. In these equations m is the electron mass and theHamiltonian (H) has terms relating to the kinetic energy ofeach electron (T), the interaction energy between the atomicnuclei and each electron (V), and finally the interaction energybetween different electrons (i and j) for N electrons (U).

ψ ψ =H E (2)

ψ ψ + + =T V U E[ ] (3)

∑ ∑ ∑ ∑ ψ ψ− ℏ ∇ + + == = = <

⎡⎣⎢⎢

⎤⎦⎥⎥m

V U Er r r2

( ) ( , )i

N

ii

N

ii

N

j ii j

2

1

2

1 1

(4)

The complete wave function is complicated, even for asimple molecule, owing to each atom having many electrons.For instance, the full electronic wave function for CO2 is a 66-dimensional function. Furthermore, the electron−electroninteraction term of the Hamiltonian (∑i = 1

N ∑j<iU(ri,rj)) iscrucial and requires considering the correlations of eachindividual electron with all the other electrons. Therefore, thewave function of the system is a complicated function that

Figure 1. Depiction of time scales captured by the simulationapproaches used in our research (a) and alternative simulationmethods for exploring a chemical energy surface (b).

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

200

Page 3: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

makes an exact or analytical solution impractical. Notably, wavefunctions cannot be directly observed; however, eq 5 illustrateshow the density of electrons (n(r)) is determined from theindividual electron wave functions, ψi.

∑ ψ ψ= *n r r r( ) 2 ( ) ( )i

i i(5)

DFT methods apply two Hohenberg−Kohn theorems toreduce the complexity of the problem immensely.12 The firststates that the ground-state energy (and other properties) isdetermined uniquely by the ground-state electron density. As aconsequence, a solution to the Schrodinger equation requires afunction of three spatial coordinates, the electron density,rather than the original wave function, which has 3N variablesfor N electrons. The second states that the correct electrondensity minimizes the overall energy functional and sub-sequently corresponds to the solution of the Schrodingerequation. This latter theorem affords a direction toward findingthe electron density of the system. Kohn and Sham appliedthese theorems to show the correct electron density can bedetermined by solving a set of single electron equations of theform illustrated in eq 6.13 The solutions of these equations arestraightforward single-electron wave functions that depend onthree spatial coordinates.

ψ ε ψ− ℏ ∇ + + + =⎡⎣⎢

⎤⎦⎥m

V V Vr r r r r2

( ) ( ) ( ) ( ) ( )i i i

22

H XC(6)

The Kohn−Sham equations contain three potentials: V, VH,and VXC. The first, V, describes the interaction between anelectron and the collection of atomic nuclei, similar to V. Thispotential has a simple analytical form. The second, VH,describes the repulsion between the single-electron and thetotal electron density. It is named the Hartree potential and isdefined by eq 7, where the electron density is given by n.Importantly, this potential includes a self-interaction contribu-tion, as the electron described by the Kohn−Sham equationalso contributes to the total electron density. Thus, VH includesa nonphysical Coulombic interaction concerning an electronand itself which is corrected for in VXC. This final potential,VXC, characterizes the electron exchange and correlationinteractions to the single-electron Kohn−Sham equations.The potential is formally defined as a “functional derivative”of the exchange−correlation energy (EXC) which is not known,except for a free electron gas. Nevertheless, approximationsexist, which permit the calculation of certain physical quantitiesaccurately.

∫= ′| − ′|

′V en

rrr

r r( )

( )dH

2 3(7)

To solve the Kohn−Sham equations, the definition ofHartree potential (VH) requires the electron density (n), whichdepends on the single-electron wave functions (ψi), which inturn requires the solution of the Kohn−Sham equations. Assuch, an iterative approach is applied to obtain a solution.14

This is routinely achieved using the self-consistent fieldmethod, an example of a simplified algorithm for this approachis outlined below:

1. Define an initial trial electron density, n(r).2. Calculate corresponding potentials, V, VH, and VXC, for

the electron density.3. Solve the Kohn−Sham equations to find the single-

electron wave functions, ψi.

4. Calculate the electron density, nKS, from the constructedKohn−Sham equations using the single-electron wavefunctions, ψi, from step 3.

5. Compare the calculated electron density, nKS(r), with theelectron density used in solving the Kohn−Shamequations, n(r). If the two densities are the same, thenthis is the ground-state electron density and it can beused to compute the total energy. If the two densities aredifferent, then a new trial electron density, n(r), is usedand the process begins again from step 2.

Reliable approximations for the exchange-correlation func-tional (EXC) are required for the solution of the Kohn−Shamequations. There are a number of approaches for treatingexchange-correlation, including local density approximation(LDA), generalized-gradient approximation (GGA), meta-GGA, and hybrid functionals. Each functional has particularstrengths and weaknesses that must be considered and proven.Ultimately, the power of DFT is simplifying a 3N-

dimensional problem, where N is the number of electrons,into a 3-dimensional problem. Consequently, DFT is a popularchoice for electronic structure calculation, of both moleculesand crystals, because of the accuracy obtained for relativelycheap computational cost. Moreover, it is readily available andwidely implemented in several open source, academic, andcommercial software packages. Recently, by comparing theequation of states for 71 elemental crystals with many DFTcodes and methods, a collection of the solid state communityhas demonstrated the reproducibility of DFT simulations.15

The results for modern codes and methods agree very well,exhibiting pairwise differences which are comparable todifferent experiments conducted with high-precision. In thefollowing, we show how we use DFT to predict theoreticalstructures of new materials, model mechanical responses byelastic constant calculations, and even study the dynamics ofmaterials on short time scales. Moreover, we describe how wetypically perform these calculations, and which technical aspectsare important for numerical efficiency and quality of results.

Modeling Crystals with DFT. The simplest type of DFTcalculation on crystal structures are “static” DFT calculationsused to explore local features of their energy landscape. Thisincludes single-point energy calculations, energy minimization,harmonic vibration modes, and more. Since most of oursystems (MOFs and zeolites) typically feature high symmetry,we employ the CRYSTAL14 software package.16 CRYSTAL14 useslocalized basis sets and is appropriate for insulating systems,such as zeolites and MOFs. This methodology is computation-ally efficient in porous systems, compared to plane waves whichwould describe void and occupied spaces with the same level ofprecision. In practice, atomic basis sets are finite and their size(and parametrization), for each element in the system, ischosen as a compromise between computational cost andaccuracy. Depending on the accuracy needed, we use either theDZVP (double-ζ valence polarized) or TZVP (triple-ζ valencepolarized) basis sets, or alternatively basis sets parametrized forchemically similar compounds. Importantly, there is a detailedlibrary of basis sets cataloged by the software developers,available online at http://www.crystal.unito.it/basis-sets.php.To determine which type of basis set to use, we usually runsingle-point energy calculations with the various candidate basissets, and check the quality of the description of the electronicstates obtained. This is typically done by comparing values ofelectron density, atomic partial charges, and band gap to our

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

201

Page 4: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

chemical knowledge of the system, or higher-level quantumchemical calculations.For DFT calculations, the level of approximation of

electronic exchange and correlation effects (EXC) is the maincause of approximation (or inaccuracy). Therefore, the choiceof exchange-correlation functional has a significant influence onhow the energy depends on atomic positions, and thus on allcalculated properties. While many options are available in theliterature and implemented in popular software, the decision onexchange-correlation functional has to be guided by acompromise between accuracy and computational cost. Welike many otherstypically use generalized-gradient approx-imation (GGA) functionals to take into account the nonlocalexchange and correlation effects, which is not possible withlocal density approximation (LDA) functionals. Furthermore,when computationally possible, we employ hybrid functionalswhich provide a better description of the electronic exchange.This is achieved by adding exact Hartree−Fock exchange,mixed with exchange and correlation terms computed at theGGA level.Generally, the appropriate functional is contingent on the

property of interest. In practice, we have found that the PBEfunctional revised for solids (PBESOL)17 in addition to ahybrid version, PBE0 (PBESOL0), produce consistent andphysical results in many porous systems.18,19 Another commonchoice of hybrid functional in the literature is B3LYP:20

although it is usually considered more appropriate formolecular calculations than in the solid state, it has beenwidely used in MOFs and appears to give reliable resultsoverall. In cases of zeolites, which are built from strongerinteractions than MOFs, we note that most properties(structure, energetics, mechanics) are well described alreadywith GGA functionals such as PBESOL. However, forsimulations of complex properties, such as band structures,the use of hybrid functionals is preferred.21

It is well-known, and should be noted here, that DFT is quitepoor at representing long-range correlation effects, typicallyincluding dispersion interactions. When required, dispersioncorrection terms are added to the functional using the D2correction scheme detailed by Grimme and co-workers,22

implemented in the CRYSTAL14 software. The D2 correction is arelatively simple function of interatomic distances. Thisincludes adjustable parameters fitted to interaction andconformational energies computed using high-level wavefunction methods. Subsequently, this correction is added tothe energy obtained by DFT and, as such, does not directlymodify electron density. Notably, there are a number ofalternative dispersion correction methods,23 such as D3 and themany-body-dispersion (MBD) scheme.24,25 Dispersion correc-tions are often essential, and we have found they can stronglyinfluence energetic and mechanical properties, especially forporous structures. Specifically, in MOF systems, dispersioncorrections often improve structural agreement with exper-imental crystal structuresalthough counter-examples can befound because Grimme’s simple and empirical “D2” correctionscheme can lead to overestimation of the dispersive interactionsat intermediate intermolecular distances (6−10 Å). Thus, foreach system, we benchmark the performance of bothapproaches, without and with dispersion, to experimentaldata: in particular, we consider the crystal density, unit cellparameters, and key intermolecular distances (π−π stacking,hydrogen bonds, etc.), crucial quantities typically affected bydispersive interactions.

