Article pubs.acs.org/JPCA

Double-Hybrid Density Functionals Free of Dispersion and Counterpoise Corrections for Non-Covalent Interactions Feng Yu* Department of Physics, School of Science, Xi’an Technological University, No. 4 Jinhua North Road, Xi’an, Shaanxi 710032, China S Supporting Information *

ABSTRACT: We have optimized two double-hybrid density functionals (DHDFs) within the frameworks of B2PLYP and mPW2-PLYP against the S22B database. These two functionals are denoted as B2NC-PLYP and mPW2NC-PLYP, where “NC” represents noncovalent interaction. The DHDFs of B2NC-PLYP and mPW2NC-PLYP are optimized free of dispersion and counterpoise corrections with triple-ζ quality basis sets. Combined with the aug-cc-pVTZ basis set, these two functionals are further assessed with the S66 database. According to our computations, both the B2NC-PLYP and mPW2NC-PLYP functionals seem to be competent for investigating noncovalent interactions. Note that the triple-ζ quality basis sets with adequate polarization and diffuse functions should be employed for practical applications. However, different exchange and correlation functionals may be selected and/or modified to reduce the amount of the Fock-exchange in the future. improves the performances of the DHDFs.55 Without additional corrections, the XYG327 and XYGJ-OS28 are valuable functionals for studying noncovalent interactions.47 With the help of the localized molecular orbital energy decomposition analyses (LMO-EDA),62 two parameters of cHF = 0.81 and cMP2 = 0.52 for the framework of the B2PLYP functional were extrapolated with five HCOOH···C6H6 complexes.56 The reference molecular geometries and interaction energies were taken from ref 63. After they were examined with the noncovalent complexation energies database (NCCE31/ 05),8,64,65 we found that these two parameters for the framework of the B2PLYP functional were appropriate for investigating noncovalent interactions,56 at least suitable for the small and medium-sized noncovalent systems. And, thus, we denoted these two parameters for the framework of the B2PLYP functional as B2N-PLYP, where “N” stood for noncovalent interaction.56 However, the training set for the B2N-PLYP functional was very small, and the corresponding parameters should be improved with a well-designed training set. In this work, we have optimized the parameters of cHF and cMP2 for the frameworks of the B2PLYP and mPW2-PLYP functionals with the S22 database.66−69 Note that the S22B69 revision has been utilized as reference. Subsequently, the optimized parameters are further assessed with the S66 database.70,71 The main purpose of this work is to treat noncovalent interactions within the framework of the DHDF,

1. INTRODUCTION Different types of noncovalent interactions are fundamental topics in chemistry, because noncovalent interactions play significant roles in cluster chemistry, supramolecular chemistry, biochemistry, solution chemistry, surface chemistry, and so on.1,2 Therefore, developing and assessing ab initio methods and density functionals for noncovalent interactions are very important for theoretical and computational chemistry.1−15 However, elucidating the natures of noncovalent interactions is also important,1−3 and the favorable computational tools should perform well on various noncovalent interactions with different essences. Double-hybrid density functionals (DHDFs) have grown to be very powerful tools for thermochemistry, thermochemical kinetics, noncovalent interaction, etc.16−57 The B2PLYP18 functional has inspired rapid development of various DHDFs,19−57 and the exchange−correlation energy Exc for the B2PLYP functional was defined as Exc = (1 − c HF)ExGGA + c HFExHF + (1 − c MP2)EcGGA + c MP2EcPT2

(1)

To study noncovalent interactions, the dispersion corrections9,58−60 have been applied to the DHDFs.20,22,23,34−36,40 The DHDFs with dispersion corrections perform very well on different types of noncovalent interactions. For example, the B2PLYP-D3 20,60 functional is a reasonable choice for investigating noncovalent interactions based on systematic evaluations.10,61 Furthermore, the DHDFs with long-range corrections perform well on noncovalent interactions,25,26,54 and the nonlocal van der Waals method also significantly © 2014 American Chemical Society

Received: January 17, 2014 Revised: April 6, 2014 Published: April 10, 2014 3175

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

3. RESULTS AND DISCUSSION 3.1. Assessment of Different (cHF, cMP2) Combinations for the Frameworks of the B2PLYP and mPW2-PLYP Functionals with the S22B Database. Different (cHF, cMP2) combinations for the framework of the B2PLYP functional have been assessed with the S22B database. The 6-311+G(2df,2p), i.e., MG3S for the S22B database, has been employed for single point energy computations. As listed in Table 1, the B2N-PLYP

and further corrections specifically designed for the DHDFs such as dispersion,20,22,23,34−36,40 long-range,25,26,54 and nonlocal van der Waals55 corrections could be ignored.

