assessment of dft methods for transition metals with the...

13
Assessment of DFT Methods for Transition Metals with the TMC151 Compilation of Data Sets and Comparison with Accuracies for Main- Group Chemistry Bun Chan,* ,Peter M. W. Gill, and Masanari Kimura Graduate School of Engineering, Nagasaki University, Bunkyo 1-14, Nagasaki 852-8521, Japan Research School of Chemistry, Australian National University, Canberra, Australian Capital Territory 2601, Australia * S Supporting Information ABSTRACT: In the present study, we have gathered a collection (that we term TMC151) of accurate reference data for transition-metal reactions for the assessment of quantum chemistry methods. It comprises diatomic dissociation energies and reaction energies and barriers for prototypical transition-metal reactions. Our assessment of a diverse range of dierent types of DFT methods shows that the most accurate functionals include ωB97M- V, ωB97X-V, MN15, and B97M-rV. Notably, they have also been previously validated to be highly robust for main-group chemistry. Nevertheless, even these methods show substantially worse accuracies for transition metals than for main-group chemistry. For less accurate methods, there is not a good correlation between their accuracies for main-group and transition-metal chemistries. Thus, in the development of new DFT, it is important to assess the accuracies for both types of data. In this regard, we have formulated the TMC34 model for ecient assessment of the performance for transition metals, which complements our previously developed MG8 model for main-group chemistry. Together, they provide a cost-eective means for initial assessment of new methodologies. INTRODUCTION Density functional theory (DFT) is an indispensable tool in modern day chemical research. With decades of development, it is now applicable to a wide range of systems and properties. 1 For instance, the emergence of hybrid DFT methods such as B3- LYP 2 has provided a useful tool for estimating thermochemical properties. The later development of kinetic/general-purpose functionals such as M06-2X 3 further enables reliable computa- tion of reaction barriers without signicantly compromising the accuracy for thermochemistry. The deciency of DFT in treating noncovalent interactions has also been largely rectied with the introduction of dispersion corrections. 4,5 In addition, the use of range-separated DFT leads to a viable means for calculating excited state properties (e.g., with LC-DFT) 6 and extended systems (with screened-exchange functionals), 7 in particular when combined with advanced techniques such as the maximum overlap method. 8 Until the recent past, the development of a functional that is equally applicable to both main-group and transition-metal chemistry remained one of the key problems facing DFT. However, with the advent of functionals (e.g., MN15 9 ) that are in part designed for transition-metal systems, this challenge has begun to be overcome. Nonetheless, most DFT methods, including many recently devised ones, are intended for the modeling of main-group chemistry. These include ωB97M-V 10 that has been formulated recently by trainingwith an extensive set of main-group molecular properties. Although ωB97M-V is tted with main-group chemistry, it has been shown in a subsequent assessment 11 that its accuracy for a transition-metal reaction is close to that achieved by MN15. This remarkable observation raises the question of whether a DFT trained with only main-group species can be as accurate for transition-metal chemistry when compared with those that perform equally well for main group but also include transition metals in their training sets. In recent years, we have also been interested in the development of new functionals, 1215 in which we generally use main-group thermochemical data 1620 for training and assessment purposes. In our other studies, we have applied established DFT methods to the study of transition-metal catalysis. 2125 As a result, the question of how accurate a main- group-trained method can be when used for transition metal is of interest to us and we believe also to a wide range of developers and users of DFT. The primary aim of the present study is to provide additional benchmarks toward answering this question. In addition, our results also yield further insights and tools for future development of DFT methods. COMPUTATIONAL DETAILS Standard computational chemistry calculations were carried out with Q-Chem 5.0. 26 Geometries and benchmark relative energies were obtained from previous studies. Notably, the ones in the MGCDB82 set 27 were taken from ref 11. Those for the TMD60 set were taken from ref 28. The geometries in ref 29 were used for the reactions in the MOR13 model, which is a statistical representation of the MOR41 test set. 29 The TMB50 set employs structures given in refs 3033. The default Received: March 8, 2019 Published: May 31, 2019 Article pubs.acs.org/JCTC Cite This: J. Chem. Theory Comput. 2019, 15, 3610-3622 © 2019 American Chemical Society 3610 DOI: 10.1021/acs.jctc.9b00239 J. Chem. Theory Comput. 2019, 15, 36103622 Downloaded by AUSTRALIAN NATL UNIV at 14:43:09:136 on June 11, 2019 from https://pubs.acs.org/doi/10.1021/acs.jctc.9b00239.

Upload: others

Post on 08-Apr-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

Assessment of DFT Methods for Transition Metals with the TMC151Compilation of Data Sets and Comparison with Accuracies for Main-Group ChemistryBun Chan,*,† Peter M. W. Gill,‡ and Masanari Kimura†

†Graduate School of Engineering, Nagasaki University, Bunkyo 1-14, Nagasaki 852-8521, Japan‡Research School of Chemistry, Australian National University, Canberra, Australian Capital Territory 2601, Australia

*S Supporting Information

ABSTRACT: In the present study, we have gathered a collection (that we term TMC151)of accurate reference data for transition-metal reactions for the assessment of quantumchemistry methods. It comprises diatomic dissociation energies and reaction energies andbarriers for prototypical transition-metal reactions. Our assessment of a diverse range ofdifferent types of DFT methods shows that the most accurate functionals include ωB97M-V, ωB97X-V, MN15, and B97M-rV. Notably, they have also been previously validated to behighly robust for main-group chemistry. Nevertheless, even these methods showsubstantially worse accuracies for transition metals than for main-group chemistry. For less accurate methods, there is not agood correlation between their accuracies for main-group and transition-metal chemistries. Thus, in the development of newDFT, it is important to assess the accuracies for both types of data. In this regard, we have formulated the TMC34 model forefficient assessment of the performance for transition metals, which complements our previously developed MG8 model formain-group chemistry. Together, they provide a cost-effective means for initial assessment of new methodologies.

■ INTRODUCTION

Density functional theory (DFT) is an indispensable tool inmodern day chemical research. With decades of development, itis now applicable to a wide range of systems and properties.1 Forinstance, the emergence of hybrid DFT methods such as B3-LYP2 has provided a useful tool for estimating thermochemicalproperties. The later development of kinetic/general-purposefunctionals such as M06-2X3 further enables reliable computa-tion of reaction barriers without significantly compromising theaccuracy for thermochemistry. The deficiency of DFT intreating noncovalent interactions has also been largely rectifiedwith the introduction of dispersion corrections.4,5 In addition,the use of range-separated DFT leads to a viable means forcalculating excited state properties (e.g., with LC-DFT)6 andextended systems (with screened-exchange functionals),7 inparticular when combined with advanced techniques such as themaximum overlap method.8

Until the recent past, the development of a functional that isequally applicable to both main-group and transition-metalchemistry remained one of the key problems facing DFT.However, with the advent of functionals (e.g., MN159) that arein part designed for transition-metal systems, this challenge hasbegun to be overcome. Nonetheless, most DFT methods,including many recently devised ones, are intended for themodeling of main-group chemistry. These include ωB97M-V10

that has been formulated recently by “training”with an extensiveset of main-group molecular properties. Although ωB97M-V isfitted with main-group chemistry, it has been shown in asubsequent assessment11 that its accuracy for a transition-metalreaction is close to that achieved by MN15. This remarkable

observation raises the question of whether a DFT trained withonly main-group species can be as accurate for transition-metalchemistry when compared with those that perform equally wellfor main group but also include transitionmetals in their trainingsets.In recent years, we have also been interested in the

development of new functionals,12−15 in which we generallyuse main-group thermochemical data16−20 for training andassessment purposes. In our other studies, we have appliedestablished DFT methods to the study of transition-metalcatalysis.21−25 As a result, the question of how accurate a main-group-trainedmethod can bewhen used for transitionmetal is ofinterest to us and we believe also to a wide range of developersand users of DFT. The primary aim of the present study is toprovide additional benchmarks toward answering this question.In addition, our results also yield further insights and tools forfuture development of DFT methods.

■ COMPUTATIONAL DETAILS

Standard computational chemistry calculations were carried outwith Q-Chem 5.0.26 Geometries and benchmark relativeenergies were obtained from previous studies. Notably, theones in the MGCDB82 set27 were taken from ref 11. Those forthe TMD60 set were taken from ref 28. The geometries in ref 29were used for the reactions in the MOR13 model, which is astatistical representation of the MOR41 test set.29 The TMB50set employs structures given in refs 30−33. The default

Received: March 8, 2019Published: May 31, 2019

Article

pubs.acs.org/JCTCCite This: J. Chem. Theory Comput. 2019, 15, 3610−3622

© 2019 American Chemical Society 3610 DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

Dow

nloa

ded

by A

UST

RA

LIA

N N

AT

L U

NIV

at 1

4:43

:09:

136

on J

une

11, 2

019

from

http

s://p

ubs.

acs.

org/

doi/1

0.10

21/a

cs.jc

tc.9

b002

39.

Page 2: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

quadrature grids of Q-Chem are used for all DFT computations.Importantly, this corresponds to the largest “SG-3” grid forMinnesota functionals that have been shown to have such arequirement.34

Unless otherwise noted, we used the def2-TZVP basis set35

and, where appropriate, associated effective core potentials inour single-point energy calculations. It has been shown that, for aset of 20 transition-metal species, the basis-set effect beyondtriple-ζ is small at the Hartree−Fock level.36 We note that,however, some DFT methods show larger than expected basisset dependences.37 We have briefly examined this aspect andfind that, while the use of a triple-ζ basis set may not yieldquantitatively converged results, the remaining basis-set effectsare small and would not lead to different qualitative conclusions(Supporting Information). Statistical analyses were carried outwith JMP Pro 14.38,39 Energies in the text are vibrationless valuesgiven in kJ mol−1. Deviations are defined as calculated valueminus reference value.