Another important consideration for DFT calculations ofcrystals is the choice of reciprocal space sampling. This iscontrolled by a shrinking factor, used to generate acorresponding grid of k-points in reciprocal space. Forstructures with very large unit cells, reciprocal space samplingcan be limited to the Γ point; in general, it should be chosen soas to reach convergence on single-point energy. Recently, theimportance of k-point convergence has been clearly describedin simulations of MIL-47(V).26 A rule of thumb we often use,as a starting point, is to have the smallest lattice parametermultiplied by the value of the shrinking parameter be around20−30 Å. In some structures with very asymmetric cells, forexample MIL-47,27 we use direction-dependent shrinkingparameters, to refine the mesh in specific directions. WithinCRYSTAL14, the space group of the system (when known) isconsidered to reduce the computational cost by reducing thenumber of spatial integrals to compute, ideally by the order ofthe space group. This can be very efficient: for example theMOF structure UiO-66, with a cubic cell and space groupF4 3m, and point group symmetries, reduces the CPU time by afactor of approximately 16 (for a single-point energy calculationon 12 processors) (Figure 2).Finally, we note that a simple and useful way to supervise

DFT calculations is to examine, during the self-consistent field(SCF) iteration scheme for the calculation of the density andenergy, the atomic partial charges (calculated by the softwarewith the Mulliken scheme). Because we have a reasonablechemical intuition of these quantities, they provide an initialand rapid feedback as to the quality of the density calculation:when the Si atoms in a zeolite acquire a net charge of 15, thesimulation is certainly diverging! Subsequently, if convergenceseems unreachable, with standard minimization schemes andparameters, one can modify the initial population of orbitals forthe first SCF cycle as to increase the damping of the SCFalgorithm, to achieve self-consistence. Alternatively, theBroyden scheme for self-consistent field calculations is veryrobust.28,29 Mulliken population analysis, commonly used toobtain atomic partial charges, has explicit basis set depend-ence30 and may not display good accuracy. As atomic partialcharges are nonphysical quantities, there are complications andsubjectivity to calculation of these values. Thus, there are anumber of alternative methods which are more suited toaccurately describing porous systems.31 However, to simplyverify calculations, we find the Mulliken method is adequate.The quantum mechanical simulations conducted by our

group primarily use the CRYSTAL14 software, which is unique intreating periodic systems with a crystalline local orbital method.However, we note many groups employ simulation codes, suchas Quantum Espresso33 and VASP,34 which use a plane-waveapproach. This method implicitly treats periodicity in thedescription of the wave function. While both approaches haveparticular advantages and disadvantages, for example, plane-wave methods require a large number of basis functions(augmenting significantly memory requirements) and crystal-line local orbitals do not efficiently treat low symmetrysystems.35 In particular, researchers have employed plane-wave approaches to investigate many nanoporous materialproperties, including the reactivity of zeolite catalysts36 andelectrical conductivity in MOFs.37

Calculation of Structural and Mechanical Properties.We highlight here the typical output we obtain from DFTcalculations of porous crystals. For further detail, one can find

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

202

Page 5: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

comprehensive reviews on the many properties calculated usingDFT simulation.38−40

The most common resultand arguably the most crucialresultof DFT calculations on porous crystals is the optimizedstructure, including cell parameters and atomic positions.Geometry optimizations (CRYSTAL14 keyword OPTGEOM)consist of iteratively updating the structure and subsequentlycomputing energies and forces. Notably, initial structures arechosen from known structures such as crystallographic data;otherwise, several plausible structures are considered to samplethe phase space. This iterative procedure is repeated untilconvergence criteria are met, i.e. until atomic displacements andforces are smaller than a chosen threshold. To determine thedirection and amplitude of each step (in the R3N space ofatomic coordinates, with N the number of atoms per cell), wegenerally use a numerical quasi-Newton method, the Broyden−Fletcher−Goldfarb−Shanno algorithm. We use the conver-gence criteria standard in CRYSTAL14 (9 × 10−4 Ry/bohr oneach component of the energy gradient). Occasionally tightercriteria are used to ensure higher accuracy, typically forproperties linked to derivatives of the energy (vibration modesor mechanical properties, see below). The number of time steps

needed to reach convergence can vary between 10 and 20,typical for small, high-symmetry systems with rather smoothenergy landscapes. However, this can be >200 in problematiccases, or if the initial structure is obviously far from optimal.This occurs if simulations begin with a poorly resolvedexperimental structure or atoms are in unrealistic proximity.Furthermore, as optimizations are constrained to the spacegroup symmetry, we often relax space group requirements andperturb the system. Consequently, we ensure a minimum oflower energy, with a lower symmetry, cannot be found and thespace group constraint is appropriate.In the case of mechanical properties calculations (keyword

ELASTCON), the main output is the 6 × 6 tensor of second-order elastic constants Cij in Voigt notation. These arecomputed by performing small deformations along each ofthe 6 deformation modes, or a subset of those in highsymmetry systems (2 modes are enough for cubic spacegroups). We then analyze it further with tensor analysis tools41

such as the ELATE web application32 (http://progs.coudert.name/elate) developed in our group to investigate themechanical properties of the material studied, including bothvolumetric and anisotropic properties. The primary propertiesare the bulk modulus, in addition to direction-dependentYoung’s moduli, linear compressibilities, shear moduli, andPoisson’s ratios.18,42 These calculations are generally morecostly than geometry optimizations, with the computation timedepending not only on the system size but also on a number ofother factors: the number of deformation modes to beperformed, from 2 in cubic systems to 6 in triclinic systems;the point group symmetries remaining after individualdeformations (for an orthorhombic system, shearing modestake much more time than compression modes, which do notbreak any symmetry); and the number of steps needed to reachconvergence, after each cell deformation, which can bedecreased by decreasing the deformation amplitude. We haveincluded a representative input file, used for mechanicalproperty simulations with CRYSTAL14, in the SupportingInformation.Aside from elastic constants, we also use DFT calculations to

analyze vibrational modes, examining frequencies and modeeigenvectors. Periodic structures have a number of quantizedmodes of vibration, in which the lattice uniformly oscillates.These normal vibrational modes, also called phonons, influencea number of physical properties, such as thermal and electricalconductivity.43 The starting point, for these simulations, mustbe a properly optimized structure, a local minimum in the spaceof atomic coordinates. A standard approach, generally sufficientfor systems with large unit cells (cell parameters ≥ 15 Å) is tocompute harmonic vibrations (keyword FREQCALC), whereall atoms with identical fractional coordinates in neighboringcells vibrate in phase. Specifically, these calculations allow forthe estimation of the contribution of vibrations to the entropyas described in eq 8, where β = 1/kBT.

