Subscriber access provided by Kaohsiung Medical University
A: Spectroscopy, Molecular Structure, and Quantum Chemistry
Density Functional Theory for Microwave Spectroscopy of Non-Covalent Complexes: A Benchmark Study Peter Kraus, and Irmgard Frank J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b03345 • Publication Date (Web): 11 May 2018 Downloaded from http://pubs.acs.org on May 15, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
Density Functional Theory for Microwave Spectroscopy of Non-Covalent Complexes: A Benchmark Study P. Kraus∗ and I. Frank Institut f¨ ur Physikalische Chemie und Elektrochemie, Leibniz Universit¨at Hannover, Callinstraße 3A, 30167 Hannover, Germany E-mail:
[email protected] Abstract In this work, we compare the results obtained with 89 computational methods for predicting non-covalent bond lengths in weakly bound complexes. Evaluations for the performance in non-covalent interaction energies and covalent bond lengths obtained from five other datasets are included. The overall best performing density functional is the ωB97M-V method, achieving balanced results across all three categories. For non-covalent geometries the best methods include B97M-V, B3LYP-D3(BJ) and DSDPBEPBE-D3(BJ). The effects of systematic improvement of the density functional approximation and of dispersion corrections are also discussed.
1
1
Introduction
2
For non-covalently bound complexes, a large portion of recent efforts in research, develop-
3
ment, and benchmarking of computational methods focuses on obtaining accurate interaction
4
energies. 1–4 A great amount of progress has been achieved in this field over the recent years.
5
At the highest levels of wavefunction theory (WFT), it has been shown that interaction ener-
6
gies computed with coupled cluster calculations including singles, doubles, and pertubative 1
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
7
triples used with a basis set extrapolated to the complete basis set limit (CCSD(T)/CBS) are
8
in excellent agreement with calculations also including core-core and core-valence correlation,
9
relativistic effects and quadruple excitations. 3 On the other hand, the advent of easy-to-use
10
empirical dispersion corrections, 5 now including three-body interactions and effects of local
11
chemical environment, 6 dynamic polarisabilities of atoms, 7 or non-local correlation, 8 allows
12
density functional theory (DFT) methods to achieve a comparatively good performance in
13
the description of interaction energies, provided large enough basis sets are applied, 4 even
14
when comparatively dated density functionals are used. Additionally, with double-hybrid
15
density functional methods, which include a part of MP2 correlation in addition to the part
16
of Hartree-Fock (HF) exchange used in single-hybrid DFT methods, accuracy comparable to
17
CCSD(T)-based benchmarks is possible. 9,10
18
However, accurate description of energies is only one aspect that makes a computa-
19
tional method generally applicable. For certain structure-sensitive applications, including
20
microwave spectroscopy, an accurate prediction of geometries is equally, if not more im-
21
portant. The data acquisition in such experiments generally starts with a computation of
22
a predicted structure, to obtain a first guess of the rotational constants. If the predicted
23
rotational constants differ significantly (> 10 MHz) from the true values, the potential to
24
waste valuable instrument time is high, especially if a broadband instrument is unavailable.
25
In addition to the data collection issues, accurate structural predictions and ab initio based
26
spectra can decrease the difficulty in interpretation of experimental data. Furthermore, there
27
is currently little guidance for an unbiased choice of computational methods, as accuracy for
28
non-covalent structures does not immediately follow from an accurate prediction of energies.
29
Therefore, in the current work, we attempt to answer the question: Which affordable com-
30
putational method is currently the best for predicting geometries of non-covalently bound
31
complexes? To help us answer this question, we have used the recently developed dataset
32
of semi-experimental non-covalent bond lengths in dimers and trimers (NCDT16) 11 and 89
33
computational methods, including the HF, MP2, CCSD and CCSD(T) wavefunction-based
2
ACS Paragon Plus Environment
Page 2 of 24
Page 3 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
34
methods. On the density functional front, 16 generalised density gradient approximation
35
(GGA) methods, 11 meta-GGA functionals depending also on the kinetic energy density, 28
36
hybrid GGA functionals incorporating HF-exchange, 20 hybrid meta-GGA functionals and
37
8 double-hybrid functionals also incorporating MP2 correlation are compared. Additionally,
38
two composite methods are included: an “affordable” PBEh-3c method intended to predict
39
good geometries 12 and a “near gold-standard” MP2/CBS + δCCSD(T)/3-ζ extrapolated-
40
basis method (shortened to MP2 + δ(T) in the following). The NCDT16 dataset comprises
41
45 non-covalent bond lengths between the heavy atoms of 16 dimers and trimers, obtained
42
using experimental rotational constants corrected for non-equilibrium effects at the double-
43
hybrid level of DFT. While angles are not included in the database, angular information is
44
included by considering more than one bond length for all non-symmetric complexes. 11 The
45
current dataset is somewhat skewed towards dispersion interactions, with 9 out of 16 com-
46
plexes having interaction energies below 4 kJ/mol. However, three complexes with dominant
47
π–π interactions and four H–bonded systems ensure that the interaction energy ranges of
48
5–10 kJ/mol and > 15 kJ/mol, respectively, are also considered. Interaction energies for
49
clusters in the NCDT16 dataset, obtained with the best-performing methods, are presented
50
at the semi-experimental (reSE ) geometries and compared to Eint results at the optimized ge-
51
ometries. To make our analysis more robust, we have carried out further calculations on five
52
other previously published benchmark datasets: a set of semi-experimental bond lengths
53
in small covalently bound compounds (CCse22), 13 its ab initio bond length counterpart
54
(CS20), 14 two datasets of highly-accurate ab initio interaction energies at the prescribed ge-
55
ometries (NCCE31 14 and A24 3 ), and a database of 21 equilibrium CCSD(T)/CBS structures
56
(A21G) 3 to serve as a counterpart to our NCDT16 set.