■ RESULTS AND DISCUSSIONDFT Methods Assessed. In the present study, we intend to

evaluate the performance of a wide range of DFT methods fortransition-metal chemistry. While one may choose a selection ofmethods based on subjective prior knowledge, we attempt to besomewhat more objective in our choice. In a recent extensivestudy,11 ∼200 DFT methods have been assessed for theiraccuracies for calculating main-group molecular properties, andwe will use a subset of these methods for our assessment.We will not consider local-density-approximation (LDA)

functionals and methods with Hartree−Fock as the onlyexchange component as candidates. This is because they arerarely an appropriate choice for calculatingmolecular properties.At the other end of the spectrum, we will not consider double-hybrid (DH) DFT methods because these methods are (in ouropinion) closer in spirit to wave function methods than to DFT.For example, the various DSD-DFT methods of Kozuch andMartin40 generally contain ∼70% HF and ∼50% MP2. Inaddition, because of the substantial advantage gained from theinclusion of MP2, when DH-DFT methods are optimally fitted,their accuracies show little dependence on the underlying DFTmethod. For instance, even an LDA-based DSD-DFT performscomparably to those based on GGA and meta-GGA methods.40

We will select the DFT methods from the rest using a fewdifferent approaches.In ref 11, the ∼200 methods are grouped into the following

categories. They are divided into nonhybrid functionals andglobal-hybrid (GH) and range-separated (RS)methods depend-ing on if and how Hartree−Fock exchange is incorporated. Eachof these three groups is subdivided into general gradientapproximation (GGA) and meta-GGA (mGGA) methodsaccording to whether a kinetic energy density component isincluded (mGGA) or not (GGA). A further division for each ofthe six types into non-dispersion-corrected and dispersion-corrected methods leads to 12 categories. Because we intend tohave a set of methods that span the range of performance for theMGCDB82 set of main-group thermochemical properties, fromeach of these subsets of methods, we include the functionals withthe lowest and highest overall mean absolute deviations(MADs) in our assessment.The resulting 24 DFTmethods selected in this way are shown

in Table 1. Importantly, among this collection are ωB97M-Vand MN15. The former is one of the most accurate functionalsformulated with just main-group properties, whereas the latter is

a highly accurate method intended for both main-group andtransition-metal chemistry. It is noteworthy that, for somecategories, the performances of the functionals with the smallestand largest MADs are in fact not very different. For example, alldispersion-corrected global-hybrid GGA and range-separatedmeta-GGA methods in Table 1 perform well, with MADs below10 kJ mol−1.We also include additional functionals chosen with an

alternative strategy. In a recent study,27 we have used standardstatistical techniques to formulate a model for accuratelyrepresenting the MGCDB82 set with just eight out of its 82subsets. In other words, a small set of chemical data can bechosen to serve as a proxy for a much wider range of data. In asimilar manner, we can use the same methodology to build amodel that includes just a small number of functionals butrepresents a large variety of methods. Our intention is to providean independent set of functionals and that this set can be utilizedwith low computational cost.To select these DFT methods, we obtain the average MADs

for the ∼200 methods for each of the 82 subsets of MGCDB82,as well as an estimated MAD (EMAD) by expressing the actualMADs as linear combinations of the MADs for a variablenumber of functionals (plus a constant term). The quality of theform of EMAD is judged by the correlation of the estimatedvalues for the 82 subsets with the actual MADs, as indicated bythe coefficient of determination (R2) for a linear fit. To arrive atan accurate EMADmodel with aminimal number of functionals,

Table 1. 24 DFT Methods Selected from Ref 11 for 12 Typesof Functionals and the Corresponding MAD (kJ mol−1) forthe MGCDB82 Data Set

method hybrid formfunctional

typedispersioncorrection MAD

GAM nonhybrid GGA not included 10.5B-OP nonhybrid GGA not included 22.6B97-D3BJ nonhybrid GGA included 8.6PBEsol-D3 nonhybrid GGA included 21.1mBEEF nonhybrid meta-GGA not included 7.9PKZB nonhybrid meta-GGA not included 19.2B97M-rV nonhybrid meta-GGA included 4.9MS0-D3 nonhybrid meta-GGA included 10.1B97-1 global-hybrid GGA not included 8.6revPBE0 global-hybrid GGA not included 15.1B97-3-D2 global-hybrid GGA included 5.1revPBE0-D30 global-hybrid GGA included 7.9MN15 global-hybrid meta-GGA not included 5.3TPSSh global-hybrid meta-GGA not included 11.8PW6-B95-D2 global-hybrid meta-GGA included 4.9M06-HF-D3 global-hybrid meta-GGA included 10.0ωB97X range-

separatedGGA not included 6.2

LRC-ωPBE range-separated

GGA not included 11.5

ωB97X-V range-separated

GGA included 4.5

N12-SX-D3BJ range-separated

GGA included 8.1

M11 range-separated

meta-GGA not included 6.4

MN12-SX range-separated

meta-GGA not included 8.6

ωB97M-V range-separated

meta-GGA included 3.3

MN12-SX-D3BJ

range-separated

meta-GGA included 7.4

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3611

Page 3: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

we exploit the forward selection technique while maintaining acriterion of R2 > 0.995. The resulting model contains just sixfunctionals out of the ∼200 candidates, and we term it the DF6model and designate the associated EMAD as EMADDF6, whichis defined by the formula

= × + ×

+ × + ×

+ × + ×

−ω

‐ ‐−

EMAD 0.2024 MAD 0.3013 MAD

0.0524 MAD 0.0649 MAD

0.1087 MAD 0.2875 MAD

0.1044 (kJ mol )

DF6 B LYP SCAN

PBEsol M06 L

PW B6K B97x rV1

There is no overlap between the six methods in DF6 and the24 functionals in Table 1, and thus the two sets togetherrepresent 30 unique DFT methods. However, neither of thesetwo compendia contain several widely used ones that we deemto be necessary inclusions for DFT benchmark. They are PBE,41

B3-LYP,2 PBE0,42 and M06-2X,3 and we additionally incorpo-rate them into our collection of methods. Furthermore, we notethat the M06 method3 is often applied to transition-metalchemistry; we therefore also include it in the present study. As aresult, we have examined a total of 35 functionals. The fullcitations of these methods are given in the SupportingInformation. We also refer readers to ref 11 for a recentdiscussion on the various types of DFT methods.Benchmark Test Sets. Transition metals have attracted the

attention of theoretical and computational chemists for anumber of reasons. From a theoretical perspective, theyrepresent a significant challenge to typical (single-reference)quantum chemistry methods because many transition-metalspecies contain considerable multireference character43−45 (seealso the Supporting Information for additional discussions).From a practical perspective, transition metals are indispensablefor modern-day chemistry due to their remarkable catalyticactivities,46 and the reliable prediction of their reactivities hasbeen a notable objective for computational chemists.47 In thepresent study, we aim to assess the chosen DFT methods fromboth theoretical and practical perspectives. Some of the systemsused in our assessment are given in Figure 1.In a recent study,28 a systematic and diverse set of bond

energies for 60 diatomic transition-metal species has beencompiled for the assessment of DFT methods (Figure 1a).

While the accurate calculations of some of these bond energieshave been shown to be reasonably straightforward, others aremore challenging. Among the 60 systems, 20 have been studiedusing high-level coupled-cluster methods up to CCSDTQ, andthe results demonstrate considerable multireference characterfor some species, as indicated by large post-CCSD(T)contributions to bond energies.36 This latter collection of 20systems has been referred to in ref 36 as the 3dMLBE20 set. Inthe present study, we will instead designate these transition-metal dimer dissociation energies as the TMD20 set for the sakeof simplicity.We will use the full set of 60 systems to assess the accuracy of

the selected DFT methods for fundamental thermochemistry.For the 20 species for which experimental values have beencritically assessed with high-level [composite post-CCSD(T) atthe complete-basis-set (CBS) limit] bond energies, we will usethe recommended values as our reference benchmark energies.For some of the remaining compounds, two sets of experimentalbond energies are given in ref 28. In these cases, we adopt theones that are believed to be more accurate; they are shown in theSupporting Information of ref 28 as the red-colored values intheir Table S1. We will designate the complete 60-data-point setas TMD60.For the assessment of practically more relevant properties, we

use two sets of data of transition-metal complexes. In a recentstudy, a set of 41 metal−organic reactions was presented withhigh-level local-CCSD(T)/CBS reference energies.29 We havesubsequently used the aforementioned statistical strategy(forward selection) to formulate a model with a subset of 13reactions (Figure 1b) to represent the complete set.27 A linearcombination of the absolute deviations for the 13 reactions isused to obtain an estimated MAD that corresponds to the 41reactions. This model, that we termed MOR13, and theassociated estimated MAD (EMADMOR13) will be employed inthe present study. The value of EMADMOR13 is defined accordingto the formula in which ΔE (kJ mol−1) is the deviation inreaction energies relative to reference values:

= × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + −

EMAD 0.1455 E 0.0839 E

0.0273 E 0.0380 E 0.0358 E

0.0152 E 0.0605 E 0.0956 E

0.0694 E 0.0647 E 0.0604 E

0.1210 E 0.2146 E 1.0116 (kJ mol )

MOR13 1 2

3 4 5

6 7 8

9 10 11

12 131

Importantly, the EMADMOR13 values correlate with the actualMADs for MOR41 with an R2 of 0.995 for a set of 41 DFTmethods.27 This model is thus a computationally efficient andaccurate alternative to conducting an assessment using the fullMOR41 set. In order to further validate MOR13 as arepresentation for MOR41, in the present study, we haveobtained the actualMADs for several DFTmethods examined inthe present study (Supporting Information). The results showthat EMADMOR13 provides a good estimate of MADMOR41 ineach of these cases, with deviations of ∼1 kJ mol−1.Another set of data that we will use is a compilation of reaction

barriers involving complexes of second- and third-row transitionmetals, which were obtained by Chen and co-workers in a seriesof investigations.30−33 A total of 50 reaction barriers are includedin this collection, for which the reference values were obtainedpreviously with CCSD(T)/CBS or local-CCSD(T)/CBSprotocols. They cover the elements Zr, Mo, W, Re, Ir, Pt, andAu, and reaction types insertion, C−H activation, hydration,

Figure 1. Systems contained in (a) the TMD60 set and (b) theMOR13model. For MOR13, ligands in the transition-metal complexes that donot participate directly are omitted for clarity. The abbreviations forchemical groups and species are “Me” for methyl, “Ac” for acetyl”, “Cy”for cyclohexyl, “dmpe” for “1,2-bis(dimethylphosphino)ethane”, and“Ph” for phenyl.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3612

Page 4: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

oxidation, reduction, metathesis, epoxidation, isomerization,and Friedel−Crafts reaction (Figure 2). For detailed descrip-

tions of these reactions, we refer readers to the original literature,specifically Scheme 1 in each of the refs 30−33. We willdesignate these 50 transition-metal barriers as the TMB50 set.In total, TMD60, MOR13, and TMB50 comprise 123 data

points, from which a total MAD can be obtained for a DFTmethod as an overall gauge of its general accuracy for thecomputation of thermochemical properties of transition-metalspecies. As mentioned above, the data points in MOR13 areintended for obtaining EMADMOR13 by a linear-combinationmodel as an estimation to the MAD for a 41-data-point set. Wewill therefore employ EMADMOR13 instead of treating thereaction energies in MOR13 as simple individual data points.With the same reasoning, we will scale EMADMOR13 accordinglywhen obtaining an overall MAD for the three collections of data.As a result, we define the overall MAD for the combinedcompilation as follows:

Figure 2. Metals and reaction types covered by the TMB50 set.

Table 2. Average Magnitudes (kJ mol−1) of the Reference Data and Mean Absolute Deviations for the MGCDB82 Set of Main-Group Thermochemical Data and the TMC151 Set and Its Subsets TMD20, TMD60, MOR13, and TMB50 for Transition-MetalChemistry

method MGCDB82 TMC151 TMD20 TMD60 MOR13a TMB50

av magnitude 187.4 196.7 302.0 365.0 130.9 48.8selected for method type and MAD for MGCDB82 (Table 1)

GAM 10.5 29.4 21.1 27.4 50.0 14.8B-OP 22.6 39.5 32.0 37.1 66.7 20.1B97-D3BJ 8.6 21.3 15.8 23.6 22.1 17.7PBEsol-D3 21.1 39.6 60.3 65.1 22.7 22.8mBEEF 7.9 24.1 29.6 34.0 23.1 13.0PKZB 19.2 31.0 33.6 37.9 42.1 13.5B97M-rV 4.9 19.8 29.1 32.0 12.5 11.1MS0-D3 10.1 21.6 24.5 32.2 15.9 13.6B97-1 8.6 21.9 15.5 25.2 32.9 8.8revPBE0 15.1 30.1 26.5 39.1 42.1 9.5B97-3-D2 5.1 20.3 17.6 32.6 18.0 7.4revPBE0-D3 7.9 21.5 26.5 39.0 12.3 8.0MN15 5.3 15.4 22.3 26.8 9.2 6.9TPSSh 11.8 23.4 30.3 31.0 28.9 9.8PW6-B95-D2 4.9 20.7 19.3 29.0 23.1 8.7M06-HF-D3 10.0 42.0 60.2 76.3 21.7 17.5ωB97X 6.2 22.7 26.3 32.2 23.5 10.7LRC-ωPBE 11.5 22.2 27.9 30.5 24.4 10.4ωB97X-V 4.5 15.3 20.4 25.7 8.9 8.0N12-SX-D3BJ 8.1 18.7 27.6 29.2 15.2 8.9M11 6.4 22.2 40.8 37.8 14.6 9.7MN12-SX 8.6 25.5 45.7 47.0 16.5 7.1ωB97M-V 3.3 16.3 21.7 30.6 8.3 5.8MN12-SX-D3BJ 7.4 26.5 45.7 46.8 20.3 7.3

statistically chosen for the DF6 modelB-LYP 18.0 35.6 31.6 36.0 54.7 19.4SCAN 9.1 22.5 30.4 35.7 13.1 14.5PBEsol 21.1 40.6 60.3 65.1 28.9 20.9M06-L 8.5 22.6 21.3 28.9 23.8 14.1PW-B6K 9.1 26.7 26.0 46.3 19.4 9.2ωB97X-rV 4.5 15.3 20.3 25.8 8.9 8.0

additional widely used functionalsPBE 14.3 31.7 38.8 45.0 32.1 15.3B3-LYP 18.0 28.1 20.5 29.6 45.9 11.8PBE0 8.8 22.1 24.6 30.7 26.1 8.5M06 6.0 25.0 25.1 39.2 24.4 8.5M06-2X 5.4 31.1 29.5 52.4 26.7 9.1

aEMAD values for MOR41 obtained by the linear-combination model given in ref 27.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3613

Page 5: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

= [ × + ×

+ × ]

MAD 60 MAD 41 EMAD

50 MAD /151TMC151 TMD60 MOR13

TMB50

Correspondingly, we will refer to the complete set of transition-metal chemistry data used in the present study as TMC151.Background Discussion on the Accuracy for

MGCDB82. In the coming sections, we will assess the 35DFT methods with the three sets of transition-metalthermochemistry data. Before we do so, let us reiterate that amotivation for the present study is based on a previousobservation11 that suggests a highly accurate DFT methoddesigned for main-group chemistry (ωB97M-V) also showspromising initial results for a transition-metal system. It is thusappropriate to first provide an overview of the accuracies of the35 functionals for main-group chemistry as a baseline forcomparison.We will use the MADs for the MGCDB82 set and its eight

categories of subsets as the benchmark for main-groupthermochemical properties. The values for MGCDB82 areshown in Table 2, alongside the results for the various transition-metal data sets. The MADs for the various categories of datapoints within MGCDB82 are given in the SupportingInformation. They contain noncovalent interactions (categoriesNCED, NCEC, and NCD), isomerization energies (IE and ID),thermochemistries (TCE and TCD), and barrier heights (BH).For the complete MGCDB82 set, ωB97M-V10 has the smallestMAD of 3.3 kJ mol−1, which is followed by closely relatedωB97X-V48 and ωB97X-rV,49 with a value of 4.5 kJ mol−1,coincidentally, for both of these two latter methods. A number ofother functionals also show good accuracies for the MGCDB82set. Specifically, five other methods (B97M-rV,49 B97-3-D2,50,51

PW6-B95-D2,51,52 M06-2X,3 and MN159) have MAD valuesthat are less than 6 kJ mol−1.Notably, B97M-rV shows a remarkably small MAD of just 4.9

kJ mol−1 for a nonhybrid functional; it is competitive with someof the most accurate hybrid-DFT methods, and it achieves thiswith a lower cost by completely excluding the computation ofHartree−Fock exchange. We also note that MN15 (MAD = 5.3kJ mol−1), which is formulated for both main-group andtransition-metal chemistries, is also among the top performers. Itis also noteworthy that the group of methods with the smallestoverall MADs also shows good performances across the boardfor the eight categories (Supporting Information). For example,ωB97M-V ranks first in six of the eight categories and second forthe remaining two. Similarly, MN15 is among the top five in fourcategories, but even in the case (IE) where it has the lowestranking among its eight rankings, theMAD is just 2.3 kJ mol−1 incomparison with 0.8 kJ mol−1 for ωB97M-V.At the other end of the scale, several methods (B-OP,53

PBEsol,54 and PBEsol-D354,55) show large MADs of over 20 kJmol−1. While these are all nonhybrid functionals, we reiteratethat, as demonstrated by the good accuracy of B97M-rV, theinclusion of Hartree−Fock may not be critical unless the highestlevel of accuracy is an absolute necessity in a main-groupthermochemistry application.Dissociation Energies for Diatomic Transition-Metal