44 This description canbe generalized to other thermodynamic quantities, includingheat capacity. In addition, we estimate Gruneisen parameters,which in turn give access to the thermal expansion coefficient.45

∑ βε= − − +

−βε

βε−

⎡⎣⎢

⎤⎦⎥S k ln(1 e )

e 1i

Ni

vib B

3i

i (8)

Vibrational mode calculations consists of 3N independentsteps (or usually less, thanks to symmetry relations). At eachstep, one atomic coordinate is slightly modified (typically by 0.3

Figure 2. An example of the use of density functional theory (DFT)calculations for porous materials: calculation of (a) the structure of adefective UiO-66(Zr) MOF with partial substitution of benzenedicarboxylate linkers by formate capping ligands;18 (b) its directionalYoung’s modulus, in units of GPa, plotted by the ELATE software32

from the DFT-computed elastic constants.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

203

Page 6: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

pm) and the resulting energy variation and forces are computedtaking into account a possible lowering of symmetry. Even forhigh-symmetry systems, these computations can be computa-tionally quite demanding since most deformations break mostsymmetries; for UiO-66,46 while a single-point and forcecalculation takes about 4 min (using 12 processors), the wholecalculation of vibration frequencies takes about 11 h, perform-ing only 18 deformations. For systems with small unit cells, it issafer to compute phonon frequencies on a finite number ofpoints of the Brillouin zone (keywords SCELPHONO andDISPERSI), rather than merely at the Γ point. Commonly,vibrational calculations result in identification of imaginaryfrequencies, a nonphysical result and indication the structure isnot a local minima. To address such spurious frequencies, wesuggest increasing the DFT convergence to extremely high-precision, such that frequencies may be more accuratelyassigned to zero or positive frequencies. However, this mayalso be an indication that the initial configuration was not welloptimized.These types of simulations are usually tractable for several

hundreds of atoms per unit cell on 16 cores. Except for specificcases, the parallelization becomes inefficient when exceeding 20to 30 processors. For example, when the shrinking parameter ishigher, using more cores becomes more efficient. Thesesimulations can be run on a desktop workstation with morethan 16 cores, local computing facilities, or national high-performance computing (HPC) centers. The largest systemever treated in the group, at the national HPC scale, is themetal−organic framework DUT-49 with 1728 atoms in aconventional cell of 46.7 Å parameter (cubic space group Pa3 ,72 atoms in the asymmetric unit).47

Ab Initio Molecular Dynamics. In addition to the study ofstructures and “local” properties (often called “zero Kelvin”properties, because they do not account for thermal motionsand entropy), the finite-temperature motion of condensedmatter systems can be examined at the quantum chemical levelusing ab initio molecular dynamics (AIMD). Although muchless frequent than local DFT calculations because of its muchhigher computational cost, we consider AIMD to be anessential tool to probe dynamical properties in situations wheresubnanosecond dynamics are relevant. It is applicable to allcondensed matter systems, not limited to crystals but alsoincluding interfaces and complex liquids, as demonstrated in arecent study of the transport mechanism of carbon dioxide inmolten carbonates.48 We highlight here our use of AIMD tostudy MOFs and zeolitic frameworks as well as porousmaterials where adsorbates are present, using the CP2Ksimulation packagea general open source quantum chemistryand solid state physics simulation package, available at https://www.cp2k.org.The idea behind molecular dynamics (MD) is to explore the

dynamic motion of molecules by numerically solving Newton’sequations of motion:

= −∇ =r U r mr

tF( ) ( )

ddi

Ni

Ni

i2

2 (9)

The force, Fi, experienced by an atom i of mass mi and positionri is equal to the negative of the gradient, ∇i, of the potentialenergy of the system defined by electron density, U(rN).Subsequently, Newton’s second law relates the force to theparticle’s mass multiplied by the second derivative of itsposition, ri, with respect to time, t, i.e. the particle’s acceleration.

In ab initioMD, the atomic forces are calculated at the quantumchemical level, typically using DFT. CP2K uses the Gaussianand plane wave (GPW) method or a Gaussian and augmentedplane wave (GAPW) method to accurately describe theelectronic density within the system.49 VandeVondele et al.have reported a complete and thorough description of thesemethods used by CP2K and other codes.50,51

The motion of the nuclei follows Newton’s equation, and isintegrated numerically with a finite size time step Δt. At eachtime step the wave function is optimized in the DFTframework, thus combining the accuracy of ab initio calculationsand the ability to study time dependent phenomena. Thismethod is thus much more costly that static calculations, butcan be used to study mechanisms where bond cleavage andbond formation are essential or where no classical approx-imation (force field) of the intra- and intermolecularinteractions is available.Simulations in several different thermodynamic ensembles

can be performed: constant number of particles, volume, andtemperature (N, V, T); constant number of particles, pressure,and temperature (N, P, T). Moreover, the simulation of rareevents such as chemical reactions and infrequent diffusion usingAIMD is challenging. By combining AIMD with constraineddynamics and metadynamics methods, we and otherresearchers have been able to elucidate properties relating torare events with high accuracy.52,53 The output from thesimulations is typically the system’s wave functionalthough itis costly to store on disk due to its large sizeand populationcharges on each atom (and for each time step). AIMDsimulations primarily result in the positions, the velocities, andthe forces on each atom of the system as well as theinstantaneous temperature, stress tensor, and, for (N, P, T)simulations, cell parameters.By analysis of the resulting trajectories we can calculate a

number of properties. First, pair distribution functions (PDFs)which, when all atoms are taken into account, can be comparedwith PDF from X-ray diffraction analyses. This is useful forvalidating simulations. By examining individual PDFs, we areable to identify the relevant interactions, which is not possiblefrom experimental data. Moreover, statistical analyses are usedto quantify events such as bond cleavage and formation;specifically, this was used to study the mechanism ofpyrocarbonate formation and its separation into carbon dioxideand carbonate anion.48 Additionally, computing coordinationnumbers aids in monitoring the stability of coordinationnetworks.54 Vibrational modes within the framework or in thepresence of a guest molecule can be obtained by effectivenormal-mode analysis,55 and preferred sites of adsorption arefound by simple analysis of the spatial distribution ofadsorbates. Extraction of fundamental free energy profiles isachieved by identifying a potential of mean force or from moresophisticated approaches using constraint molecular dynamicssuch as the Blue Moon Ensemble approach.56,57 Thesemethods are quite powerful; for example, we have recentlyused them to uncover the mechanism by which a carbondioxide molecule can permeate a zeolite where cations blocknarrow pores for other gases (Figure 3).Importantly, these methods are quite costly in terms of

computational power. For simulations in the (N, V, T)ensemble of MOFs with approximately 300 atoms, we cansimulate between 2000 and 6000 fs in 10 h using 256 cores onnational high-performance computing systems. Simulations inthe (N, P, T) ensemble are often three times slower than

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

204

Page 7: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

(N,V,T) simulations, as the temperature and pressure convergemuch slower. Despite this, we were able to accurately describetwo structurally different phases of a very flexible MOF.58 Inparticular, massive parallelization is not possible for this type ofcalculation; specifically, an increase from 64 to 128 cores has anefficiency of about 40%, and from 128 to 256 cores about 50%,but increasing from 256 to 512 cores can even make thecalculation 20% slower. Scaling studies are crucial beforerunning simulations in order to be computationally efficient, orat least not to waste computational resources.During simulations, we check that the chosen parameters

allow us to represent the system correctly. Conservation ofenergy and other conserved quantities are essential, in additionto convergence of temperature, as it shows that the thermostathas the desired effect, and the same goes for the pressure(keeping in mind that pressure fluctuations on such smallsystems can be very large (∼10000 bar for a volume of 4000Å3)). We also check that the cell parameters and the chargesremain plausible and they do not diverge or converge toobviously nonphysical values. The number of self-consistentfield steps needed to converge the energy at each time step isalso a good hint of whether or not the input parameters, such asthe time step, have been well-chosen. It is also crucial to checkthat no unrealistic phenomena occur, such as a zeoliteframework breaking covalent bonds in usual conditions oftemperature and pressure.Additionally, we have attached a representative input file for

an MD simulation in the Supporting Information. Thiscorresponds to a MD run for zeolite Na-RHO with carbondioxide adsorbed.

■ CLASSICAL SIMULATIONHerein we have shown the wealth of information obtainable byquantum mechanical simulation; however, for longer timescales and larger systems, classical molecular simulations areemployed. Typically, classical simulations can be performedroutinely on systems containing thousands of atoms, or in moreextreme cases for systems up to 100 000 atoms59althoughdue to their periodic nature, it is not common in simulations ofcrystalline materials. Simulation times in classical dynamics canalso be much larger that in ab initio MD, ranging from a fewnanoseconds to hundreds of nanoseconds. This allows

molecular level insight into macroscopic material propertiessuch as adsorption, diffusion, and framework dynamics.60−62

Force Field Parameters. Potential energy (U(rN)) inclassical simulations is calculated from an interatomic potentialenergy function that is described by parameters from the forcefield. Importantly, the choice and implementation of force fieldsis crucial to the accuracy obtained by this method.Force fields are parametrized such as to reproduce the

molecular geometry or thermodynamic properties reportedexperimentally or described by higher level ab initiocalculations. The general functional form of the potentialenergy function in classical molecular simulation includesbonded terms for interactions of atoms that are linked bycovalent bonds (or metal−ligand bonds), and nonbonded ornoncovalent terms that describe long-range electrostatic andvan der Waals forces illustrated in Figure 4. The specific

decomposition of the terms depends on the force field applied,but a general form for the total energy in an additive force fieldis illustrated in eqs 10, 11, and 12.