3
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 4 of 24
Table 1: An overview of the computational methods used in this work, grouped according to their classification (“Jacob’s ladder”): GGAs (F (∇ρ[r])), meta-GGAs (F (∇ρ[r], τ [r])), single hybrids, single meta-GGA hybrids, double hybrids, and coupled cluster methods. GGA BLYP BLYP-D2 BLYP-D3(BJ) BLYP-D3m(BJ) PBE PBE-D2 PBE-D3(BJ) PBE-D3m(BJ) HCTH/120 HCTH/120-D3(BJ) VV10 B97-D2 B97-D3(BJ) B97-D3m(BJ) N12 N12-D3(BJ)
meta-GGA M06-L M06-L-D3 M11-L M11-L-D3(BJ) MN12-L MN12-L-D3(BJ) MN15-L MN15-L-D3 TPSS TPSS-D3(BJ) B97M-V
XHF + GGA HF B3LYP B3LYP-D2 B3LYP-D3 B3LYP-D3(BJ) B3LYP-D3m(BJ) B3PW91 B3PW91-D3(BJ) ωPBE ωPBE-D3(BJ) ωPBE-D3m(BJ) PBE0 PBE0-D2 PBE0-D3 PBE0-D3(BJ) PBE0-D3m(BJ) ωPBE0 LC-VV10 B97 ωB97X ωB97X-D ωB97X-V BHandH BHandHLYP HSE06 HSE06-D3(BJ) SOGGA11-X SOGGA11-X-D3(BJ) PBEh-3c
XHF + meta-GGA M05 M05-D3 M05-2X M05-2X-D3 M06 M06-D3 M06-2X M06-2X-D3 M06-HF M06-HF-D3 M08-HX M08-HX-D3 M08-SO M11 M11-D3(BJ) MN15 MN15-D3(BJ) TPSSH TPSSH-D3(BJ) ωB97M-V
XHF + CMP2 + GGA MP2 B2PLYP B2PLYP-D3(BJ) B2PLYP-D3m(BJ) PBE0-2 DSD-BLYP-D3(BJ) DSD-PBEP86-D3(BJ) DSD-PBEPBE-D3(BJ)
CC CCSD CCSD(T) MP2 + δ(T)
57
2
58
All ab initio calculations were performed using a development version (ver. > 1.2a1.dev781)
59
of the electronic structure code ψ 4 , 15 coupled to the DFT exchange-correlation library
60
Libxc v. 3.0.0. 16 Our modifications to the ψ 4 codebase consisted of adding numerous density
61
functionals which are a part of Libxc v. 3.0.0; they are now included in the development
62
version. The use of such open-source programs is in our view critical for reproducibility and
63
auditing of computational data. As a result, the driver routines used in the current work are
64
available in the Supporting information code archive.
65
Computational methods
All electronic structure calculations were carried out with the density-fitted variants of 4
ACS Paragon Plus Environment
Page 5 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
66
WFT (including MP2, CCSD or CCSD(T)) and DFT methods. An “ultra fine” (99, 590)
67
integration grid was used; for cases where SCF convergence was difficult, an even larger
68
(250, 974) grid was applied with a density damping of 20%. The convergence thresholds in ge-
69
ometry optimisations were tightened to ∆E < 1µEh, Fmax < 60 µEh/a0 , FRM S < 40 µEh/a0 ,
70
∆Rmax < 1 ma0 and ∆RRM S < 1 ma0 . In some complexes, such as for the Ne–OCS dimer,
71
results vary significantly when the default (looser) convergence thresholds are used. For
72
all methods with the exception of coupled cluster and composite calculations, the def2-
73
TZVPD basis set was used 17,18 due to its availability for heavier elements. For the CCSD and
74
CCSD(T) calculations, the may-cc-pVTZ basis set was used 19 to reduce computational cost.
75
The PBEh-3c composite method has been applied with its modified double-ζ basis set. 12
76
Finally, the extrapolated composite method denoted “MP2 + δ(T)” comprises a frozen-core
77
density-fitted MP2 calculation extrapolated towards the basis set limit 20 by using triple- and
78
quadruple-ζ correlation-consistent basis sets (ie. MP2/aug-cc-pV[TQ]Z), 21,22 with further
79
correlation effects (“δ(T)”) approximated from the difference between frozen-core density-
80
fitted CCSD(T) and MP2 energies at the may-cc-pVTZ level. The use of quadruple-ζ basis
81
in the MP2 extrapolation is critical, as the extrapolation from double- and triple-ζ basis
82
sets (ie. MP2/aug-cc-pV[DT]Z) performs considerably worse than the MP2/def2-TZVPD
83
method. An overview of the computational methods used in the current work is shown in
84
Table 1. For full references of all listed methods, see the Supporting information.
85
3
86
The mean absolute errors (MAEs) of selected non-covalent bond lengths (|∆reSE |) averaged
87
88
Results and discussion
over the 16 complexes comprising the NCDT16 dataset (|∆reSE |) are shown in Figure 1 √ along with their associated standard errors (σ(|∆reSE |)/ 16). The best performer is the
89
meta-GGA functional B97M-V, 23 with |∆reSE | = 48 m˚ A. This corresponds to an accuracy
90
better than several double hybrid functionals and even the considerably more expensive
5
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
0
25
50
75
100
100
150
200
HSE06-D3(BJ)
M05-2X ωPBE-D3m(BJ) B3LYP-D3 MP2 PBE-D3m(BJ) MN15-D3(BJ) B3LYP-D3m(BJ)
50
75
100
0
50
300
400
B97-D2
M06 B97 BHandHLYP M06-D3 B3PW91-D3(BJ) N12-D3(BJ) ωPBE0 ωPBE SOGGA11-X-D3(BJ) MN12-L MN12-L-D3(BJ) BHandH TPSSH TPSS M11-L-D3(BJ)
443 461 788 936 1302 1400 1495 1694
M11-L SOGGA11-X HF N12 B3LYP B3PW91 BLYP
B3LYP-D2 M06-HF-D3 PBEh-3c M08-HX-D3 HSE06 B97-D2
MN15 ωB97X PBE-D3(BJ) HSE06-D3(BJ)