Species. We begin our discussion on the accuracies of thevarious DFT methods for transition-metal chemistry byexamining the results for the TMD60 set (Table 2). In additionto the MADs for the entire set, we have also obtained thecorresponding ones for the TMD20 subset for which highlyaccurate reference values are available. By and large, the trends in

the MADs between the two sets of data follow one another, withan R2 of 0.80 and a standard error of 5.5 kJ mol−1 for the MADvalues for TMD60 over the range of 23.6−76.3 kJ mol−1.The reasonable correlation with the TMD20 subset is

important because some of the reference data in TMD60carry no or large quoted uncertainties,28 and the faircorrespondence provides a degree of reassurance for the useof the full TMD60 set for benchmark purposes. A notabledifference between the 20- and 60-point data sets is that theMADs for TMD60 are generally larger, sometimes by asignificant margin. Large differences in the MAD values aremore often associated with large MADs. Conversely, forfunctionals with small MADs for TMD20, the correspondingMADs for TMD60 are usually not significantly larger. Thissuggests that the most accurate methods examined here show acertain robustness across the different systems, though theaccuracies themselves are not outstanding, as we shall discuss.Let us now examine the accuracies of the DFT methods for

these transition-metal diatomics dissociation energies. Animmediate observation is that the MAD values are much largerthan those for MGCDB82. This is the case even for the mostaccurate methods found for main-group chemistry. For example,ωB97M-V hasMADs of 21.7 and 30.6 kJ mol−1, respectively, forTMD20 and TMD60, in comparison with 3.3 kJ mol−1 forMGCDB82. Nonetheless, such an accuracy is in fact similar tothat forMN15 (22.3 and 26.8 kJmol−1, respectively, for TMD20and TMD60), which has been designed for both main-groupand transition-metal chemistries.In addition to the comparison with the overall MADs for

MGCDB82, it is instructive to compare the accuracies forTMD60 with the difficult thermochemical data withinMGCDB82. In this regard, the MADs of ωB97M-V andMN15 for the TCD (thermochemistry, difficult) category are14.2 and 21.0 kJ mol−1. This puts the TMD60 set of dissociationenergies at a similar level of difficulty to some of the mostchallenging main-group systems. In passing, we also note thatB97M-rV also shows reasonable accuracy for TMD20 andTMD60 in comparison with ωB97M-V and MN15, withcorresponding MADs of 29.1 and 32.0 kJ mol−1. Functionalswith the smallest MADs for these two sets of systems are severalB97-type methods (e.g., B97-D3BJ,56,57 B97-1,58ωB97X-V, andωB97X-rV). Their values for TMD20 are∼15−20 kJ mol−1, andthe ones for TMD60 are ∼25 kJ mol−1.At the other end of the scale, methods that show the largest

deviations are PBEsol, PBEsol-D3, and M06-HF-D3,59 withMAD values that are larger than 60 kJ mol−1 for both TMD20and TMD60 sets. We note that a common practice for thetreatment of transition metals is to use nonhybrid functionals orhybrid methods with a small proportion of Hartree−Fockexchange. In this regard, the poor performance of M06-HF-D3may appear to be supportive of this preference. However, thenegative impact on the accuracy for transition-metal systems byincluding a large amount of Hartree−Fock exchange60 appearsto be smaller than commonly believed, as demonstrated by therelatively good performances achieved by functionals with largeproportions of Hartree−Fock exchange, such as MN15 (44%)and the various ωB97-type methods (15−100%).At this point, we note that many of the methods examined

here that yield the smallest MADs for TMD20 and TMD60(e.g., B97-type methods such as ωB97X-V) were devised withonly main-group data. Their decent accuracies for TMD20 andTMD60 may be a sign of characteristics in transition metals

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3614

Page 6: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

being partially and implicitly captured by the protocol forformulating these “main-group” functionals.Reaction Energies and Barriers in Catalytic Systems.

We now turn our attention to the systems that are of practicalrelevance to transition-metal catalysis. In general, the EMADsfor the MOR13 model of reaction energies and the MADs forthe TMB50 set of barriers are smaller than those for thedissociation energies (Table 2). To illustrate this quantitatively,the averages of the MADs/EMADs for TMD20, TMD60,MOR13, and TMC50 are 30.0, 37.2, 25.1, and 11.7 kJ mol−1,respectively. Of course, if one were to take the magnitudes of thereference data into account, it can also be argued that thedissociation energies represent a smaller percentage deviation(10%) than those for reaction energies (19%) and barriers(24%).A comparison of the trends between the three types of data

shows that TMD60, MOR13, and TMC50 do not correspondparticularly well with one another. The R2 values for the linearcorrelations for TMD60/MOR13, MOR13/TMB50, andTMD60/TMB50 are 0.002, 0.26, and 0.20, respectively. Despitethe generally poor correlations, methods that have small MADvalues for TMD60 also show fair to good accuracies moregenerally. The smallest MADTMD60, EMADMOR13, andMADTMB50 values are 23.6 (B97-D3BJ), 8.3 (ωB97M-V), and5.8 (ωB97M-V) kJ mol−1, respectively. In comparison, forωB97X-V the corresponding values are 25.7, 8.9, and 8.0 kJmol−1, and for MN15 they are 26.8, 9.2, and 6.9 kJ mol−1. Tofurther put these numbers into perspective, we can comparethem with the average MAD/EMAD values mentioned above[37.2 (TMD60), 25.1 (MOR13), and 11.7 (TMC50)], as wellas the largest MADTMD60, EMADMOR13, and MADTMB50 of 76.3(M06-HF-D3), 66.7 (B-OP), and 22.8 (PBEsol-D3) kJ mol−1,respectively.Overall Performance andComparisonwithMGCDB82.

When the three sets of transition-metal chemistry data arecombined to give the TMC151 set, we arrive at a single metricfor assessing the DFT methods within the scope of the presentstudy. As discussed above, a DFT that shows decent accuracy forone of the three subsets of TMC151 is also generally quite goodfor the other two sets. As a result, some of the best performingmethods for the complete transition-metal test set remain thefairly accurate ones that we have mentioned above for thesubsets (Table 2).Specifically, the six functionals with the smallest MADTMC151

values are, in ascending order of their MADs, ωB97X-V48 andωB97X-rV49 (15.3 kJ mol−1), MN159 (15.4), ωB97M-V10

(16.3), N12-SX-D3BJ61 (18.7), and B97M-rV49 (19.8). Thegeneral accuracies of the first four methods are, in our opinion,indistinguishable for TMC151. As noted in our discussion of theprevious results for MGCDB82, their accuracies for main-groupsystems are also quite similar [MAD = 4.5 (ωB97X-V andωB97X-rV), 5.3 (MN15), and 3.3 (ωB97M-V) kJmol−1]. Thesefunctionals appear to be the most dependable choice for thecomputation of molecular thermochemical quantities coveringboth main-group and transition-metal elements.The next two methods (N12-SX-D3BJ and B97M-rV) have

MADTMC151 values of ∼19 kJ mol−1, which are larger than ∼16kJ mol−1 achieved by the four most accurate methods but notsignificantly so. Between N12-SX-D3BJ and B97M-rV, the latterhas a better accuracy for MGCDB82 (MAD = 4.9 kJ mol−1

versus 8.1 for N12-SX-D3BJ). In principle, it is also computa-tionally less demanding for it is a nonhybrid functional, whereasN12-SX-D3BJ is a screened-exchange DFT. We therefore deem

B97M-rV to be an attractive option when the use of a global-hybrid functional (MN15) or a long-range-corrected DFT(ωB97X-V, ωB97X-rV, and ωB97M-V) is computationallyimpractical for a particular application to molecular systems. Wealso note that functionals that are least accurate for MGCDB82(e.g., PBEsol, MAD = 21.1 kJ mol−1) generally do not fare wellfor the transition-metal systems examined in the present study(MAD = 40.6 kJ mol−1 for PBEsol for TMC151).Given that the most/least accurate batch of methods for

MGCDB82 is also found to be the most/least accurate forTMC151, it is of interest to inspect whether the former set ofmain-group chemical properties is more generally a goodindicator for the latter compilation for transition-metalchemistry. We use the MADs for the 35 DFT methods for theTMC151 and MGCDB82 sets and their subsets/subcategoriesas a probe. Table 3 shows the R2 values when drawing

correlations between the transition-metal data sets and themain-group sets. It is apparent that, in an overall sense, there is amoderate correlation between the two types of quantities, withan R2 value of 0.62 between TMC151 and MGCDB82. Withthat being said, for functionals that do not belong to the mostand least accurate extremes, there is considerable uncertainty indeducing the performance for one type of data by looking atanother type (Supporting Information). Interestingly, the TCDsubset of MGCDB82 appears to be a better indicator for theperformance for TMC151, with an R2 value of 0.76.We now examine the correlations between the subsets of

TMC151 and those of MGCDB82. We can see that the TCD-type of main-group data has the largest R2 values for thecorrelations with TMD20 and TMD60, but these R2 values of∼0.5 are still poor. In comparison, the EMADs of MOR13correlate reasonably with the MADs for the NCED subset ofnoncovalent interactions (R2 = 0.74). This is somewhat curiousbecause one might expect that the metal−organic reactions inMOR13, which can be considered as changes in covalent bonds,would correlate better with the main-group thermochemistrydata sets (TCE andTCD). The stronger correlation withNCEDis associated with the large sizes of the metal complexes, whichlead to large additional noncovalent interactions upon bindingof substrates, as pointed out in ref 29. Lastly, we find a good R2

Table 3. Correlations (R2) between theMADs/EMADs of the35 DFT Methods for the TMC151 Set and Its Subsets andThose for the MGCDB82 Set and Its Subcategories of DataPointsa