= +U U Utotal bonded nonbonded (10)

= + +U U U Ubonded bond angle dihedral (11)

= +U U Unonbonded electrostatic van der Waals (12)

There are a great variety of options for force fields, with earlyexamples developed to reproduce the geometry of small organicmolecules and later adapted to treat more complex function-alized molecules.63 Moreover, there are a number of generalforce fields that have been developed to treat all atoms in theperiodic table,64 biomolecules,65 and condensed matter.66

Simulations employing these general force fields have beenshown to be applicable to many porous systems; in particular,universal force field (UFF)64 parameters have been ubiquitousin calculating intermolecular energies (Uvan der Waals) for gasadsorption simulations.67

Force fields for porous structures with a high degree offlexibility require good accuracy in the description ofintramolecular interactions in the structure, in particular forthe low-frequency phonon modes present, which is often notcaptured by generic force fields.68 To this end, researchers haveused ab initio simulation to derive accurate system specific forcefields. van Speybroeck and co-workers have developed QuickFFfor the easy derivation of new force fields from ab initiosimulations,69 which has been used to describe a number of

Figure 3. AIMD simulations allowed us to uncover the mechanism bywhich a carbon dioxide molecule can permeate a zeolite where cationsblock narrow pores for other gases by taking advantage of cationsthermal motion.52 In gray: Na-RHO zeolitic framework; blue: Na+

cations; CO2 (cyan and red) guest molecules can bypass the sodiumions and diffuse through the zeolite’s windows, while methane (cyanand white) cannot.

Figure 4. Examples of the structural parameters explicitly described byforce field models used in classical molecular simulations.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

205

Page 8: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

flexible MOFs and explore thermodynamics.70 Alternatively,Heine et al. have extended UFF for use in MOF systems byrevising and adding parameters relating to common transitionmetal nodes.71 Nevertheless, the choice of force field applied inclassical simulation must be initially tested against experimentalor quantum chemistry methods to demonstrate the accuracy ofthe method.Methodologies and Implementation. Though classical

simulation, like quantum mechanical simulation, can be used toobtain single-point energies and local minima, we primarily useclassical methods to investigate the dynamics of molecules andadsorption processes.Molecular dynamics is achieved by the same iterative process

described previously for AIMD. However, owing to the cheapercomputational costs of classical simulation, we can followprocesses on a larger scale, such as water confined in zeolites.72

In addition to the choice of force field, as discussed previously,the options of time step and thermostat (or in the case of (N, P,T) simulations thermostat and barostat) are crucial. Forgreatest computational efficiency, the time step should bechosen as large as possible such to minimize the number ofcalculation steps for a given time period. However, the timestep must be verified, so that it does not result in drifts or largefluctuations in energy and other conserved quantities.73

Notably, this is also true for AIMD simulations. Moreover,molecular dynamics explicitly represents the (N, V, E)ensemble; thus, to compare to experimental systems, dynamicconstraints (thermostat and barostat), applied to the movementor particles or cell parameters, are added. This results in a fixedaveraged value of temperature or pressure resulting in the (N,V, T) and (N, P, T) ensembles. There are many approaches tofixing these averages, each with particular advantages anddisadvantages that must be considered before application.74,75

In particular, we routinely use the Nose−Hoover thermostatand barostat.76−78 When using this method, there is a choice ofrelaxation time, which defines the time scale for the relaxationof temperature or pressure. We often choose relaxation times of1−5 ps, ensuring the temperature and pressure do not fluctuatesignificantly and the equilibration time is not unnecessarilylong.The adsorption process can be simulated by Monte Carlo

(MC) simulations which consist of trial displacement, insertion,and removal of gas molecules in the framework structure.79 Thecriteria for accepting a trial displacement are often described bythe Metropolis algorithm.80 In the Metropolis algorithm in asimple application, the below steps are followed until a move isaccepted.

1. Calculate the potential energy (Ui) of the initial state.2. Choose a trial displacement of a random molecule from a

uniform random distribution.3. Calculate the potential energy of the new state (Uf).4. If Uf ≤ Ui, accept the move.5. If Uf > Ui, select a random number, w, where w ∈ [0,1]

from a uniform distribution and if exp[−β(Uf − Ui)] > w,accept the move.

6. Repeat steps 2−5.In grand-canonical Monte Carlo, the number of molecules

adsorbed in the porous structure evolves over the course of thesimulation, and after a period of equilibration, it will fluctuatearound an equilibrium value imitating the experimental system,in which the adsorbed phase is at equilibrium with a gasreservoir. Conducting Monte Carlo simulations can be very

straightforward; however, there are a number of importantoptions and analyses to consider before conducting asimulation, including the relative probability of different trialmoves (rotation, displacement, reinsertion, and more) andensuring convergence in energy or number of adsorbedparticles. Dubbeldam et al. have produced an excellent resourcefor the application of MC in this area.79

In particular, we have employed classical molecular dynamicsto examine the stability of 18 zeolitic imidazolate frameworks(ZIFs), displayed in Figure 5.81 By using a classical approach,

we were able to simulate the framework dynamics of the 18frameworks at multiple temperatures for more than 5 ns, a scalenot computationally feasible using ab initio methods. Unit cellfluctuations were measured in response to changes intemperature and pressure to quantify the relative stability ofthe topologies. Notably, we observed a number of the examinedZIFs displaying significant negative thermal expansion (Figure5), and many of the included hypothetical topologies are notmechanically stable at reasonable temperatures.As computational chemists, our findings rely on software

implementations of the simulation algorithms described brieflyabove. The two main categories of software used in simulationare small, homegrown software that are used at the scale of oneor two research groups, and very big software, used byhundreds (if not thousands) of researchers. Most of theclassical molecular dynamics codes around belong to thesecond category, such as LAMMPS,82 GROMACS,83 and DL-POLY.84 This situation arose mainly because of the complexityof the algorithms used in molecular dynamics, where very-large-scale systems are simulated on supercomputers, and significant

Figure 5. Components of zeolitic imidazolate frameworks (ZIFs) andtheir possible topologies examined (a). Resulting volume thermalexpansion coefficients, simulated by classical molecular dynamics, forthe ZIF frameworks (b). This is compared to values reported in theliterature for other metal−organic frameworks, portrayed in a darkershade. This figure was adapted from ref 81.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

206

Page 9: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

development effort is required to get an efficient implementa-tion. On the other hand, most publicly available MC codes,such as RASPA85 and many others, are homegrown software.This can be explained by the diversity of the algorithms, and thecapacity to tailor a specific MC move for a specific system. Forexample, the GIBBS code86 was initially developed for thesimulation of hydrocarbons in zeolites, and as such proposesspecific MC moves such as chain reptation, partial rotation ofmolecules, and biased insertion/deletion of chain molecules.Sometimes, the existing software does not provide what we

need for our studies. For example, the coupling betweenadsorption and deformation of the guest framework is hard tostudy with the current methods. This coupling is at the originof the complex and fascinating behaviors of some MOF,including breathing,87 gate-opening,88 or more complexbehaviors.47 The natural ensemble for simulating thesephenomena is the osmotic ensemble,89 which is an extensionof the grand-canonical ensemble. Molecular dynamics cannotbe used easily for simulations in the grand-canonical ensemble,because most of the algorithms rely on an Hamiltoniandynamic, which is hard to obtain with a varying number ofparticles. (Some schemes have been proposed for grand-canonical molecular dynamics,90,91 but it is not yet clearwhether the ensemble simulated using these scheme is thegrand-canonical ensemble or not.) On the other hand, MCmethods can easily sample the grand-canonical ensemble, buthave a lot more difficulties to reproduce large amplitude andcollective movements, which naturally occur in thesephenomena. One solution to get efficient sampling of theosmotic ensemble is to use a hybrid Monte-Carlo simulation,where short runs of molecular dynamics are used as MC moves.Unfortunately, hybrid MC simulations are not implemented

in any of the current widely used molecular simulation software,and implementing such a feature in existing codes wouldnecessitate large flexibility of the code base.92 As a result, wehave begun to write a new simulation code, with the explicitgoals of being modular and to ease the new methods. We aimto make this code easy to read and to modify, and reasonablyfast. At the time of writing, the code is able to perform energyminimization, molecular dynamics simulations in the (N, V, E),(N, V, T), and (N, P, T) ensemble, MC simulations in the (N,V, T), (N, P, T), grand-canonical ensemble, and hybrid MCsimulation in all ensembles, including the osmotic ensemble. Allthese simulation methods share the same implementation ofenergy evaluation, including Ewald sum and Wolf sum for thetreatment of electrostatic interactions.93,94 The software isundergoing a cleaning phase before we share it openly with thecommunity.Currently the development of scientific software is often

overlooked, with software bugs that either give wrong results orprevent reproducibility.95 To improve software quality andmaintainability, scientists can and should use proven techniquesfrom software engineering.96 These techniques are modernsanswers to an old problem: humans are fallible, how do wemake sure that the code is correct? Methods we use for thedevelopment of our in-house software include version controlusing git97 and Github;98 unit tests, which check the behavior ofone function only; and integration tests, which test thatmultiple parts of the code work together (for example, is theenergy constant during a molecular dynamics simulation, do weget the pressure of the system right in (N, P, T) MC). All thesetests are run each time new code is written, using multiplecompilers and operating systems. Services like Travis99 provide

server farms for free for open-source projects to build and runtests using Linux or OS X and various compilers (GCC andclang). Additional code review ensures that the code is read andunderstood by different persons, catching some additionalmistakes. Together, theses methods help to ensure thereproducibility of the simulations and improve the usability ofthe resulting software by removing a lot of bugs.