200
PBE PBE0 ωB97X-D HCTH/120
VV10 M05 CCSD M05-D3 PBE0-D2 B2PLYP M08-SO M06-L-D3 MN12-SX PBE-D2 M06-HF BLYP-D2 MN15-L MN15-L-D3 M06-L TPSS-D3(BJ) MN12-SX-D3(BJ)
M05-2X-D3 PBE0-D3(BJ) DSD-BLYP-D3(BJ)
100
M08-HX B97-D3(BJ) B97-D3m(BJ)
M11 HCTH/120-D3(BJ) TPSSH-D3(BJ) BLYP-D3(BJ) M11-D3(BJ)
PBE0-D3 M06-2X M06-2X-D3 DSD-PBEP86-D3(BJ) PBE0-D3m(BJ) BLYP-D3m(BJ)
0
B97M-V
ωPBE-D3(BJ)
ωB97M-V ωB97X-V B3LYP-D3(BJ) MP2 + δ(T) CCSD(T) B2PLYP-D3m(BJ)
25
50
B97M-V
B97M-V DSD-PBEPBE-D3(BJ) PBE0-2 LC-VV10 B2PLYP-D3(BJ)
0
0
Page 6 of 24
100 |∆reSE |
150 ˚] / [mA
200
0
100
200
300
Figure 1: The MAE of various computational methods in selected non-covalent bond lengths averaged over the 16 complexes in the NCDT16 dataset. The colour denotes the class of the method (dark gray: GGA, light gray: meta-GGA, blue: global hybrid, light blue: rangeseparated hybrid, purple: meta-GGA hybrid, pink: range-separated meta-GGA hybrid, orange: double hybrid, yellow: coupled cluster). √ Accuracy decreases from left to right. Error SE bar shows one standard error (σ(|∆re |)/ 16). 91
CCSD(T) method. The only other method with a MAE below 50 m˚ A is the double-hybrid
92
DSD-PBEPBE-D3(BJ). 24 Other methods achieving lower MAE than CCSD(T), as well as
93
the extrapolated MP2 + δ(T) composite method, include the dispersion corrected global
94
hybrid B3LYP-D3(BJ), 25–27 two range-separated functionals ωB97X-V 28 and ωB97M-V 29
95
related to the best performing meta-GGA, the original double-hybrid with dispersion cor-
96
rection B2PLYP-D3(BJ), 9,27 and the range-separated functional with non-local correlation
97
LC-VV10. 8 Somewhat surprisingly, the dispersionless double hybrid PBE0-2 10 also performs 6
ACS Paragon Plus Environment
400
Page 7 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
0
A) Ne–complexes 20 40 60
80
B) dispersion dominated 0 8 16 24 32
0
C) π–π interactions 11 22 33 44
0
D) H–bonded 11 22 33
B3LYP-D3(BJ)
B97M-V
BLYP-D3(BJ)
MN12-SX
DSD-PBEPBE-D3(BJ)
ωB97X-V
B3PW91-D3(BJ)
MN12-L
PBE0-D3
BLYP-D3m(BJ)
B97-D3(BJ)
MP2 + δ(T)
BLYP-D3m(BJ)
M05-2X
N12-D3(BJ)
MN12-SX-D3(BJ)
CCSD(T)
B3LYP-D3(BJ)
LC-VV10
MN12-L-D3(BJ)
PBE0-D3m(BJ)
PBE0-D2
B3LYP-D3(BJ)
MN15-D3(BJ)
PBE-D3(BJ)
ωB97M-V
B2PLYP-D3(BJ)
M08-SO
PBE0-2
B2PLYP-D3m(BJ)
ωPBE-D3(BJ)
MN15
B97M-V
B2PLYP-D3(BJ)
ωB97X-V
M06-D3
B3LYP-D3m(BJ)
DSD-PBEPBE-D3(BJ)
TPSSH-D3(BJ)
ωB97X-V
0
20
40
60
80
0
8
16
24
32
0
11
|∆reSE | / [m˚ A]
22
33
44
0
11
22
33
Figure 2: The MAE of the ten best computational methods in subsets of the NCDT16 dataset: A) dispersion dominated complexes including Ne, B) other dispersion dominated complexes, C) complexes with significant π–π interactions and D) H–bonded complexes. Colours as in Fig. 1. 98
very well.
99
A more detailed look into the top performers in the four categories of complexes compris-
100
ing the NCDT16 dataset is shown in Fig. 2. Detailed results for all methods are listed in the
101
Supporting information. The ranking of the methods varies significantly, with only B3LYP-
102
D3(BJ) 25–27 achieving a “top 10” result in three out of the four categories – its performance
103
for the H–bonded dataset is average. The lowest MAE for the five Ne–containing complexes
104
is significantly higher than in the other three categories, with meta-GGA’s performing com-
105
parably poorly: the best meta-GGA hybrid M06-2X-D3 30,31 is only rank 17. On the other
106
hand, for H–bonded complexes, the more recent meta-GGA’s of Truhlar and co-workers 32,33
107
top the field – the overall MAE of the range-separated meta-GGA MN12-SX, 34 which is
108
the top method in this category, is hampered only by the large MAE for the Ne–Ne–N2 O
109
trimer. On the other hand, the highest ranked double hybrid for the H–bonded complexes
110
is the PBE0-2 functional 10 at number 13. The other two classifications show more mixed
111
results. The excellent performance of dispersion-corrected GGA’s for π–π bonded systems is
112
somewhat surprising, especially in the case of the 4th best N12-D3(BJ) 35,36 functional which
113
has a 10× larger overall |∆reSE | at ∼ 270 m˚ A. Also notable are the rather average results 7
ACS Paragon Plus Environment
44
44
The Journal of Physical Chemistry
0
50
100
150
200
Ne–C2 H4 Ne–OCS
Ne–Ne–N2 O Ne–Ar–N2 O
Ne–complexes
Ne–Ar–HCl Ar–C2 H4 Ar–OCS Kr–OCS Ar–Ar–N2 O
dispersion
CS2 –OCS HCCH–HCCH CH2 ClF–HCCH
π–π
HF–H2 CO HCl–H2 CO HCN–H2 CO HF–HF-NH3
0
50
100 |∆reSE | / [m˚ A]
150
H–bonded
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 24
200
Figure 3: The absolute deviations of the non-covalent bond lengths calculated with B97M-V (gray) and MP2 + δ(T) (red) compared to the semi-experimental uncertainties (error bars).
114
of MP2 + δ(T) for π–π interactions, at rank 30.