TMC151 TMD20 TMD60 MOR13 TMB50

MGCDB82 0.62 0.22 0.15 0.54 0.57NCED 0.28 0.01 0.00 0.74 0.15NCEC 0.36 0.11 0.06 0.45 0.24NCD 0.49 0.32 0.19 0.22 0.69IE 0.40 0.05 0.02 0.71 0.29ID 0.37 0.02 0.01 0.61 0.56TCE 0.49 0.38 0.31 0.12 0.56TCD 0.76 0.53 0.49 0.22 0.69BH 0.47 0.22 0.13 0.24 0.87

aNCED = noncovalent interaction, easy, dimer; NCEC = noncovalentinteraction, easy, cluster; NCD = noncovalent interaction, difficult; IE= isomerization energy, easy; ID = isomerization energy, difficult;TCE = thermochemistry, easy; TCD = thermochemistry, difficult; andBH = barrier height.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3615

Page 7: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

value of 0.87 for the correlation between TMB50 and the main-group BH subset.Taking all the above results into consideration, we propose

that a small MAD for the MGCDB82 data set may indicaterobustness for both main-group and transition-metal chem-istries. We have previously suggestedMADMGCDB82≲ 4 kJ mol−1

to be a general guideline,27 and we believe that this wouldcontinue to be relevant given the seemingly broadened scopebeyond main-group chemistry.We have previously formulated a model (termed MG8 with

64 data points) for estimating MADMGCDB82, which may be usedfor rapid evaluation of the accuracies of new quantum chemistrymethods.27 As an aside but in a similar vein, if we obtain theEMADs with the DF6model for all 123 data points in TMC151,they correlate reasonably with the MADs calculated using thedeviations for the other 29models, with an R2 of 0.91. This resultserves as a validation for the use of the DF6 model as a probe forthe difficulty of a particular problem for an “average” functional.Investigation into Some Systems with Large Devia-

tions.We now proceed to examine the MADs (for the 29 DFTmethods) for the various systems in the TMC151 set.Within theTMD60 subset of BDEs, the MADs vary from a modest 12.9 kJmol−1 (for ZnS) to a very substantial 112.0 kJ mol−1 (for ScCl)(Table 4). Overall, two species have MADs that are larger than90 kJ mol−1 [ZnF (94.5) and ScCl (112.0)], whereas for otherspecies the MADs are∼60 kJ mol−1 or smaller. We will examinethe cases of ZnF and ScCl in greater detail.For ScCl, it has been suggested that the reference

experimental value (BDE of 341.1 kJ mol−1) is too low,28 asindicated by an alternative high-level theoretical value of 447.7kJ mol−1.62 We note that several of the best performing DFTs,for example, ωB97M-V and MN15, show large positivedeviations for ScCl (+99.6 and +107.3 kJ mol−1, respectively).Most of the DFT methods examined also yield BDEs that aresignificantly larger than the experimental value; the meandeviation (MD) for the 29 DFT methods is +111.2 kJ mol−1,

which is virtually identical in magnitude to the MAD. Theagreements between the theoretical reference BDE and the DFTvalues are quite reasonable, with a mild MAD of 12.1 kJ mol−1.The story for the BDE of ZnF is similar to that for the ScCl

BDE. It is characterized by a large MD of −94.5 kJ mol−1 that isvirtually identical in magnitude when compared with the MADand by a moderate standard deviation (SD) of the deviations(29.2 kJ mol−1, cf. 17.1 kJ mol−1 for ScCl). A high-leveltheoretical study has reported a value (292.3 kJ mol−1) that isvery different from the experimental BDE of 381.9 kJ mol−1.63

The DFT values agree considerably better with the theoreticalreference value, with an MAD of 20.8 kJ mol−1. For both ScCland ZnF, it seems desirable to re-evaluate the experimentalBDEs.We also note that systems such as ScBr, FeH, NiF, and NiBr

are also associated with large MADs and large magnitudes forthe MDs (>50 kJ mol−1) but relatively modest SDs (<30 kJmol−1). We believe that re-examination of these BDEs may alsobe desirable. When these six systems are excluded from ouranalysis, the average of the MADs for the TMD60 set reducesfrom 37.2 kJ mol−1 to 30.1 kJ mol−1. The correlation betweenthe TMD20 and TMD60 sets also improves, with an R2 value of0.86 (versus 0.80 when all 60 data points are used).The SD values for the oxides (and to a somewhat lesser extent,

the sulfides) are quite large. Indeed, the largest SD for theTMD60 set can be seen for MnO (74.2 kJ mol−1). In this case,high-level multireference computations support the experimen-tal reference value.64 The deviations for ωB97M-V and MN15are −42.3 and +8.2 kJ mol−1, which is large in the former casebut, arguably, not excessively larger than its MAD of 30.6 kJmol−1 for the TMD60 set (Table 2). TheMD value for MnO forthe set of DFT methods is just +10.6 kJ mol−1, which isconsiderably smaller in magnitude than the MAD of 54.0 kJmol−1. Thus, the nonsystematic component for the deviations ismuch larger than the systematic component. Taking all of thesestatistical measures into account, it seems that the BDEs of

Table 4. Mean Absolute Deviations (kJ mol−1) over 29 DFT Methods for the Individual Systems in the TMD60 Set and theCorresponding Mean Deviations and Standard Deviations (SD) of the Deviationsa

Sc Ti V Cr Mn Fe Co Ni Cu Zn

mean absolute deviationH 22.9 36.0 47.3 19.1 32.3 58.5 40.5 17.5 15.0 15.2F 20.1 31.6 52.3 44.2 36.2 28.6 33.9 57.3 29.1 94.5Cl 112.0 32.5 30.3 19.2 28.5 21.8 36.1 16.9 19.9 17.5Br 58.4 25.1 61.5 21.7 36.0 40.1 33.9 50.8 20.3 30.3O 39.0 47.3 50.3 40.8 54.0 45.1 54.7 47.8 29.5 33.5S 27.4 50.9 38.4 27.1 36.4 29.7 28.0 48.2 17.4 12.9

mean deviationH 22.6 25.0 40.3 15.5 15.8 58.3 38.5 8.0 −3.4 8.3F −2.3 −5.5 −44.2 −43.2 −18.5 −15.4 −31.4 −57.3 −25.8 −94.5Cl 112.0 23.3 16.5 5.0 4.9 11.8 33.3 −1.2 −12.2 −6.6Br −58.4 4.8 −60.0 8.7 −27.9 37.3 −31.6 −50.8 −11.6 28.3O −15.8 −13.4 −3.3 −21.2 10.6 −10.0 −39.7 −20.6 −22.0 −27.3S −7.8 41.4 −1.4 −10.5 18.3 −6.9 −11.4 −38.3 −2.5 −1.0

SD of the deviationsH 15.2 34.8 36.8 21.7 36.2 24.7 19.4 23.1 22.8 19.9F 26.6 42.0 43.2 25.5 41.0 32.2 25.0 25.5 22.8 29.2Cl 17.1 32.1 34.3 26.8 33.6 27.4 24.6 22.3 21.3 22.9Br 15.7 30.6 32.6 30.4 34.8 33.7 23.4 21.2 20.5 27.9O 49.1 62.4 71.8 44.5 74.2 58.1 60.8 61.7 33.1 28.3S 36.5 48.6 52.0 32.3 47.6 37.8 38.5 49.8 24.6 16.3

aMethods included in Table 2, excluding those for the DF6 model.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3616

Page 8: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

diatomic transition-metal oxides such as MnO represent achallenging case for DFT methods, including some of the bestperforming ones.TheMADs for the individual reaction energies in theMOR13

model are generally quite modest, with values of ∼10−30 kJmol−1. However, two reactions (Figure 3) are associated with

large MADs of 55.3 and 42.4 kJ mol−1. It is noteworthy thatthese two reactions are also associated with large deviations forthe methods assessed in ref 29. An examination of the individualdeviations that contribute to these MADs shows that, in bothcases, the non-dispersion-corrected methods generally performmuch poorer than the dispersion-corrected ones. In the case ofthe association reaction (a), the MAD for non-dispersion-corrected methods is very substantial (81.9 kJ mol−1), which isin stark contrast to the modest value of 17.5 kJ mol−1 for thedispersion-corrected functionals. We see a similar but less

dramatic difference between the two types of methods for theligand substitution reaction (b).These observations again demonstrate the importance of

adequate treatment of dispersion interactions for organometallicspecies, in which large ligands and substrates that can lead tosubstantial dispersion interactions are often part of the system.With that being said, we note that some methods that do notinclude external dispersion corrections, such as MN15, performquite well for these two reactions [deviations of −9.7 (a) and+11.2 (b) kJ mol−1]. In comparison, the deviations for reactions(a) and (b) for ωB97M-V are −22.2 and +6.6 kJ mol−1,respectively.For the TMB50 set of barriers, the MADs for most of the

reactions are considerably below 20 kJ mol−1. Six reactions, onthe other hand, have larger MADs that are above this threshold(Figure 4). Four of these are (de)oxygenation reactions (9, 16,17, and 24), whereas two are pericyclic-type reactions (15 and45). We note that, for the pericyclic-type reactions, thedeviations for ωB97M-V and MN15 are modest. Thecorresponding deviations for the de(oxygenation) reactionsare more substantial, particularly forωB97M-V despite it havinga smallMAD of 5.8 kJmol−1 for the complete TMB50 set (Table4).We have inspected the deviations for other functionals for

these four reactions. By comparing the results for PBE [with noHartree−Fock exchange (HFX)] and PBE0 (25% HFX), itappears that these barriers are fairly sensitive to the proportionof HFX in the functional. This would account for the largedeviations forωB97M-V, which contains 15%HFX in the short-range, but it increases rapidly to over 40% at an interelectronicdistance of ∼1 Å and over 90% at ∼4 Å. The MN15 method,

Figure 3. Reactions in the MOR13 model that generally show largedeviations from benchmark reaction energies for a set of 29 DFTmethods assessed and MADs for all 29 methods and those for non-dispersion-corrected and dispersion-corrected subsets of functionals.