■ MESOSCALE SIMULATIONWe have shown how characterization at the quantum andatomistic scale can elucidate interesting phenomena exhibitedby porous materials. However, there are limits to the efficiencyand system size accessible by these methods. To achieveexamination of the mesoscopic properties of porosity and large-scale transport processes, we use a number of tools which takefurther approximations to the chemistry structure.

Pore Structure Identification. Surface area and poremetrics are vital characteristics of porous materials, particularlyfor applications in gas storage and separation, where theperformance of materials is often correlated with thesequantities.100 Notably, MOFs are reported to have very highsurface areas, over 7000 m2 g−1;101 however, surface area cannotbe measured directly from experiments. The surface area of aporous material is commonly obtained by application ofBrunauer−Emmett−Teller (BET) theory to Ar or N2adsorption isotherms.102,103 The BET method relies upon anumber of assumptions: adsorption occurs on a homogeneoussurface, no lateral interactions occur between adsorbedmolecules, and Langmuir theory is applicable to each adsorbedlayer. While for a number of cases these assumptions may nothold true,104 and while such surface area measurements give nomicroscopic insight into the nature of the porosity, being astandard analysis technique, they offer a widely used, simple,and reproducible characterization tool for nanoporous materi-als.There are a number of algorithms and tools that are used to

obtain pore metrics from a well-defined periodic structure. Forthe most direct method, the surface area and pore volume canbe calculated geometrically from the crystal structure.105

Popular through application in the Accelrys Materials Studio,the surface area of a material corresponds to the area traced outby the center of a probe particle as the probe particle is movedacross the surface of the framework atoms, illustrated in Figure6a. These accessible areas calculated with a probe moleculeequivalent to the size of nitrogen are in good agreement withthe experimental BET surface areasalthough exceptions canarise because of the dynamic nature of the framework, or poreblocking by guest molecules due to incomplete activation of theexperimental samples. Alternatively, there are a number ofmore complex methods to describe the detailed porosity of aperiodic system including Poreblazer,106 ZEOMICS,107 andZeo++.108 Each of these methods can identify the pore sizedistribution, maximum pore size, and pore limiting diameter toresult in a clear picture of the system’s pore structure. Anexample of the pore metrics obtainable using Zeo++ isrepresented in Figure 6b.The methods outlined here provide a crucial role of bridging

the microscopic structure, elucidated by quantum mechanicalor classical simulation, to the surface area or porosity observedexperimentally.

Fluid Flow and Adsorption. Understanding the interplaybetween transport of species and adsorption in porousmaterials, at a mesoscopic scale, is one of the main steps to

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

207

Page 10: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

improve the efficiency of porous materials for applications suchas ion capture, ion exchange, and phase separation. The numberof atoms or molecules needed to perform simulations withmolecular dynamics or Monte Carlo models at large lengthscales (micrometer) makes atomistic approaches computation-ally unfeasible. Macroscopic methods used in classicalcomputational fluid dynamics can be used to simulate fluidbehavior, but lack in their description of local phenomena, suchas adsorption. In between these two scales, there exist a fewmesoscale lattice-based methods, although they are much lesscommonly used than atomistic simulations and computationalfluid dynamics. In our group, in order to computationally studyliquid flow and adsorption in porous materials, we use aLattice−Boltzmann (LB) model recently expanded to accountfor adsorption.109 Its basic principles are illustrated in Figure 7.The Lattice−Boltzmann scheme is a computational fluiddynamic (CFD) model working on a 3D mesh of equallyspaced nodes, i.e. a grid of cubic voxels. It allows us to computeefficiently the fluid behavior in the material. Physical quantitiesrelated to the fluid flow (such as the fluid velocity field) and theadsorption (concentration of adsorbed and free species) arestored for each node of the mesh. Then, instead of solving theNavier−Stokes equations as is done in standard CFD methods,these fields are numerically integrated in time through adiscrete formulation of the Boltzmann transport equation. Thealgorithm we are using here is based on the schemesummarized by Ladd and Verberg.110 For a given externalforce, equivalent to a drop pressure, it computes the flux anddensity field of fluid. Here we investigate mesoporous materials,and the regime is assumed to be linear with a no slip boundarycondition at the interface solid/liquid.The LB model finds its roots in the lattice gas cellular

automata method and statistical physics with the introductionof a Boltzmann equation.111−113 It works on the propagation ofa single particle velocity distribution function f(r,c,t) represent-ing the probability of one particle to be at position r with thevelocity c at time t. This function is propagated on the latticeusing the discrete velocity model D3Q19 (3 dimensions, 19

speeds). Other velocity models are available, such as theD3Q15 or the D3Q27, but the D3Q19 offers the bestcompromise between speed and precision.114 The dynamics ofthe LB scheme follows this propagation described in eq 13,where f i

e represents the local Maxwell−Boltzmann equilibriumdistribution, τ is the relaxation time, and Fi

ext accounts forexternal forces. For a better understanding of the fundamentalsof the LB model, refer to the work of Succi.115

τ

+ Δ + Δ

= +−

+

f r c t t t

f r tf r t f r t

F

( , )

( , )( ( , ) ( , ))

i i

iie

iiext

(13)

The LB model is coupled with a moment propagationmethod116,117 simulating the motion of species called ”tracers”inside the fluid. The adsorption part of the algorithm, takingplace at the solid/liquid interface, is introduced in the momentpropagation method.109 It has been recently improved toaccount for saturation of the tracers on the adsorption sites andthus reproduce a Langmuir adsorption type.118 The dispersioncoefficient, the diffusion coefficient, the quantity adsorbed, andthe density adsorbed are computed in this part. Thiscoefficients give interesting information about the motion oftracers inside the fluid.The LB scheme employed here has the particularity to be

adimensional. Inputs and output values are expressed as afunction of the distance between nodes and the time step. Weuse an internal code to create input files and compute results.This code includes a library of typical geometries automatically

Figure 6. Representation of the geometric method used to measurethe pore volume and surface area in porous materials (a). Additionally,the pore information obtained by software such as Zeo++, limitingpore diameter (LPD), and maximum pore diameters (MPD) for porechannels and maximum void diameters (MVD) can be found forinaccessible voids present within the structure (b).

Figure 7. 2D graphic representation of the Lattice−Boltzmann schemeexpanded to account for adsorption. The solid part of the material is ingray, the liquid part is in blue, and the colored gradients show theadsorbed density. The figure below is a 3D binarized image of a porousmaterial obtained with X-ray tomography we may use as input for oursimulations.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

208

Page 11: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

generated, such as slit pore, cylinder, honeycomb, and inverseopal.X-ray tomography imaging (see Figure 7) can be easily used

to create an input of real materials with a simple binarizationprocess. The definition of the lattice (3D cubic grid) of the LBscheme corresponds exactly to the voxel definition of a 3Dimage. This type of geometry gives a powerful tool to simulateflow and adsorption in real materials.The code is parallelized using OpenMP. We simulate

geometries up to 64 × 106 nodes on conventional multi-threaded computers. In the case of small geometries (100−1000 nodes), we run a simulation on 4 to 8 cores. For biggeometries (>1,000,000 nodes) we run simulations on 20 to 30cores. We usually use convergence criteria on the mean fluidvelocity, fraction adsorbed, velocity autocorrelation function,and dispersion coefficient. The dispersion coefficient iscomputed from the velocity autocorrelation function. However,we always consider carefully the convergence of the dispersioncoefficient because a good convergence on the velocityautocorrelation function does not necessarily imply aconvergence of the dispersion coefficient. A representativeinput file used for LB simulations is demonstrated in theSupporting Information.

■ CONCLUSION

Computational chemistry is a useful toolbox for understandingthe very large diversity of physical and chemical phenomenathat can be observed among nanoporous materials. Herein, wehave outlined the wide variety of methods our group uses toelucidate their physical and chemical properties, in addition tothose of fluids adsorbed in the materials’ nanosized pores. Wehave described simulations at different “levels of theory” anddifferent time and length scales. Computationally demandingsimulations are used to investigate complex electronicprocesses, by explicitly treating the electrons of each frame-work. Alternatively, larger systems and gas adsorption can beexamined by classical potentials that describe atoms or groupsof atoms. Finally, we have illustrated statistical physics andgeometric approaches that we use to explore porosity and large-scale transport. By using this toolbox of methods, in synergywith experimental observations in general, and in situ and inoperando characterization in particular, one can obtain a broadand complete picture of the system or process, as well asunparalleled understanding into the microscopic behavior ofporous materials.