115
The absolute deviations of the calculated bond lengths from the semi-experimental
116
equilibrium values (|∆reSE |) are shown for the top density functional B97M-V and the
117
MP2 + δ(T) methods in Figure 3. The reported uncertainties in the semi-experimental bond
118
lengths 11 are also included. For Ne–containing complexes, both methods struggle with the
119
Ne–OCS complex and with the position of Ne in the Ne–Ar–N2 O and Ne–Ar–HCl trimers. In
120
other dispersion dominated complexes, the rather poor performance of the composite method
121
for the Kr–OCS dimer could be due to the frozen core approximation or neglection of rela-
122
tivistic effects in the basis sets used. The results of the two methods for π–π interactions is
8
ACS Paragon Plus Environment
Page 9 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 B97M-V DSD-PBEPBE-D3(BJ) PBE0-2 LC-VV10 B2PLYP-D3(BJ) ωB97M-V ωB97X-V B3LYP-D3(BJ) CCSD(T) PBE0-D3 BLYP-D3m(BJ) M05-2X MN15-D3(BJ) BLYP-D3(BJ) M08-SO MN12-SX B97-D3(BJ) MN12-L B3PW91-D3(BJ) N12-D3(BJ)
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 |∆Eint | / [kJ/mol]
Figure 4: MAE values of counterpoise-corrected interaction energies obtained with the listed methods with respect to the MP2 + δ(T) data at reSE geometries (blue) and at optimized geometries (orange). 123
comparable, with the large deviation in one of the CS2 –OCS parameters due to a different
124
angle between the two monomers. In H–bonded complexes, the composite method gener-
125
ally performs better than B97M-V, with the latter method significantly overpredicting the
126
H2 C=O–HCN angle compared to both MP2 + δ(T) and the semi-experimental uncertainty.
127
3.1
128
Counterpoise-corrected interaction energies of the 16 weakly-bound clusters have been cal-
129
culated at the semi-experimental geometries, as well as at the optimised geometries, with
130
the ten best-performing methods. The full results are listed in the Supporting information
131
files. The MAE values with respect to the MP2 + δ(T) results are shown in Fig. 4. The
132
average difference between Eint values at reSE and at the MP2 + δ(T)–optimized geometries
133
is just below 1.1 kJ/mol, with the highest difference (10.8 kJ/mol or 43%) obtained for the
134
HF–H2 CO complex. This increase in interaction energies and associated errors should be
135
considered when further components of binding energies are required, as zero point vibra-
136
tional corrections have to be obtained at the minima.
Interaction energies of the NCDT16 dataset
9
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
137
The MN15-D3(BJ) functional performs rather well for both sets of interaction energies
138
and it is the best out of the 20 methods studied. Among the best methods for non-covalent
139
geometries (top part of Fig. 4), the error in the interaction energies at the minima increases
140
compared to the error at the reSE structures, mainly due to a further overprediction of the
141
interaction energies in H–bonded complexes. Methods without empirical or non-local disper-
142
sion terms generally underpredict the interactions for dispersion-dominated complexes, even
143
when MP2 correlation is included (cf. PBE0-2). The only exception among the functionals
144
shown is M05-2X, which on average overpredicts Eint of Ne–containing complexes by about
145
0.5 kJ/mol. The large absolute error in CCSD(T) data compared to the MP2 + δ(T) results
146
is a systematic underprediction of the interaction energies due to the smaller basis set used,
147
possibly further exacerbated by the application of counterpoise corrections. 37
148
3.2
149
The performance of the methods listed in Table 1 for the accuracy of non-covalent bond
150
lengths has been correlated with the performance observed for five other benchmark datasets,
151
with the overall results shown in Table 2. The six datasets can be split into three pairs:
152
non-covalent semi-experimental and theoretical equilibrium bond lengths (NCDT16 11 and
153
A21G, 3 respectively), counterpoise-corrected interaction energies (A24 3 and NCCE31 14 ),
154
and covalent semi-experimental and theoretical equilibrium bond lengths (CCse22 13 and
155
CS20, 14 respectively). Despite using different basis sets, our results calculated for the five
156
datasets are generally comparable to literature data, obtaining qualitative agreement with
157
the A21G data of Mardirossian and Head-Gordon, 28 an agreement within 0.64 kJ/mol with
158
the results of Peverati and Truhlar 14 and Roch 38 for the A24 and NCCE31 datasets, or
159
within 1 m˚ A for the CS20 results of Peverati and Truhlar. 14
Correlations with other datasets
160
The average errors of the methods for the two datasets of non-covalent bond lengths are
161
strongly correlated, with ρNCDT16, A21G = 0.94. The MAEs obtained for the A21G dataset
162
are much lower than for the NCDT16 dataset, as in the A21G benchmark the whole bond 10
ACS Paragon Plus Environment
Page 10 of 24
Page 11 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
Table 2: Pearson’s correlation coefficient between studied benchmark datasets.
NCDT16 A21G A24 NCCE31 CCse22
A21G 0.94
A24 0.84 0.87
NCCE31 0.56 0.62 0.74
CCse22 0.21 0.21 0.31 0.54
CS20 0.24 0.19 0.38 0.48 0.81
163
matrix is assessed. The only bond length present in both datasets – the r(C–C) of the
164
T-shaped HCCH–HCCH dimer – differs by 70 m˚ A, with CCSD(T)/CBS underpredicting
165
A from our CCSD(T) and the reSE . This difference is consistent with the |∆reSE | ∼ 65 m˚
166
MP2 + δ(T) results.
167
The correlation between the two datasets of counterpoise-corrected interaction ener-
168
gies is also strong (ρA24, NCCE31 = 0.74).
Similarly to the NCDT16 data above, our
169
counterpoise-corrected CCSD(T)/may-cc-pVTZ calculation consistently overpredicts the ref-
170
erence CP-CCSD(T)/CBS binding energies with |∆Eint | = 1.5 kJ/mol. As our extrapolated
171
MP2 + δ(T) results have a much lower |∆Eint | = 0.1 kJ/mol, the difference is attributed
172
to the basis set incompleteness error. The correlation between the covalent bond length
173
datasets (ρCCse22, CS20 = 0.81) is weaker than for the non-covalent bond datasets.