Figure 4. Reactions in the TMB50 set that generally show large deviations from benchmark reaction barriers for a set of 29 DFTmethods assessed andMADs and deviations (kJ mol−1) for selected functionals.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3617

Page 9: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

which incorporates a constant 44% HFX, has somewhat smallerdeviations.Overall, the observations on the systems with large deviations

carry three messages: (1) while transition metals representchallenges to modern DFT methods, the source of deviationsfrom experimental values may well be the experimental valuesthemselves; (2) adequate account of (midrange) dispersions,either by explicit inclusion of dispersion corrections or byimplicit parametrization, is indispensable; and (3) Hartree−Fock exchange represents a useful inclusion in modernfunctionals, but achieving a proportion that is optimal in allcases is still elusive. These messages are by no means news, butour results, including those for some of the currently mostaccurate and robust methods, show that these understandingsdeveloped in the past are still relevant. It is worth remindingourselves of them when formulating new methods and applyingexisting ones.Small Data-Set Models for Rapid Assessment. The

MG8model mentioned previously provides a convenient meansfor assessing the accuracy of a computational chemistry methodfor main-group thermochemistry. As noted above, the resultsdiscussed so far suggest that, when developing a new method,

the achievement of an excellent EMADMG8 could be a usefulinitial indicator for robustness for not only main-group but alsotransition-metal chemistry. However, such a target may beelusive during the exploration stage of development. It wouldtherefore be useful to additionally have a rapid assessmentprotocol for transition-metal chemistry, which may provideadditional physical insights. To enable such a task, we herein usethe same statistical methodology (forward selection) that weused to formulate MG8, MOR13, and DF6 to devise small-data-set models for TMD60 and TMB50.For each set, we build a minimal subset of data points such

that, for the 35 functionals, the EMADs obtained as a linearcombination of the absolute deviations correlate with the actualMADs with R2 > 0.995. For TMD60, this gives a group of 10constituent deviations for dissociation energies (ΔD), namelythose for TiBr, VO, CrS, MnBr, FeCl, FeS, CoH, CoF, CuBr,and ZnS. We term this model TMD10, and the correspondingEMADTMD10 is defined by

Figure 5. Reactions contained in the TMB11 model of barrier heights.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3618

Page 10: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

= × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + −

EMAD 0.0372 D 0.1410 D

0.0670 D 0.0434 D 0.1447 D

0.0457 D 0.0629 D 0.1857 D

0.0552 D 0.0833 D 9.7526 (kJ mol )

TMD10 TiBr VO

CrS MnBr FeCl

FeS CoH CoF

CuBr ZnS1

For the TMB50 set, a collection of 11 reaction barriers ischosen statistically, and we designate the model as TMB11accordingly. The associated EMADTMB11 can be obtained by thefollowing formula using deviations for the barriers (ΔB), withthe included reactions shown in Figure 5.

= × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + × |Δ |

+ × |Δ | + × |Δ | + −

EMAD 0.0878 B 0.0450 B 0.0693 B

0.0577 B 0.1444 B 0.0220 B

0.0674 B 0.1223 B 0.1631 B

0.0650 B 0.0911 B 0.9963 (kJ mol )

TMB11 1 7 9

11 21 22

23 33 35

38 481

As a whole, the combination of the TMD10, MOR13, andTMB11 models employs 34 data points to represent TMC151.We term this collection of 34 energies and the associated linear-combination equations the TMC34 model. We deem it to be acost-effective means for assessing the robustness of a quantumchemistry method when it is used for cross-validating thefindings from the results for the MG8 model of main-groupchemistry data.Prospects for Expansion. At this final stage of our

discussion, let us reiterate that a question that we seek toanswer is whether the accuracy of a functional for main-groupproperties would reflect its accuracy for transition-metalchemistry. The preliminary answer is in part positive. When amethod is shown to be highly robust for an extensive and diverseset of main-group data, the robustness may also extend totransition metals. However, such a connection may be weakerfor less accurate methods.To further test the validity of this proposition, assessments

with an expanded collection of data sets and a wider range oftransition-metal properties would be desirable. One of us has inthe past experimentally examined a wide range of transition-metal-catalyzed reactions, and we have begun to computation-ally elucidate themechanistic details for some of these processes.They include, for example, the use of relatively inertnucleophiles such as alkenes, alkynes, and conjugated dienesfor regio- and stereoselective CO2 fixation.

65−67 In due course,the results for the studies on some of these catalytic reactionsmay provide suitable benchmark data for assessing quantumchemistry methods.It is also noteworthy that, in addition to homogeneous

catalysis, heterogeneous catalysis is also a major use of transitionmetals. Highly accurate data that can serve as suitable referencefor benchmark purpose is, to the best of our knowledge, notavailable at present. Nonetheless, CCSD(T) reaction energiesand barriers for some symbolic reactions of small metal clustershave been obtained in recent years.68,69 Further works along thisline of research would be desirable. In relation to the chemistryof metal particles, we have in the past obtained benchmark-quality atomization energies of small gold clusters.70 A largercollection of highly accurate fundamental properties of thesespecies would also be valuable for the evaluation of quantumchemistry methods. We will expand the present study uponemergence of suitable additional data in the future.

■ CONCLUDING REMARKSIn the present study, we have gathered a collection of accuratereference data for transition-metal reactions for the assessmentof DFT methods. This compilation, termed TMC151,comprises the TMD60 set of 60 dissociation energies fordiatomic species of first-row transition metals, the MOR13model for reaction energies of common catalytic processes, andthe TMB50 set of 50 barriers for prototypical transition-metalreactions. With MOR13 being a statistically devised model thatrepresents 41 reactions in a previously formulated MOR41 set,TMC151 accounts for 151 relative energies for transition-metal-containing compounds.We have used this compendium to assess a diverse range of

different types of DFT methods. The most accurate functionalsfound for TMC151 include ωB97M-V, ωB97X-V, MN15, andB97M-rV. While MN15 has been optimized for both main-group and transition-metal species, ωB97M-V, ωB97X-V, andB97M-rV are devised with only main-group data. Our resultssuggest that a functional that is highly accurate and robust formain-group data may also be among the most accurate fortransition-metal chemistry. Nonetheless, even the most accuratefunctionals show worse accuracies for transition metals than formain-group chemistry. We are therefore still far from the end ofpursuit toward universally applicable DFT.We find that, for the most accurate methods examined, the

accuracies for main-group chemistry appear to be a reasonableindicator for the accuracies for transition metals. However, sucha correlation is less apparent for functionals that perform lesswell. In the development of new DFT, it is thus important toassess the accuracies for both types of data along the way ofachieving an outstanding accuracy. We have previously usedobjective statistical techniques to devise the MG8 model (withjust 64 data points) for reliable and rapid evaluation of theaccuracy for MGCDB82 (with ∼4400 data points). Similarly, inthe present study, we have formulated the TMC34 model as arepresentation of TMC151 for efficient assessment of theperformance for transition metals. We deem the combination ofMG8 and TMC34 to be a cost-effective means for initialdevelopment of new quantum chemistry methods.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.jctc.9b00239.

Additional discussions and references (PDF)Geometries for the species in the data sets of TMD10,MOR13, and TMB1 (ZIP)Templates for obtaining EMADs with these models(XLSX)

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] Chan: 0000-0002-0082-5497Peter M. W. Gill: 0000-0003-1042-6331Masanari Kimura: 0000-0003-2900-9554FundingWe gratefully acknowledge research funding from the JapanSociety for the Promotion of Science (JSPS) (Program forAdvancing Strategic International Networks to Accelerate the

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3619

Page 11: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

Circulation of Talented Researchers) and generous grants ofcomputer time from the RIKEN Advanced Center forComputing and Communication (ACCC) (Project Q19266),Japan, and National Computational Infrastructure (NCI),Australia.NotesThe authors declare no competing financial interest.