■ ASSOCIATED CONTENT

*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.chemma-ter.6b02994.

Input files for DFT calculation, ab initio moleculardynamics, and Lattice−Boltzmann simulation (PDF)

■ AUTHOR INFORMATION

Corresponding Author*E-mail: [email protected].

NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSAlthough no original calculations were performed in the writingof this review, a large part of the work reviewed here requiresaccess of scientists in the field to large supercomputer centers.We acknowledge GENCI for high-performance computingCPU time allocations (projects x2016087069 andt201608s037). The group’s work benefited from the supportof ANRT (theses CIFRE 2015/0268 et 2013/1262), PSLResearch University (project DEFORM, grant ANR-10-IDEX-0001-02), and the Ecole normale superieure. F.-X.C. thanksAnne Boutin and Alain Fuchs for a long-standing andcontinuing collaboration on the fascinating topic of modelingof nanoporous materials. J.-M.V. acknowledges Joel Lachambrefor the X-ray tomography acquisition and 3D reconstructionperformed on the sample shown here. We thank the WOS Bar,Le Bombardier, Picard, and De Clercq for providing moralsupport in liquid and solid form.

■ REFERENCES(1) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption andseparation in metal−organic frameworks. Chem. Soc. Rev. 2009, 38,1477.(2) Morris, R. E.; Wheatley, P. S. Gas Storage in NanoporousMaterials. Angew. Chem., Int. Ed. 2008, 47, 4966−4981.(3) Bell, A. T. The Impact of Nanoscience on HeterogeneousCatalysis. Science 2003, 299, 1688−1691.(4) Della Rocca, J. D.; Liu, D.; Lin, W. Nanoscale Metal−OrganicFrameworks for Biomedical Imaging and Drug Delivery. Acc. Chem.Res. 2011, 44, 957−968.(5) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. TheChemistry and Applications of Metal-Organic Frameworks. Science2013, 341, 1230444−1230444.(6) Davis, M. E. Ordered porous materials for emerging applications.Nature 2002, 417, 813−821.(7) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal−Organic Frameworks. Chem. Rev. 2012, 112, 673−674.(8) Coudert, F.-X. Responsive Metal−Organic Frameworks andFramework Materials: Under Pressure, Taking the Heat, in theSpotlight, with Friends. Chem. Mater. 2015, 27, 1905−1916.(9) Sholl, D.; Steckel, J. Density Functional Theory: A PracticalIntroduction; Wiley: 2011.(10) Frenkel, D.; Smit, B. Understanding Molecular Simulation, 2nded.; Academic Press, Inc.: Orlando, FL, USA, 2001.(11) Hoover, W. G. Nonequilibrium Molecular Dynamics. Annu. Rev.Phys. Chem. 1983, 34, 103−127.(12) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys.Rev. 1964, 136, B864−B871.(13) Kohn, W.; Sham, L. J. Self-Consistent Equations IncludingExchange and Correlation Effects. Phys. Rev. 1965, 140, A1133−A1138.(14) Parr, R.; Weitao, Y. Density-Functional Theory of Atoms andMolecules; International Series of Monographs on Chemistry; OxfordUniversity Press: 1994.(15) Lejaeghere, K.; et al. Reproducibility in density functional theorycalculations of solids. Science 2016, 351, aad3000−aad3000.(16) Dovesi, R.; Orlando, R.; Erba, A.; Zicovich-Wilson, C. M.;Civalleri, B.; Casassa, S.; Maschio, L.; Ferrabone, M.; Pierre, M. D. L.;D’Arco, P.; Noel, Y.; Causa, M.; Rerat, M.; Kirtman, B. C RYSTAL14:A program for the ab initio investigation of crystalline solids. Int. J.Quantum Chem. 2014, 114, 1287−1317.(17) Perdew, J. P.; Ruzsinszky, A.; Csonka, G. I.; Vydrov, O. A.;Scuseria, G. E.; Constantin, L. A.; Zhou, X.; Burke, K. Restoring theDensity-Gradient Expansion for Exchange in Solids and Surfaces. Phys.Rev. Lett. 2008, 100, 136406.(18) Thornton, A. W.; Babarao, R.; Jain, A.; Trousselet, F.; Coudert,F.-X. Defects in metalorganic frameworks: a compromise betweenadsorption and stability? Dalton Trans. 2016, 45, 4352−4359.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

209

Page 12: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

(19) Lee, S. J.; Doussot, C.; Baux, A.; Liu, L.; Jameson, G. B.;Richardson, C.; Pak, J. J.; Trousselet, F.; Coudert, F.-X.; Telfer, S. G.Multicomponent MetalOrganic Frameworks as Defect-TolerantMaterials. Chem. Mater. 2016, 28, 368−375.(20) Ortiz, A. U.; Boutin, A.; Fuchs, A. H.; Coudert, F.-X.Metalorganic frameworks with wine-rack motif: What determinestheir flexibility and elastic properties? J. Chem. Phys. 2013, 138,174703.(21) Heyd, J.; Scuseria, G. E. Efficient hybrid density functionalcalculations in solids: Assessment of the Heyd−Scuseria−Ernzerhofscreened Coulomb hybrid functional. J. Chem. Phys. 2004, 121, 1187.(22) Grimme, S. Semiempirical GGA-type density functionalconstructed with a long-range dispersion correction. J. Comput.Chem. 2006, 27, 1787−1799.(23) Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C.Dispersion-Corrected Mean-Field Electronic Structure Methods.Chem. Rev. 2016, 116, 5105−5154.(24) Moellmann, J.; Grimme, S. DFT-D3 Study of Some MolecularCrystals. J. Phys. Chem. C 2014, 118, 7615−7621.(25) Tkatchenko, A.; DiStasio, R. A.; Car, R.; Scheffler, M. Accurateand Efficient Method for Many-Body van der Waals Interactions. Phys.Rev. Lett. 2012, 108, 236402.(26) Vanpoucke, D. E. P.; Lejaeghere, K.; Speybroeck, V. V.;Waroquier, M.; Ghysels, A. Mechanical Properties from Periodic PlaneWave Quantum Mechanical Codes: The Challenge of the FlexibleNanoporous MIL-47(V) Framework. J. Phys. Chem. C 2015, 119,23752−23766.(27) Barthelet, K.; Marrot, J.; Riou, D.; Ferey, G. A Breathing HybridOrganic−Inorganic Solid with Very Large Pores and High MagneticCharacteristics. Angew. Chem., Int. Ed. 2002, 41, 281−284.(28) Broyden, C. G. A class of methods for solving nonlinearsimultaneous equations. Mathematics of Computation 1965, 19, 577−577.(29) Johnson, D. D. Modified Broyden’s method for acceleratingconvergence in self-consistent calculations. Phys. Rev. B: Condens.Matter Mater. Phys. 1988, 38, 12807−12813.(30) Reed, A. E.; Weinstock, R. B.; Weinhold, F. Natural populationanalysisa. J. Chem. Phys. 1985, 83, 735.(31) Manz, T. A.; Sholl, D. S. Chemically Meaningful AtomicCharges That Reproduce the Electrostatic Potential in Periodic andNonperiodic Materials. J. Chem. Theory Comput. 2010, 6, 2455−2468.(32) Gaillac, R.; Pullumbi, P.; Coudert, F.-X. ELATE: an open-sourceonline application for analysis and visualization of elastic tensors. J.Phys.: Condens. Matter 2016, 28, 275201.(33) Giannozzi, P.; et al. QUANTUM ESPRESSO: a modular andopen-source software project for quantum simulations of materials. J.Phys.: Condens. Matter 2009, 21, 395502.(34) Hafner, J. Ab-initio simulations of materials using VASP:Density-functional theory and beyond. J. Comput. Chem. 2008, 29,2044−2078.(35) Wills, J. M.; Eriksson, O.; Andersson, P.; Delin, A.; Grechnyev,O.; Alouani, M. Full-Potential Electronic Structure Method; Springer:Berlin Heidelberg, 2010.(36) Sierka, M.; Sauer, J. Structure and reactivity of silica and zeolitecatalysts by a combined quantum mechanics[ndash ]shell-modelpotential approach based on DFT. Faraday Discuss. 1997, 106, 41−62.(37) Sun, L.; Hendon, C. H.; Minier, M. A.; Walsh, A.; Dinca, M.Million-Fold Electrical Conductivity Enhancement in Fe 2 (DEBDC)versus Mn 2 (DEBDC) (E = S, O). J. Am. Chem. Soc. 2015, 137,6164−6167.(38) Odoh, S. O.; Cramer, C. J.; Truhlar, D. G.; Gagliardi, L.Quantum-Chemical Characterization of the Properties and Reactivitiesof Metal−Organic Frameworks. Chem. Rev. 2015, 115, 6051−6111.(39) van Speybroeck, V.; Hemelsoet, K.; Joos, L.; Waroquier, M.;Bell, R. G.; Catlow, C. R. A. Advances in theory and their applicationwithin the field of zeolite chemistry. Chem. Soc. Rev. 2015, 44, 7044−7111.