2. COMPUTATIONAL METHODS The databases of S22,66−69 S66,70,71 and ADIM6 (interaction energy database for six n-alkane dimers)40,72,73 for noncovalent interactions were employed in the present work. The S22B69 revision of the S22 database was used as the training set for the optimizations of the cHF and cMP2 parameters for the frameworks of the B2PLYP and mPW2-PLYP functionals. And then, the S6670,71 database was utilized as a validation set, and the corresponding reference interaction energies were taken from ref 71. As to the methane dimer in the S22 database, the dispersion component for the total interaction energy was −0.89 kcal/mol, whereas the corresponding electrostatic component was only −0.19 kcal/mol 74 from accurate symmetry-adapted perturbation theory (SAPT)75,76 computation. Therefore, we deduced that the dispersion components were extremely important for the dimers in the ADIM6 database. Accordingly, the ADIM6 database was tested in this work. Additionally, the interaction potential wells for some rare gas dimers77 were also tested. The basis sets of 6-31+G(d,p),78−82 6-311+G(d,p),82−84 6311+G(2df,2p), 82−84 6-311+G(2df,2pd), 82−84 6-311++G(3df,3pd),82−84 aug-cc-pVDZ,85−88 aug-cc-pVTZ,85−88 ccpVTZ,85 def2-TZVPP,89 ma-TZVPP,90 def2-QZVP,89 and maQZVP90 were employed for single point energy computations. The basis set superposition error (BSSE) was not corrected with the counterpoise (CP) method91 unless otherwise noted. For the S22 database, the 6-311+G(2df,2p) basis set was identical to the MG3S basis set defined by Truhlar and coworkers,92 which was accurate enough for relative energy computations. Therefore, the basis set of 6-311+G(2df,2p) was utilized in the optimization process of the cHF and cMP2 parameters within the framework of B2PLYP. In detail, the computations with the 6-31+G(d,p), 6-311+G(d,p), 6-311+G(2df,2p), and aug-cc-pVDZ basis sets were performed with the General Atomic and Molecular Electronic Structure System (GAMESS) program,93,94 and other computations with the 6311+G(2df,2pd), 6-311++G(3df,3pd), aug-cc-pVTZ, cc-pVTZ, def2-TZVPP, ma-TZVPP, def2-QZVP and ma-QZVP basis sets were carried out with the ORCA program.95 As to the computations related to the framework of mPW2-PLYP, only the ORCA program was employed. The corresponding cHF and cMP2 parameters were optimized with the 6-311+G(2df,2pd) basis set. Comparisons between the GAMESS and ORCA programs on single point energy computations were presented in the Supporting Information (Table S1). It should be emphasized that the chemical core electrons were frozen in the present work. In addition, the Lebedev (96,302) integration grid96 was employed for the computations performed with the GAMESS program, whereas the Grid4 was utilized for the computations carried out with the ORCA program. On the basis of our previous work,97 these two integration grids were suitable for single point energy computations. The performances of the DHDFs with different basis sets were evaluated with the signed mean deviation (MD), mean absolute deviation (MAD), and mean absolute percent deviation (MA%D). The corresponding definitions had been described in ref 10.