■ REFERENCES(1) Sholl, D.; Steckel, J. A. Density Functional Theory: A PracticalIntroduction; Wiley: Hoboken, 2009; DOI: 10.1002/9780470447710.(2) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. AbInitio Calculation of Vibrational Absorption and Circular DichroismSpectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98,11623−11627.(3) Zhao, Y.; Truhlar, D. G. TheM06 Suite of Density Functionals forMain Group Thermochemistry, Thermochemical Kinetics, Non-covalent Interactions, Excited States, and Transition Elements: TwoNew Functionals and Systematic Testing of Four M06-Class Func-tionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215−241.(4) Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C.Dispersion-Corrected Mean-Field Electronic Structure Methods.Chem. Rev. 2016, 116, 5105−5154.(5) Bursch, M.; Caldeweyher, E.; Hansen, A.; Neugebauer, H.; Ehlert,S.; Grimme, S. Understanding and Quantifying London DispersionEffects in Organometallic Complexes. Acc. Chem. Res. 2019, 52, 258−266.(6) Tsuneda, T.; Hirao, K. Long-Range Correction for DensityFunctional Theory. WIREs Comput. Mol. Sci. 2014, 4, 375−390.(7) Henderson, T. M.; Paier, J.; Scuseria, G. E. Accurate Treatment ofSolids with the HSE Screened Hybrid. Phys. Status Solidi B 2011, 248,767−774.(8) Gilbert, A. T. B.; Besley, N. A.; Gill, P. M. W. Self-Consistent FieldCalculations of Excited States Using the Maximum Overlap Method(MOM). J. Phys. Chem. A 2008, 112, 13164−13171.(9) Yu, H. S.; He, X.; Li, S. L.; Truhlar, D. G. MN15: A Kohn-ShamGlobal-Hybrid Exchange-Correlation Density Functional with BroadAccuracy for Multi-Reference and Single-Reference Systems andNoncovalent Interactions Chem. Sci. 2016, 7, 5032−5051.(10) Mardirossian, N.; Head-Gordon, M. ωB97M-V: A Combinato-rially Optimized, Range-Separated Hybrid, Meta-GGA Density Func-tional with VV10 Nonlocal Correlation. J. Chem. Phys. 2016, 144,214110.(11) Mardirossian, N.; Head-Gordon, M. Thirty Years of DensityFunctional Theory in Computational Chemistry: An Overview andExtensive Assessment of 200 Density Functionals.Mol. Phys. 2017, 115,2315−2372.(12) Chan, B.; Song, J. W.; Kawashima, Y.; Hirao, K. Toward theComplete Range Separation of Non-Hybrid Exchange−CorrelationFunctional. J. Comput. Chem. 2015, 36, 871−877.(13) Chan, B.; Song, J. W.; Kawashima, Y.; Hirao, K. Performance oftheOPCorrelation Functional in Relation to its Formulation: Influenceof the Exchange Component and the Effect of Incorporating Same-SpinCorrelations. J. Comput. Chem. 2016, 37, 1306−1312.(14) Chan, B.; Kawashima, Y.; Hirao, K. Correlation Functional inScreened-Exchange Density Functional Theory Procedures. J. Comput.Chem. 2017, 38, 2307−2315.(15) Chan, B.; Kawashima, Y.; Hirao, K. The reHISS Three-RangeExchange Functional with an Optimal Variation of Hartree−Fock andIts Use in the reHISSB-D Density Functional Theory Method. J.Comput. Chem. 2019, 40, 29.(16) Chan, B.; Deng, J.; Radom, L. G4(MP2)-6X: A Cost-EffectiveImprovement to G4(MP2). J. Chem. Theory Comput. 2011, 7, 112−120.(17) Chan, B.; Radom, L. ObtainingGood Performance with Triple-ζ-Type Basis Sets in Double-Hybrid Density Functional TheoryProcedures. J. Chem. Theory Comput. 2011, 7, 2852−2863.

(18) Chan, B.; Radom, L. W1X-1 and W1X-2: W1-Quality Accuracywith an Order of Magnitude Reduction in Computational Cost. J.Chem. Theory Comput. 2012, 8, 4259−4269.(19) Chan, B.; Gill, P. M. W.; Radom, L. Performance of Gradient-Corrected and Hybrid Density Functional Theory: Role of theUnderlying Local Density Approximation and the Gradient Correction.J. Chem. Theory Comput. 2012, 8, 4899−4906.(20) Chan, B.; Karton, A.; Raghavachari, K.; Radom, L. Restricted-Open-Shell G4(MP2)-Type Procedures. J. Phys. Chem. A 2016, 120,9299−9304.(21) Yap, A. J.; Chan, B.; Yuen, A. K. L.; Ward, A. J.; Masters, A. F.;Maschmeyer, T. A Palladium-Catalyzed Multicascade Reaction: FacileLow-Temperature Hydrogenolysis of Activated Nitriles and RelatedFunctional Groups. ChemCatChem 2011, 3, 1496−1502.(22) Chan, B.; Ball, G. E. A Benchmark Ab Initio and DFT Study ofthe Structure and Binding of Methane in the σ−Alkane ComplexCpRe(CO)2(CH4). J. Chem. Theory Comput. 2013, 9, 2199−2208.(23) McGuinness, D. S.; Chan, B.; Britovsek, G. J. P.; Yates, B. F.Ethylene Trimerisation with Cr-PNP Catalysts: A TheoreticalBenchmarking Study and Assessment of Catalyst Oxidation State.Aust. J. Chem. 2014, 67, 1481−1490.(24) McSkimming, A.; Chan, B.; Bhadbhade, M. M.; Ball, G. E.;Colbran, S. B. Bio-Inspired Transition Metal-Organic HydrideConjugates for Catalysis of Transfer Hydrogenation: Experiment andTheory. Chem. - Eur. J. 2015, 21, 2821−2834.(25) Chan, B.; Luo, Y.; Kimura,M.Mechanism for Three-ComponentNi-Catalyzed Carbonyl−Ene Reaction for CO2 Transformation: WhatPractical Lessons Do We Learn from DFT Modelling? Aust. J. Chem.2018, 71, 272−278.(26) Shao, Y.; Gan, Z.; Epifanovsky, E.; Gilbert, A. T. B.; Wormit, M.;Kussmann, J.; Lange, A. W.; Behn, A.; Deng, J.; Feng, X.; Ghosh, D.;Goldey, M.; Horn, P. R.; Jacobson, L. D.; Kaliman, I.; Khaliullin, R. Z.;Kus, T.; Landau, A.; Liu, J.; Proynov, E. I.; Rhee, Y. M.; Richard, R. M.;Rohrdanz, M. A.; Steele, R. P.; Sundstrom, E. J.; Woodcock, H. L., III;Zimmerman, P. M.; Zuev, D.; Albrecht, B.; Alguire, E.; Austin, B.;Beran, G. J. O.; Bernard, Y. A.; Berquist, E.; Brandhorst, K.; Bravaya, K.B.; Brown, S. T.; Casanova, D.; Chang, C.-M.; Chen, Y.; Chien, S. H.;Closser, K. D.; Crittenden, D. L.; Diedenhofen, M.; DiStasio, R. A., Jr.;Dop, H.; Dutoi, A. D.; Edgar, R. G.; Fatehi, S.; Fusti-Molnar, L.;Ghysels, A.; Golubeva-Zadorozhnaya, A.; Gomes, J.; Hanson-Heine,M.W. D.; Harbach, P. H. P.; Hauser, A. W.; Hohenstein, E. G.; Holden, Z.C.; Jagau, T.-C.; Ji, H.; Kaduk, B.; Khistyaev, K.; Kim, J.; Kim, J.; King,R. A.; Klunzinger, P.; Kosenkov, D.; Kowalczyk, T.; Krauter, C.M.; Lao,K. U.; Laurent, A.; Lawler, K. V.; Levchenko, S. V.; Lin, C. Y.; Liu, F.;Livshits, E.; Lochan, R. C.; Luenser, A.; Manohar, P.; Manzer, S. F.;Mao, S.-P.; Mardirossian, N.; Marenich, A. V.; Maurer, S. A.; Mayhall,N. J.; Oana, C. M.; Olivares-Amaya, R.; O’Neill, D. P.; Parkhill, J. A.;Perrine, T. M.; Peverati, R.; Pieniazek, P. A.; Prociuk, A.; Rehn, D. R.;Rosta, E.; Russ, N. J.; Sergueev, N.; Sharada, S. M.; Sharmaa, S.; Small,D. W.; Sodt, A.; Stein, T.; Stuck, D.; Su, Y.-C.; Thom, A. J. W.;Tsuchimochi, T.; Vogt, L.; Vydrov, O.; Wang, T.; Watson, M. A.;Wenzel, J.; White, A.; Williams, C. F.; Vanovschi, V.; Yeganeh, S.; Yost,S. R.; You, Z.-Q.; Zhang, I. Y.; Zhang, X.; Zhou, Y.; Brooks, B. R.; Chan,G. K. L.; Chipman, D. M.; Cramer, C. J.; Goddard, W. A., III; Gordon,M. S.; Hehre, W. J.; Klamt, A.; Schaefer, H. F., III; Schmidt, M. W.;Sherrill, C. D.; Truhlar, D. G.; Warshel, A.; Xua, X.; Aspuru-Guzik, A.;Baer, R.; Bell, A. T.; Besley, N. A.; Chai, J.-D.; Dreuw, A.; Dunietz, B. D.;Furlani, T. R.; Gwaltney, S. R.; Hsu, C.-P.; Jung, Y.; Kong, J.;Lambrecht, D. S.; Liang, W.; Ochsenfeld, C.; Rassolov, V. A.;Slipchenko, L. V.; Subotnik, J. E.; Van Voorhis, T.; Herbert, J. M.;Krylov, A. I.; Gill, P. M. W.; Head-Gordon, M. Advances in MolecularQuantum Chemistry Contained in the Q-Chem 4 Program Package.Mol. Phys. 2015, 113, 184−215.(27) Chan, B. Formulation of Small Test Sets Using Large Test Setsfor Efficient Assessment of Quantum Chemistry Methods. J. Chem.Theory Comput. 2018, 14, 4254−4262.(28) Moltved, K. A.; Kepp, K. P. Chemical Bond Energies of 3dTransition Metals Studied by Density Functional Theory. J. Chem.Theory Comput. 2018, 14, 3479−3492.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3620