(40) Coudert, F.-X.; Fuchs, A. H. Computational characterizationand prediction of metal−organic framework properties. Coord. Chem.Rev. 2016, 307, 211−236.(41) Marmier, A.; Lethbridge, Z. A.; Walton, R. I.; Smith, C. W.;Parker, S. C.; Evans, K. E. ElAM: A computer program for the analysisand representation of anisotropic elastic properties. Comput. Phys.Commun. 2010, 181, 2102−2115.(42) Ortiz, A. U.; Boutin, A.; Fuchs, A. H.; Coudert, F.-X.Anisotropic Elastic Properties of Flexible Metal-Organic Frameworks:How Soft are Soft Porous Crystals? Phys. Rev. Lett. 2012, 109, 195502.(43) Simon, S. H. The Oxford solid state basics; Oxford Univ. Press:Oxford, UK, 2013.(44) Fultz, B. Vibrational thermodynamics of materials. Prog. Mater.Sci. 2010, 55, 247−352.(45) Gava, V.; Martinotto, A. L.; Perottoni, C. A. First-PrinciplesMode Gruneisen Parameters and Negative Thermal Expansion in α -ZrW 2 O 8. Phys. Rev. Lett. 2012, 109, 195503.(46) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.;Bordiga, S.; Lillerud, K. P. A New Zirconium Inorganic Building BrickForming Metal Organic Frameworks with Exceptional Stability. J. Am.Chem. Soc. 2008, 130, 13850−13851.(47) Krause, S.; Bon, V.; Senkovska, I.; Stoeck, U.; Wallacher, D.;Tobbens, D. M.; Zander, S.; Pillai, R. S.; Maurin, G.; Coudert, F.-X.;Kaskel, S. A pressure-amplifying framework material with negative gasadsorption transitions. Nature 2016, 532, 348−352.(48) Corradini, D.; Coudert, F.-X.; Vuilleumier, R. Carbon dioxidetransport in molten calcium carbonate occurs through an oxo-Grotthuss mechanism via a pyrocarbonate anion. Nat. Chem. 2016, 8,454−460.(49) Lippert, G.; Hutter, J.; Parrinello, M. The Gaussian andaugmented-plane-wave density functional method for ab initiomolecular dynamics simulations. Theor. Chem. Acc. 1999, 103, 124−140.(50) VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.;Chassaing, T.; Hutter, J. Quickstep: Fast and accurate densityfunctional calculations using a mixed Gaussian and plane wavesapproach. Comput. Phys. Commun. 2005, 167, 103−128.(51) VandeVondele, J.; Hutter, J. Gaussian basis sets for accuratecalculations on molecular systems in gas and condensed phases. J.Chem. Phys. 2007, 127, 114105.(52) Coudert, F.-X.; Kohen, D. Molecular Insight into CO2 TrapdoorAdsorption in Zeolite Na-RHO. Submitted for publication, 2016.(53) De Wispelaere, K. D.; Ensing, B.; Ghysels, A.; Meijer, E. J.;Speybroeck, V. V. Complex Reaction Environments and CompetingReaction Mechanisms in Zeolite Catalysis: Insights from AdvancedMolecular Dynamics. Chem. - Eur. J. 2015, 21, 9385−9396.(54) Haigis, V.; Coudert, F.-X.; Vuilleumier, R.; Boutin, A.; Fuchs, A.H. Hydrothermal Breakdown of Flexible Metal−Organic Frameworks:A Study by First-Principles Molecular Dynamics. J. Phys. Chem. Lett.2015, 6, 4365−4370.(55) Martinez, M.; Gaigeot, M.-P.; Borgis, D.; Vuilleumier, R.Extracting effective normal modes from equilibrium dynamics at finitetemperature. J. Chem. Phys. 2006, 125, 144106.(56) Carter, E.; Ciccotti, G.; Hynes, J. T.; Kapral, R. Constrainedreaction coordinate dynamics for the simulation of rare events. Chem.Phys. Lett. 1989, 156, 472−477.(57) Sprik, M.; Ciccotti, G. Free energy from constrained moleculardynamics. J. Chem. Phys. 1998, 109, 7737.(58) Haigis, V.; Belkhodja, Y.; Coudert, F.-X.; Vuilleumier, R.;Boutin, A. Challenges in first-principles NPT molecular dynamics ofsoft porous crystals: A case study on MIL-53(Ga). J. Chem. Phys. 2014,141, 064703.(59) Evans, J. D.; Huang, D. M.; Hill, M. R.; Sumby, C. J.; Sholl, D.S.; Thornton, A. W.; Doonan, C. J. Molecular Design of AmorphousPorous Organic Cages for Enhanced Gas Storage. J. Phys. Chem. C2015, 119, 7746−7754.(60) Evans, J. D.; Coudert, F.-X. Microscopic Mechanism of ChiralInduction in a Metal−Organic Framework. J. Am. Chem. Soc. 2016,138, 6131−6134.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

210

Page 13: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

(61) Evans, J. D.; Huang, D. M.; Hill, M. R.; Sumby, C. J.; Thornton,A. W.; Doonan, C. J. Feasibility of Mixed Matrix Membrane GasSeparations Employing Porous Organic Cages. J. Phys. Chem. C 2014,118, 1523−1529.(62) Trousselet, F.; Boutin, A.; Coudert, F.-X. Novel PorousPolymorphs of Zinc Cyanide with Rich Thermal and MechanicalBehavior. Chem. Mater. 2015, 27, 4422−4430.(63) Allinger, N. L.; Chen, K.; Lii, J.-H. An improved force field(MM4) for saturated hydrocarbons. J. Comput. Chem. 1996, 17, 642−668.(64) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.;Skiff, W. M. UFF, a full periodic table force field for molecularmechanics and molecular dynamics simulations. J. Am. Chem. Soc.1992, 114, 10024−10035.(65) Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K.M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.;Kollman, P. A. A Second Generation Force Field for the Simulation ofProteins, Nucleic Acids, and Organic Molecules. J. Am. Chem. Soc.1995, 117, 5179−5197.(66) Sun, H. COMPASS: An ab Initio Force-Field Optimized forCondensed-Phase ApplicationsOverview with Details on Alkane andBenzene Compounds. J. Phys. Chem. B 1998, 102, 7338−7364.(67) Getman, R. B.; Bae, Y.-S.; Wilmer, C. E.; Snurr, R. Q. Reviewand Analysis of Molecular Simulations of Methane, Hydrogen, andAcetylene Storage in Metal−Organic Frameworks. Chem. Rev. 2012,112, 703−723.(68) Bureekaew, S.; Amirjalayer, S.; Tafipolsky, M.; Spickermann, C.;Roy, T. K.; Schmid, R. MOF-FF - A flexible first-principles derivedforce field for metal-organic frameworks. Phys. Status Solidi B 2013,250, 1128−1141.(69) Vanduyfhuys, L.; Vandenbrande, S.; Verstraelen, T.; Schmid, R.;Waroquier, M.; van Speybroeck, V. QuickFF: A program for a quickand easy derivation of force fields for metal-organic frameworks fromab initio input. J. Comput. Chem. 2015, 36, 1015−1027.(70) Wieme, J.; Vanduyfhuys, L.; Rogge, S. M. J.; Waroquier, M.; vanSpeybroeck, V. Exploring the Flexibility of MIL-47(V)-Type MaterialsUsing Force Field Molecular Dynamics Simulations. J. Phys. Chem. C2016, 120, 14934−14947.(71) Addicoat, M. A.; Vankova, N.; Akter, I. F.; Heine, T. Extensionof the Universal Force Field to Metal−Organic Frameworks. J. Chem.Theory Comput. 2014, 10, 880−891.(72) Fogarty, A. C.; Coudert, F.-X.; Boutin, A.; Laage, D.Reorientational Dynamics of Water Confined in Zeolites. Chem-PhysChem 2014, 15, 521−529.(73) Fincham, D. Choice of timestep in molecular dynamicssimulation. Comput. Phys. Commun. 1986, 40, 263−269.(74) Hunenberger, P. H. Advanced Computer Simulation; SpringerScience + Business Media: 2005; pp 105−149.(75) Rogge, S.; Vanduyfhuys, L.; Ghysels, A.; Waroquier, M.;Verstraelen, T.; Maurin, G.; Speybroeck, V. V. A Comparison ofBarostats for the Mechanical Characterization of Metal−OrganicFrameworks. J. Chem. Theory Comput. 2015, 11, 5583−5597.(76) Nose, S. A unified formulation of the constant temperaturemolecular dynamics methods. J. Chem. Phys. 1984, 81, 511.(77) Hoover, W. G. Canonical dynamics: Equilibrium phase-spacedistributions. Phys. Rev. A: At., Mol., Opt. Phys. 1985, 31, 1695−1697.(78) Martyna, G. J.; Tobias, D. J.; Klein, M. L. Constant pressuremolecular dynamics algorithms. J. Chem. Phys. 1994, 101, 4177.(79) Dubbeldam, D.; Torres-Knoop, A.; Walton, K. S. On the innerworkings of Monte Carlo codes. Mol. Simul. 2013, 39, 1253−1292.(80) HASTINGS, W. K. Monte Carlo sampling methods usingMarkov chains and their applications. Biometrika 1970, 57, 97−109.(81) Bouessel du Bourg, L. B.; Ortiz, A. U.; Boutin, A.; Coudert, F.-X.Thermal and mechanical stability of zeolitic imidazolate frameworkspolymorphs. APL Mater. 2014, 2, 124110.(82) Plimpton, S. Fast Parallel Algorithms for Short-Range MolecularDynamics. J. Comput. Phys. 1995, 117, 1−19.(83) Abraham, M. J.; Murtola, T.; Schulz, R.; Pall, S.; Smith, J. C.;Hess, B.; Lindahl, E. GROMACS: High performance molecular