174
The SOGGA11-X 39 method has been previously highlighted as an excellent functional
175
for covalent bond lengths. 40 However, for non-covalent bond lenghts, its performance is poor
176
(|∆reSE | > 750 m˚ A, 300 m˚ A with D3(BJ)) even when compared to non-dispersion corrected
177
single hybrids, such as PBE0 (|∆reSE | = 182 m˚ A). The lack of transferability between er-
178
rors in covalent and non-covalent bonds is also confirmed by the low correlation coefficient
179
ρNCDT16,CCse22 = 0.21. On the other hand, the comparably high correlation coefficient be-
180
tween interaction energies and non-covalent bonds (ρNCDT16,A24 = 0.84) shows that accurate
181
description of energetics generally corresponds to accurate geometries – the notable excep-
182
tions with poorly predicted geometries despite |∆Eint | < 1 kJ/mol are ωB97X-D, 41 B97-
183
D3, 6,42 TPSS-D3(BJ) 27,43 and M08-HX(-D3). 44,45 Curiously, the two half-and-half hybrid 11
ACS Paragon Plus Environment
The Journal of Physical Chemistry
2.0
|∆Eint | / [kJ/mol]
1.5
0
15
150
200 (A)
PBE-D3m(BJ) PBE0-2
M06-L-D3 VV10 M05
7
PBEh-3c B97-D2
MP2 MN15-D3(BJ) M06-2X-D3 5 HSE06-D3(BJ) M08-HX-D3 LC-VV10 1 M08-SO 4 B97M-V TPSS-D3(BJ) 6 TPSSH-D3(BJ) 2 ωB97M-V 3 ωPBE-D3m(BJ) ωB97X-V M05-2X-D3 MP2+δ(T)
(B)
(A): PBE0-D3 BLYP-D3m(BJ) (B): M06-2X-D3 PBE-D3m(BJ) (C): M05-2X-D3 MN15-L (D): MN15-D3(BJ) (E): TPSSH-D3(BJ) VV10 B97-D2 (F): ωB97M-V MN12-SX (G): ωPBE-D3m(BJ) (H): DSD-PBEPBE-D3(BJ) TPSS-D3(BJ) M08-SO (I): MP2 + δ(T)
10
LC-VV10 B3LYP-D3(BJ)
PBEh-3c
M11
M06-HF HCTH/120-D3(BJ) B97M-V C G MP2 M05 B ωB97X-V D E M06-L-D3 PBE0-2 A CCSD H F HSE06-D3(BJ) I ωB97X B2PLYP-D3(BJ) DSD-PBEP86-D3(BJ) CCSD(T) DSD-BLYP-D3(BJ)
5 0
100
B2PLYP-D3(BJ) DSD-BLYP-D3(BJ) M06-HF DSD-PBEP86-D3(BJ) DSD-PBEPBE-D3(BJ) MN15-L BLYP-D3m(BJ) CCSD(T) B3LYP-D3(BJ) M11 ωB97X HCTH/120-D3(BJ)
PBE0-D3
0.5 0.0 20
50
(1): (2): (3): (4): (5): (6): (7):
1.0
SE | / [m˚ |∆rcov A]
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 24
0
50
100 |∆reSE |
M08-HX-D3
150
200
/ [m˚ A]
Figure 5: Correlation diagrams for performance in prediction of non-covalent bonds (|∆reSE |, NCDT16 dataset): (A) correlation with MAE of counterpoise-corrected interaction energies SE |, CCse22). The dashed (|∆Eint |, A24); (B) correlation with MAE of covalent bonds (|∆rcov line shows average performance of the dataset. For clarity, only selected methods are shown. Colours as in Figure 1. 184
functionals 25 (BHandH with 50% Slater exchange and BHandHLYP with 50% B88 exchange)
185
perform much better than both three-parameter functionals (B3PW91 25 and B3LYP 25,26 ).
186
This is due to the poor description of dispersion in the three-parameter functionals in all rare
187
gas complexes (see Supporting information for details). However, with dispersion corrections,
188
the B3LYP-D3(BJ) method 25–27 is one of the top methods.
189
The latter two correlations are shown in graphical form in Figure 5. Only methods with
190
|∆reSE | < 200 m˚ A in the NCDT16 dataset are listed, and only the best-performing dispersion
191
correction is shown for each functional, if it is better than the parent functional. While the
12
ACS Paragon Plus Environment
Page 13 of 24
192
B97M-V 23 functional is the best studied method for non-covalent bond lengths, the related
193
range-separated ωB97M-V 29 offers a more balanced performance for accurate interaction
194
energies (Figure 5 (A), |∆Eint | = 0.3 kJ/mol) and covalent bond lengths (Figure 5 (B),
195
SE | = 4 m˚ |∆rcov A). The most balanced double hybrid functional remains DSD-PBEPBE-
196
D3(BJ), 24 as its somewhat large |∆Eint | of 0.5 kJ/mol in the A24 dataset is offset by its
197
remarkably good performance for the NCCE31 dataset. For non-meta-GGA single hybrid
198
functionals, good overall functionals are the range-separated ωB97X-V 28 and global hybrid
199
B3LYP-D3(BJ). 25–27 With the exception of B97M-V, the meta-GGA functionals achieve
200
worse results than dispersion-corrected GGA functionals. The most balanced GGA func-
201
SE | = 9 m˚ A, as the tional is HCTH/120-D3(BJ) 27,46 with |∆Eint | = 1.3 kJ/mol and |∆rcov
202
dispersion-corrected forms of PBE and BLYP have comparably large deviation in covalent
203
SE | > 16 m˚ A). bond lengths (|∆rcov
204
3.3
Climbing the Jacob’s ladder
|∆reSE | / [m˚ A]
200 150
(A)
100 50 GGA
GSH
RSH DH
DSD
1.5 GGA
GSH
RSH DH
DSD
|∆Eint | / [kJ/mol]
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
1.0 0.5 (B)
0.0
Figure 6: Performance of related functionals in non-covalent interactions: (A) bond lengths, (B) interaction energies. Colour denotes functional family (green: BLYP, red: PBE, blue: B97, cyan: meta-B97) and dispersion corrections (yellow: empirical, black: VV10). Vertical lines delineate GGA’s from global single hybrids, range-separated hybrids, double hybrids and spin-component scaled double hybrids.