Page 12: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

(29) Dohm, S.; Hansen, A.; Steinmetz, M.; Grimme, S.; Checinski, M.P. Comprehensive Thermochemical Benchmark Set of RealisticClosed-Shell Metal Organic Reactions. J. Chem. Theory Comput.2018, 14, 2596−2608.(30) Kang, R.; Lai, W.; Yao, J.; Shaik, S.; Chen, H. How Accurate Cana Local Coupled Cluster Approach Be in Computing the ActivationEnergies of Late-Transition-Metal-Catalyzed Reactions with Au, Pt,and Ir? J. Chem. Theory Comput. 2012, 8, 3119−3127.(31) Sun, Y.; Chen, H. Performance of Density Functionals forActivation Energies of Zr-Mediated Reactions. J. Chem. Theory Comput.2013, 9, 4735−4743.(32) Sun, Y.; Chen, H. Performance of Density Functionals forActivation Energies of Re-CatalyzedOrganic Reactions. J. Chem. TheoryComput. 2014, 10, 579−588.(33) Hu, L.; Chen, H. Assessment of DFT Methods for ComputingActivation Energies of Mo/W-Mediated Reactions. J. Chem. TheoryComput. 2015, 11, 4601−4614.(34) Dasgupta, S.; Herbert, J. M. Standard Grids for High-PrecisionIntegration of Modern Density Functionals: SG-2 and SG-3. J. Comput.Chem. 2017, 38, 869−882.(35) Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence,Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn:Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7,3297−3305.(36) Cheng, L.; Gauss, J.; Ruscic, B.; Armentrout, P. B.; Stanton, J. F.Bond Dissociation Energies for Diatomic Molecules Containing 3dTransition Metals: Benchmark Scalar-Relativistic Coupled-ClusterCalculations for 20 Molecules. J. Chem. Theory Comput. 2017, 13,1044−1056.(37) Mardirossian, N.; Head-Gordon, M. Characterizing andUnderstanding the Remarkably Slow Basis Set Convergence of SeveralMinnesota Density Functionals for Intermolecular InteractionEnergies. J. Chem. Theory Comput. 2013, 9, 4453−4461.(38) Jones, B.; Sall, J. JMP Statistical Discovery Software. WIREsComp. Stat. 2011, 3, 188−194.(39) JMP, Version 14; SAS Institute Inc.: Cary, NC, 2018.(40) Kozuch, S.; Martin, J. M. L. Spin-Component-Scaled DoubleHybrids: An Extensive Search for the Best Fifth-Rung FunctionalsBlending DFT and Perturbation Theory. J. Comput. Chem. 2013, 34,2327−2344.(41) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized GradientApproximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.(42) Adamo, C.; Barone, V. Toward Reliable Density FunctionalMethods Without Adjustable Parameters: The PBE0Model. J. Chem.Phys. 1999, 110, 6158−6170.(43) Chan, B.; Karton, A.; Raghavachari, K.; Radom, L. Heats ofFormation for CrO, CrO2, and CrO3: An Extreme Challenge for Black-Box Composite Procedures. J. Chem. Theory Comput. 2012, 8, 3159−3166.(44) Jiang, W.; DeYonker, N. J.; Wilson, A. K. MultireferenceCharacter for 3d Transition-Metal-Containing Molecules. J. Chem.Theory Comput. 2012, 8, 460−468.(45) Wang, J.; Manivasagam, S.; Wilson, A. K. MultireferenceCharacter for 4d Transition-Metal-Containing Molecules. J. Chem.Theory Comput. 2015, 11, 5865−5872.(46) Steinborn, D. Fundamentals of Organometallic Catalysis; Wiley-VCH: Weinheim, 2012; DOI: 10.1002/anie.201201768.(47) Sameera, W. M. C.; Maseras, F. Transition Metal Catalysis byDensity Functional Theory and Density Functional Theory/MolecularMechanics. WIREs Comput. Mol. Sci. 2012, 2, 375−385.(48) Mardirossian, N.; Head-Gordon, M. ωB97X-V: a 10-Parameter,Range-Separated Hybrid, Generalized Gradient Approximation Den-sity Functional With Nonlocal Correlation, Designed by a Survival-of-the-Fittest Strategy. Phys. Chem. Chem. Phys. 2014, 16, 9904−9924.(49) Mardirossian, N.; Pestana, L. R.; Womack, J. C.; Skylaris, C.-K.;Head-Gordon, T.; Head-Gordon, M. Use of the rVV10 NonlocalCorrelation Functional in the B97M-V Density Functional: DefiningB97M-rV and Related Functionals. J. Phys. Chem. Lett. 2017, 8, 35−40.

(50) Keal, T. W.; Tozer, D. J. Semiempirical Hybrid Functional withImproved Performance in an Extensive Chemical Assessment. J. Chem.Phys. 2005, 123, 121103.(51) Grimme, S. Semiempirical GGA-Type Density FunctionalConstructed with a Long-Range Dispersion Correction. J. Comput.Chem. 2006, 27, 1787−1799.(52) Zhao, Y.; Truhlar, D. G. Design of Density Functionals that areBroadly Accurate for Thermochemistry, Thermochemical Kinetics, andNonbonded Interactions. J. Phys. Chem. A 2005, 109, 5656−5667.(53) Tsuneda, T.; Suzumura, T.; Hirao, K. A New One-ParameterProgressive Colle−Salvetti-Type Correlation Functional. J. Chem. Phys.1999, 110, 10664−10678.(54) 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.(55) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent andAccurate Ab Initio Parametrization of Density Functional DispersionCorrection (DFT-D) for the 94 Elements H−Pu. J. Chem. Phys. 2010,132, 154104.(56) Becke, A. D. Density-Functional Thermochemistry. V. System-atic Optimization of Exchange−Correlation Functionals. J. Chem. Phys.1997, 107, 8554−8560.(57) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the DampingFunction in Dispersion Corrected Density Functional Theory. J.Comput. Chem. 2011, 32, 1456−1465.(58) Hamprecht, F. A.; Cohen, A. J.; Tozer, D. J.; Handy, N. C.Development and Assessment of New Exchange-Correlation Func-tionals. J. Chem. Phys. 1998, 109, 6264−6271.(59) Zhao, Y.; Truhlar, D. G. Density Functional for Spectroscopy:No Long-Range Self-Interaction Error, Good Performance for Rydbergand Charge-Transfer States, and Better Performance on Average thanB3LYP for Ground States. J. Phys. Chem. A 2006, 110, 13126−13130.(60) Schultz, N. E.; Zhao, Y.; Truhlar, D. G. Databases for TransitionElement Bonding: Metal−Metal Bond Energies and Bond Lengths andTheir Use ToTest Hybrid, HybridMeta, andMeta Density Functionalsand Generalized Gradient Approximations. J. Phys. Chem. A 2005, 109,4388−4403.(61) Peverati, R.; Truhlar, D. G. Screened-Exchange DensityFunctionals with Broad Accuracy for Chemistry and Solid-StatePhysics. Phys. Chem. Chem. Phys. 2012, 14, 16187−16191.(62) Kardahakis, S.; Mavridis, A. First-Principles Investigation of theEarly 3d Transition Metal Diatomic Chlorides and Their Ions, ScCl0,±,TiCl0,±, VCl0,±, and CrCl0,±. J. Phys. Chem. A 2009, 113, 6818−6840.(63) Hayashi, S.; Leonard, C.; Chambaud, G. Ab Initio Study of theLow Lying Electronic States of ZnF and ZnF−. J. Chem. Phys. 2008, 129,044313.(64) Miliordos, E.; Mavridis, A. Electronic Structure and Bonding ofthe Early 3d-TransitionMetal Diatomic Oxides and Their Ions: ScO0,±,TiO0,±, CrO0,±, and MnO0,±. J. Phys. Chem. A 2010, 114, 8536−8572.(65) Mori, Y.; Kawabata, T.; Onodera, G.; Kimura, M. RemarkablySelective Formation of Allenyl and Dienyl Alcohols via Ni-CatalyzedCoupling Reaction of Conjugated Enyne, Aldehyde, and OrganozincReagents. Synthesis 2016, 48, 2385−2395.(66)Mori, Y.; Shigeno, Y.; Luo, Y.; Chan, B.; Onodera, G.; Kimura,M.Ni-Catalyzed Formal Carbonyl-Ene Reaction of Terminal Alkenes viaCarbon Dioxide Insertion. Synlett 2018, 29, 742−746.(67) Chan, B.; Luo, Y.; Kimura,M.Mechanism for Three-ComponentNi-Catalyzed Carbonyl−Ene Reaction for CO2 Transformation: WhatPractical Lessons Do We Learn from DFT Modelling? Aust. J. Chem.2018, 71, 272−278.(68) Flores, L. A.; Murphy, J. G.; Copeland, W. B.; Dixon, D. A.Reaction of CO2 with Groups 4 and 6 TransitionMetal Oxide Clusters.J. Phys. Chem. A 2017, 121, 8719−8727.(69) Fang, Z.; Vasiliu, M.; Peterson, K. A.; Dixon, D. A.Computational Study of Molecular Hydrogen Adsorption over Small(MO2)n Nanoclusters (M = Ti, Zr, Hf; n = 1 to 4). J. Phys. Chem. A2018, 122, 4338−4349.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3621

Page 13: Assessment of DFT Methods for Transition Metals with the …rsc.anu.edu.au/~pgill/papers/186_TMC.pdf · 2019-06-12 · Assessment of DFT Methods for Transition Metals with the TMC151

(70) Chan, B.; Yim, W. L. Accurate Computation of CohesiveEnergies for Small to Medium-Sized Gold Clusters. J. Chem. TheoryComput. 2013, 9, 1964−1970.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.9b00239J. Chem. Theory Comput. 2019, 15, 3610−3622

3622