simulations through multi-level parallelism from laptops to super-computers. SoftwareX 2015, 1−2, 19−25.(84) Smith, W.; Todorov, I. T. A short description of DL_POLY.Mol. Simul. 2006, 32, 935−943.(85) Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q. RASPA:molecular simulation software for adsorption and diffusion in flexiblenanoporous materials. Mol. Simul. 2016, 42, 81−101.(86) Design, M. MedeA GIBBS: Thermodynamics of fluids andadsorption. http://www.materialsdesign.com/medea/gibbs, 2016.(87) Triguero, C.; Coudert, F.-X.; Boutin, A.; Fuchs, A. H.; Neimark,A. V. Mechanism of Breathing Transitions in Metal−OrganicFrameworks. J. Phys. Chem. Lett. 2011, 2, 2033−2037.(88) Coudert, F.-X.; Mellot-Draznieks, C.; Fuchs, A. H.; Boutin, A.Prediction of Breathing and Gate-Opening Transitions Upon BinaryMixture Adsorption in Metal-Organic Frameworks. J. Am. Chem. Soc.2009, 131, 11329−11331.(89) Coudert, F.-X.; Jeffroy, M.; Fuchs, A. H.; Boutin, A.; Mellot-Draznieks, C. Thermodynamics of Guest-Induced Structural Tran-sitions in Hybrid Organic-Inorganic Frameworks. J. Am. Chem. Soc.2008, 130, 14294−14302.(90) Eslami, H.; Muller-Plathe, F. Molecular dynamics simulation inthe grand canonical ensemble. J. Comput. Chem. 2007, 28, 1763−1773.(91) Kuznetsova, T.; Kvamme, B. Grand canonical moleculardynamics for TIP4P water systems. Mol. Phys. 1999, 97, 423−431.(92) Gupta, A.; Chempath, S.; Sanborn, M. J.; Clark, L. A.; Snurr, R.Q. Object-oriented Programming Paradigms for Molecular Modeling.Mol. Simul. 2003, 29, 29−46.(93) Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N·log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993,98, 10089.(94) Fukuda, I.; Yonezawa, Y.; Nakamura, H. Consistent MolecularDynamics Scheme Applying the Wolf Summation for CalculatingElectrostatic Interaction of Particles. J. Phys. Soc. Jpn. 2008, 77, 114301.(95) Soergel, D. A. W. Rampant software errors may underminescientific results. F1000Research 2015, 303.(96) Wilson, G.; Aruliah, D. A.; Brown, C. T.; Hong, N. P. C.; Davis,M.; Guy, R. T.; Haddock, S. H. D.; Huff, K. D.; Mitchell, I. M.;Plumbley, M. D.; Waugh, B.; White, E. P.; Wilson, P. Practices forScientific Computing. PLoS Biol. 2014, 12, e1001745.(97) Torvalds, L.; Hamano, J. Git. http://www.git-scm.com, 2016.(98) Preston-Wernerm, T.; Wanstrath, C.; Hyett, P. Github. http://www.github.com, 2016.(99) Fuchs, S. Travis CI. http://www.travis-ci.com, 2016.(100) Bae, Y.-S.; Snurr, R. Q. Development and Evaluation of PorousMaterials for Carbon Dioxide Separation and Capture. Angew. Chem.,Int. Ed. 2011, 50, 11586−11596.(101) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer,C. E.; Sarjeant, A. A.; Snurr, R. Q.; Nguyen, S. T.; Ozgur Yazaydın, A.;Hupp, J. T. Metal−Organic Framework Materials with UltrahighSurface Areas: Is the Sky the Limit? J. Am. Chem. Soc. 2012, 134,15016−15021.(102) Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of Gases inMultimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309−319.(103) Rouquerol, J.; Llewellyn, P.; Rouquerol, F. Stud. Surf. Sci.Catal.; Elsevier BV: 2007; pp 49−56.(104) Wang, T. C.; Bury, W.; Gomez-Gualdron, D. A.; Vermeulen,N. A.; Mondloch, J. E.; Deria, P.; Zhang, K.; Moghadam, P. Z.;Sarjeant, A. A.; Snurr, R. Q.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K.Ultrahigh Surface Area Zirconium MOFs and Insights into theApplicability of the BET Theory. J. Am. Chem. Soc. 2015, 137, 3585−3591.(105) Sarkisov, L.; Kim, J. Computational structure characterizationtools for the era of material informatics. Chem. Eng. Sci. 2015, 121,322−330.(106) Sarkisov, L.; Harrison, A. Computational structure character-isation tools in application to ordered and disordered porous materials.Mol. Simul. 2011, 37, 1248−1257.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

211

Page 14: Computational Chemistry Methods for Nanoporous Materials · 2020-05-29 · Computational Chemistry Methods for Nanoporous Materials Jack D. Evans,† Guillaume Fraux,† Romain Gaillac,†,‡

(107) First, E. L.; Gounaris, C. E.; Wei, J.; Floudas, C. A.Computational characterization of zeolite porous networks: anautomated approach. Phys. Chem. Chem. Phys. 2011, 13, 17339.(108) Willems, T. F.; Rycroft, C. H.; Kazi, M.; Meza, J. C.;Haranczyk, M. Microporous Mesoporous Mater. 2012, 149, 134−141.(109) Levesque, M.; Duvail, M.; Pagonabarraga, I.; Frenkel, D.;Rotenberg, B. Accounting for adsorption and desorption in latticeBoltzmann simulations. Phys. Rev. E 2013, 88, 013308.(110) Ladd, A. J. C.; Verberg, R. Lattice-Boltzmann Simulations ofParticle-Fluid Suspensions. J. Stat. Phys. 2001, 104, 1191−1251.(111) Frisch, U.; d’Humieres, D.; Hasslacher, B.; Lallemand, P.;Pomeau, Y.; Rivet, J.-P. Lattice gas hydrodynamics in two and threedimensions. Complex Systems 1987, 1, 649−707.(112) d’Humieres, D.; Lallemand, P.; Frisch, U. Lattice gas model for3D hydrodynamics. Europhys. Lett. 1986, 2, 291−297.(113) Frisch, U.; Hasslacher, B.; Pomeau, Y. Lattice-Gas Automatafor the Navier-Stokes equation. Phys. Rev. Lett. 1986, 56, 1505−1509.(114) Mei, R.; Shyy, W.; Yu, D.; Luo, L.-S. Lattice BoltzmannMethod for 3-D Flows with Curved Boundary. J. Comput. Phys. 2000,161, 680−699.(115) Succi, S. The Lattice-Boltzmann equation for fluid dynamics andbeyond; Oxford Science Publications: 2001.(116) Lowe, C.; Frenkel, D. The super long-time decay of velocityfluctuations in a two-dimensional fluid. Phys. A 1995, 220, 251−260.(117) Lowe, C. P.; Frenkel, D. Do Hydrodynamic DispersionCoefficients Exist? Phys. Rev. Lett. 1996, 77, 4552−4555.(118) Vanson, J.-M.; Coudert, F.-X.; Rotenberg, B.; Levesque, M.;Tardivat, C.; Klotz, M.; Boutin, A. Unexpected coupling between flowand adsorption in porous media. Soft Matter 2015, 11, 6125−6133.

Chemistry of Materials Review

DOI: 10.1021/acs.chemmater.6b02994Chem. Mater. 2017, 29, 199−212

212