13
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
205
The large amount of collected data allows us to investigate the effects of systematic
206
improvement of DFT methods by climbing the so-called “Jacob’s ladder”. The resulting av-
207
erage errors in non-covalent bond lengths and interaction energies are shown in Figure 6. In
208
the BLYP-based series (BLYP-D3(BJ), B3LYP-D3(BJ), B2PLYP-D3(BJ) and DSD-BLYP-
209
D3(BJ); green), a significant improvement in prediction of the structure is achieved by in-
210
clusion of HF exchange, while the inclusion of MP2 correlation shows mixed results. In
211
the PBE-based series (PBE-D3(BJ), PBE0-D3(BJ), ωPBE-D3(BJ), LC-VV10, PBE0-2 and
212
DSD-PBEPBE-D3(BJ); red), inclusion of HF-exchange leads to a better prediction of inter-
213
action energies in addition to the structural parameters, especially in the range-separated
214
form and with inclusion of the VV10 non-local correlation. The dispersionless PBE0-2 dou-
215
ble hybrid functional performs well for structures, but interaction energies are not improved
216
from the related single hybrids, as a consequence of the lack of dispersion corrections. In
217
the B97-based data (blue), the results are mixed. Without dispersion corrections, the range-
218
separated hybrid ωB97X performs quite poorly for interaction energies - worse than the
219
original B97-D2 form. The empirical dispersion correction included in ωB97X-D improves
220
the agreement in interaction energies, but significantly degrades the structural performance;
221
inclusion of the VV10 non-local correlation in ωB97X-V remedies this problem.
222
3.4
223
It is widely accepted 47 that correctly scaled dispersion corrections can greatly improve the
224
agreement in non-covalent interaction energies. For non-covalent bond lengths, near-MP2
225
quality can be achieved with dispersion corrected GGAs. 27 This statement is further con-
226
firmed by our overall results (cf. Figure 1) and is especially pronounced for π–π interactions
227
(cf. Figure 2). For the “worst case” of BLYP and B3LYP functionals, which do not describe
228
dispersion interactions well, the inclusion of D2 correction 5 already reduces |∆reSE | at least
229
by a factor of 7. However, the D3 and D3(BJ) corrections 6,27 perform much better, with
230
further reduction in MAE’s by a factor of 2. The the revised version (D3m(BJ)) 48 shows no
Role of dispersion corrections
14
ACS Paragon Plus Environment
Page 14 of 24
Page 15 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
231
systematic improvement compared to the original Becke-Johnson-damped version (D3(BJ)).
232
However, the most accurate method for improving non-covalent bond lengths is the non-local
233
correlation in VV10, 8 with 4 out of 5 methods including this component performing better
234
than CCSD(T). In light of this, the recently developed spin-component scaled double hybrids
235
with non-local correlation in place of the empirical dispersion term (ie. DSD-PBEPBE-NL)
236
are likely to achieve good results, 49 however the availability of such methods is limited. On
237
the other hand, the non-local correlation approach has a higher computational cost than
238
empirical dispersion corrections, especially as the implementation of analytical gradients
239
for these functionals is not widespread. It would be interesting to compare the current re-
240
sults to methods including with the recently-developed D4 correction, 7 unfortunately their
241
availability is also currently limited.
242
The importance of developing density functionals with systematic consideration of dis-
243
persion corrections can also be highlighted. The dispersion-corrected spin-component-scaled
244
double hybrids of Kozuch and Martin 24,50 as well as the recent functionals of Mardirossian
245
and Head-Gordon 23,28,29 were systematically parametrised with dispersion corrections, and
246
as a result show good performance for dispersion dominated complexes. However, some dis-
247
persionless functionals trained on datasets including non-covalent interactions, such as the
248
functionals from the Truhlar group, 30,33,34,51 are excellent for H–bonded systems even with-
249
out empirical dispersion terms. While applying such further corrections to these functionals
250
is possible, 36,45 it does not lead to a systematic improvement of non-covalent geometries.
251
4
252
In the current work, we have benchmarked 89 ab initio methods using the NCDT16 dataset 11
253
to establish their performance for prediction of non-covalent bond lengths, and five other
254
reference datasets to analyse correlations with accuracy for interaction energies and cova-
255
lent bonds. The best overall method is the ωB97M-V range-separated meta-GGA hybrid
Conclusions
15
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
256
functional of Mardirossian and Head-Gordon, 29 thanks to its balanced performance for non-
257
covalent structures, interaction energies and covalent bond lengths. The best results in non-
258
covalent structures were achieved by the related meta-GGA B97M-V. 23 Other recommended
259
methods include the single hybrid B3LYP-D3(BJ) 25–27 with excellent results for complexes
260
with binding energies below 10 kJ/mol and its widespread availability, and the double hy-
261
brid DSD-PBEPBE-D3(BJ) 24 which is considerably more accurate in interaction energies.
262
For π–π interactions, the B97-D3 method 6,42 performs rather well, while for H–bonded com-
263
plexes the newer Minnesota functionals from the MN12 family 32,34 and MN15-D3(BJ) 33,45
264
are excellent performers.
265
The results also highlight that the ranking of a method for covalent bonds is not neces-
266
sarily transferable to non-covalent bonds, while a low average error in interaction energies
267
generally correlates with the quality of prediction of non-covalent bond lengths. Additionally,
268
at the highest level of theory used in the current work (MP2/CBS + δCCSD(T)/3-ζ), the
269
differences in interaction energies between the semi-experimental and optimized geometries
270
can be as large as 40% (HF–H2 CO complex). The MAE between the two sets of structures of
271
62 m˚ A confirms the importance of validation of computational methods against experimental
272
observables.
273
Furthermore, systematic improvements of the density functional approximations by
274
range-separation, MP2 correlation and empirical dispersion generally improve the agree-
275
ment with reference interaction energies. However, the most consistent functionals for non-
276
covalent bond lengths contain the non-local correlation from the VV10 8 functional, and are
277
systematically parametrised with the dispersion terms taken into account.
278
Acknowledgements
279
PK would like to thank Dr. S. Lehtola for his help with Libxc, Dr. D.G.A. Smith and
280
Dr. L.A. Burns for their help with ψ 4 , and Dr. D.A. Obenchain for his support and helpful
16
ACS Paragon Plus Environment
Page 16 of 24
Page 17 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
281
comments. This work was partially carried out on the Leibniz Universit¨at Hannover compute
282
cluster, which is funded by the Leibniz Universit¨at Hannover, the Lower Saxony Ministry of
283
Science and Culture (MWK) and the German Research Association (DFG).
284
Supporting information
285
A spreadsheet with the MAE values for all six datasets and all complexes in the NCDT16
286
dataset, a document containing detailed references for all functionals, an archive of Python
287
driver routines for performing the benchmarks, an archive of optimized structures with se-
288
lected methods.
289
References
290
(1) Zhao, Y.; Truhlar, D. G. Benchmark databases for nonbonded interactions and their
291
use to test density functional theory. J. Chem. Theory Comput. 2005, 1, 415–432.
292
(2) Hohenstein, E. G.; Sherrill, C. D. Wavefunction methods for noncovalent interactions.
293
Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 304–326.
294
ˇ aˇc, J.; Hobza, P. Describing noncovalent interactions beyond the common approx(3) Rez´
295
imations: How accurate is the “gold standard,” CCSD(T) at the complete basis set
296
limit? J. Chem. Theory Comput. 2013, 9, 2151–2155.
297
298
299
300
301
ˇ aˇc, J.; Hobza, P. Benchmark calculations of interaction energies in noncovalent (4) Rez´ complexes and their applications. Chem. Rev. 2016, 116, 5038–5071. (5) Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. (6) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio
17
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
302
parametrization of density functional dispersion correction (DFT-D) for the 94 elements
303
H-Pu. J. Chem. Phys. 2010, 132, 154104.
304
305
306
307
308
309
310
311
(7) Caldeweyher, E.; Bannwarth, C.; Grimme, S. Extension of the D3 dispersion coefficient model. J. Chem. Phys. 2017, 147, 034112. (8) Vydrov, O. A.; Van Voorhis, T. Nonlocal van der Waals density functional: The simpler the better. J. Chem. Phys. 2010, 133, 244103. (9) Grimme, S. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 2006, 124, 034108. (10) Chai, J. D.; Mao, S. P. Seeking for reliable double-hybrid density functionals without fitting parameters: The PBE0-2 functional. Chem. Phys. Lett. 2012, 538, 121–125.
312
(11) Kraus, P.; Obenchain, D. A. D.; Frank, I. Benchmark-quality semiexperimental struc-
313
tural parameters of van der Waals complexes. J. Phys. Chem. A 2018, 122, 1077–1087.
314
(12) Grimme, S.; Brandenburg, J. G.; Bannwarth, C.; Hansen, A. Consistent structures and
315
interactions by density functional theory with small atomic orbital basis sets. J. Chem.
316
Phys. 2015, 143, 054107.
317
(13) Piccardo, M.; Penocchio, E.; Puzzarini, C.; Biczysko, M.; Barone, V. Semi-experimental
318
equilibrium structure determinations by employing B3LYP/SNSD anharmonic force
319
fields: Validation and application to semirigid organic molecules. J. Phys. Chem. A
320
2015, 119, 2058–2082.
321
(14) Peverati, R.; Truhlar, D. G. Quest for a universal density functional: the accuracy
322
of density functionals across a broad spectrum of databases in chemistry and physics.
323
Philos. Trans. A. Math. Phys. Eng. Sci. 2014, 372, 20120476.
324
(15) Parrish, R. M. et al. Psi4 1.1: An open-source electronic structure program emphasizing
18
ACS Paragon Plus Environment
Page 18 of 24
Page 19 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
325
automation, advanced libraries, and interoperability. J. Chem. Theory Comput. 2017,
326
13, 3185–3197.
327
(16) Lehtola, S.; Steigemann, C.; Oliveira, M. J.; Marques, M. A. Recent developments in
328
LIBXC – A comprehensive library of functionals for density functional theory. Soft-
329
wareX 2018, 7, 1–5.
330
331
(17) Rappoport, D.; Furche, F. Property-optimized Gaussian basis sets for molecular response calculations. J. Chem. Phys. 2010, 133, 134105.
332
(18) Hellweg, A.; Rappoport, D. Development of new auxiliary basis functions of the Karl-
333
sruhe segmented contracted basis sets including diffuse basis functions (def2-SVPD,
334
def2-TZVPPD, and def2-QVPPD) for RI-MP2 and RI-CC calculations. Phys. Chem.
335
Chem. Phys. 2015, 17, 1010–1017.
336
(19) Papajak, E.; Zheng, J.; Xu, X.; Leverentz, H. R.; Truhlar, D. G. Perspectives on basis
337
sets beautiful: Seasonal plantings of diffuse basis functions. J. Chem. Theory Comput.
338
2011, 7, 3027–3034.
339
(20) Halkier, A.; Helgaker, T.; Jørgensen, P.; Klopper, W.; Olsen, J. Basis-set convergence
340
of the energy in molecular Hartree–Fock calculations. Chem. Phys. Lett. 1999, 302,
341
437–446.
342
(21) Kendall, R. A.; Dunning, T. H.; Harrison, R. J. Electron affinities of the first-row
343
atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96,
344
6796–6806.
345
(22) Woon, D. E.; Dunning, T. H. Gaussian basis sets for use in correlated molecular cal-
346
culations. X. The atoms aluminum through argon revisited. J. Chem. Phys. 1993, 98,
347
1358–1371.
19
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
348
(23) Mardirossian, N.; Head-Gordon, M. Mapping the genome of meta-generalized gradient
349
approximation density functionals: The search for B97M-V. J. Chem. Phys. 2015, 142,
350
074111.
351
(24) Kozuch, S.; Martin, J. M. L. Spin-component-scaled double hybrids: An extensive
352
search for the best fifth-rung functionals blending DFT and perturbation theory. J.
353
Comput. Chem. 2013, 34, 2327–2344.
354
355
(25) Becke, A. D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377.
356
(26) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab initio calculation
357
of vibrational absorption and circular dichroism spectra using density functional force
358
fields. J. Phys. Chem. 1994, 98, 11624–11627.
359
360
(27) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465.
361
(28) Mardirossian, N.; Head-Gordon, M. ωB97X-V: A 10-parameter, range-separated hy-
362
brid, generalized gradient approximation density functional with nonlocal correlation,
363
designed by a survival-of-the-fittest strategy. Phys. Chem. Chem. Phys. 2014, 16, 9904.
364
(29) Mardirossian, N.; Head-Gordon, M. ωB97M-V: A combinatorially optimized, range-
365
separated hybrid, meta-GGA density functional with VV10 nonlocal correlation. J.
366
Chem. Phys. 2016, 144, 0–23.
367
(30) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group ther-
368
mochemistry, thermochemical kinetics, noncovalent interactions, excited states, and
369
transition elements: Two new functionals and systematic testing of four M06-class
370
functionals and 12 other function. Theor. Chem. Acc. 2008, 120, 215–241.
20
ACS Paragon Plus Environment
Page 20 of 24
Page 21 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
371
(31) Goerigk, L.; Grimme, S. A thorough benchmark of density functional methods for gen-
372
eral main group thermochemistry, kinetics, and noncovalent interactions. Phys. Chem.
373
Chem. Phys. 2011, 13, 6670–6688.
374
(32) Peverati, R.; Truhlar, D. G. An improved and broadly accurate local approximation
375
to the exchangecorrelation density functional: The MN12-L functional for electronic
376
structure calculations in chemistry and physics. Phys. Chem. Chem. Phys. 2012, 14,
377
13171–13174.
378
(33) Yu, H. S.; He, X.; Li, S. L.; Truhlar, D. G. MN15: A Kohn–Sham global-hybrid
379
exchange-correlation density functional with broad accuracy for multi-reference and
380
single-reference systems and noncovalent interactions. Chem. Sci. 2016, 7, 5032–5051.
381
(34) Peverati, R.; Truhlar, D. G. Screened-exchange density functionals with broad accuracy
382
for chemistry and solid-state physics. Phys. Chem. Chem. Phys. 2012, 14, 16187–16191.
383
(35) Peverati, R.; Truhlar, D. G. Exchange-correlation functional with good accuracy for
384
both structural and energetic properties while depending only on the density and its
385
gradient. J. Chem. Theory Comput. 2012, 8, 2310–2319.
386
387
388
389
(36) Goerigk, L. Treating London-dispersion effects with the latest Minnesota density functionals: Problems and possible solutions. J. Phys. Chem. Lett. 2015, 6, 3891–3896. (37) Mentel, M.; Baerends, E. J. Can the counterpoise correction for basis set superposition effect be justified? J. Chem. Theory Comput. 2014, 10, 252–267.
390
(38) Roch, L. M.; Baldridge, K. K. Dispersion-corrected spin-component-scaled double-
391
hybrid density functional theory: Implementation and performance for non-covalent
392
interactions. J. Chem. Theory Comput. 2017, 13, 2650–2666.
393
(39) Peverati, R.; Truhlar, D. G. Communication: A global hybrid generalized gradient
394
approximation to the exchange-correlation functional that satisfies the second-order 21
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
395
density-gradient constraint and has broad applicability in chemistry. J. Chem. Phys.
396
2011, 135, 191102.
397
´ Savarese, M.; Su, N. Q.; P´erez-Jim´enez, A. ´ J.; Xu, X.; Sancho(40) Br´emond, E.;
398
Garc´ıa, J. C.; Adamo, C. Benchmarking density functionals on structural parameters of
399
small-/medium-sized organic molecules. J. Chem. Theory Comput. 2016, 12, 459–465.
400
(41) Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with
401
damped atomatom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–
402
6620.
403
404
(42) Becke, A. D. Density-functional thermochemistry. V. Systematic optimization of exchange-correlation functionals. J. Chem. Phys. 1997, 107, 8554–8560.
405
(43) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the density functional
406
ladder: Non-Empirical meta-generalized gradient approximation designed for molecules
407
and solids. Phys. Rev. Lett. 2003, 91, 146401.
408
(44) Zhao, Y.; Truhlar, D. G. Exploring the limit of accuracy of the global hybrid meta den-
409
sity functional for main-group thermochemistry, kinetics, and noncovalent interactions.
410
J. Chem. Theory Comput. 2008, 4, 1849–1868.
411
(45) Goerigk, L.; Hansen, A.; Bauer, C. A.; Ehrlich, S.; Najibi, A.; Grimme, S. A look
412
at the density functional theory zoo with the advanced GMTKN55 database for gen-
413
eral main group thermochemistry, kinetics and noncovalent interactions. Phys. Chem.
414
Chem. Phys. 2017, 19, 32184–32215.
415
416
417
(46) Boese, A. D.; Doltsinis, N. L.; Handy, N. C.; Sprik, M. New generalized gradient approximation functionals. J. Chem. Phys. 2000, 112, 1670–1678. (47) Kruse, H.; Goerigk, L.; Grimme, S. Why the standard B3LYP/6-31G* model chemistry
22
ACS Paragon Plus Environment
Page 22 of 24
Page 23 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Physical Chemistry
418
should not be used in DFT calculations of molecular thermochemistry: Understanding
419
and correcting the problem. J. Org. Chem. 2012, 77, 10824–10834.
420
(48) Smith, D. G. A.; Burns, L. A.; Patkowski, K.; Sherrill, C. D. Revised damping parame-
421
ters for the D3 dispersion correction to density functional theory. J. Phys. Chem. Lett.
422
2016, 7, 2197–2203.
423
(49) Kesharwani, M. K.; Karton, A.; Martin, J. M. Benchmark ab initio conformational
424
energies for the proteinogenic amino acids through explicitly correlated methods. As-
425
sessment of density functional methods. J. Chem. Theory Comput. 2016, 12, 444–454.
426
(50) Kozuch, S.; Martin, J. M. L. DSD-PBEP86: in search of the best double-hybrid DFT
427
with spin-component scaled MP2 and dispersion corrections. Phys. Chem. Chem. Phys.
428
2011, 13, 20104–20107.
429
430
(51) Peverati, R.; Truhlar, D. G. Improving the accuracy of hybrid meta-GGA density functionals by range separation. J. Phys. Chem. Lett. 2011, 2, 2810–2817.
23
ACS Paragon Plus Environment
The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
431
Page 24 of 24
TOC Graphic NCDT16 reSE
CCS 5 D(T) B97M-V N1 J) PBE M B3LYP B ( 0 -2 -D3
+
24
ACS Paragon Plus Environment
=
1 2