Table 1. Performances of Different (cHF, cMP2) Combinations for the Framework of the B2PLYP Functional with the S22B Databasea DHDF B2PLYP B2T-PLYP B2GP-PLYP B2K-PLYP B2N-PLYP

(cHF,cMP2) (0.53, (0.60, (0.65, (0.72, (0.81,

0.27) 0.31) 0.36) 0.42) 0.52)

MD

MAD

MA%D

1.62 1.26 0.96 0.57 0.05

1.64 1.29 0.99 0.63 0.18

39.3 31.8 25.1 17.0 6.4

a

The 6-311+G(2df,2p) basis set is utilized for single point energy computations, and the MD (kcal/mol), MAD (kcal/mol), and MA%D are listed as comparisons.

functional performs relatively well with the S22B database. On the basis of the performance of the B2N-PLYP functional, we have examined a total of 35 (cHF, cMP2) combinations around the B2N-PLYP functional against the S22B database. As shown in Figure 1a, the range of the MADs of the tested (cHF, cMP2) combinations is about 0.15−0.30 kcal/mol. As a result, the MAD of the (cHF = 0.81, cMP2 = 0.55) combination is only 0.15 kcal/mol, and the corresponding MA%D is 5.9%. Therefore, we denote the (cHF = 0.81, cMP2 = 0.55) combination for the framework of the B2PLYP functional as B2NC-PLYP, where “NC” represents noncovalent interaction. The B2NC-PLYP functional is an improved revision for the B2N-PLYP functional. Similarly, a total of 45 (cHF, cMP2) combinations for the framework of the mPW2-PLYP functional were also assessed with the 6-311+G(2df,2pd) basis set against the S22B database. As shown in Figure 1b, the MADs of the tested (cHF, cMP2) combinations float within about 0.17−0.36 kcal/mol. The MAD of the (cHF = 0.67, cMP2 = 0.49) combination is 0.17 kcal/ mol, and the corresponding MA%D is 4.8%. And hence, we denote the (cHF = 0.67, cMP2 = 0.49) combination for the framework of the mPW2-PLYP functional as mPW2NC-PLYP. 3.2. Basis Set Effects of the B2NC-PLYP and mPW2NCPLYP Functionals. As comparisons, the basis sets of 631+G(d,p), 6-311+G(d,p), 6-311+G(2df,2pd), 6-311++G(3df,3pd), aug-cc-pVDZ, aug-cc-pVTZ, cc-pVTZ, def2TZVPP, ma-TZVPP, def2-QZVP, and ma-QZVP are systematically assessed for the B2NC-PLYP and mPW2NC-PLYP functionals against the S22B database. As listed in Table 2, the 6-311+G(2df,2pd) and aug-cc-pVTZ basis sets are reasonable choices for the B2NC-PLYP and mPW2NC-PLYP functionals to investigate noncovalent interactions. As the basis set utilized in the optimization process of the B2NC-PLYP functional, the 6-311+G(2df,2p) basis set is definitely suitable for the B2NCPLYP functional. Furthermore, according to the performances of the cc-pVTZ and aug-cc-pVTZ basis sets, the corresponding diffuse functions play important roles for these two functionals to study noncovalent interactions. The performances of these two functionals with the def2-QZVP and ma-QZVP basis sets are worse than those with the def2-TZVPP and ma-TZVPP 3176

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

Table 2. Basis Set Effects of the B2NC-PLYP and mPW2NCPLYP Functionals with the S22B Databasea basis set

MD

6-31+G(d,p) 6-311+G(d,p) 6-311+G(2df,2p) 6-311+G(2df,2pd) 6-311++G(3df,3pd) aug-cc-pVDZ aug-cc-pVTZ cc-pVTZ def2-TZVPP ma-TZVPP def2-QZVP ma-QZVP 6-31+G(d,p) 6-311+G(d,p) 6-311+G(2df,2pd) 6-311++G(3df,3pd) aug-cc-pVDZ aug-cc-pVTZ cc-pVTZ def2-TZVPP ma-TZVPP def2-QZVP ma-QZVP

MAD

MA%D

0.42 0.50 0.15 0.21 0.45 0.97 0.23 0.46 0.31 0.29 0.39 0.40

12.5 12.7 5.9 6.8 11.5 22.2 5.7 10.9 9.3 8.3 10.5 10.7

0.37 0.52 0.17 0.30 0.79 0.16 0.51 0.36 0.38 0.44 0.46

10.7 12.0 4.8 8.6 20.4 3.6 11.2 9.1 8.4 10.2 10.4

B2NC-PLYP −0.21 −0.16 −0.01 −0.12 −0.40 −0.97 −0.16 −0.19 0.13 0.25 0.31 0.35 mPW2NC-PLYP −0.09 −0.20 −0.01 −0.27 −0.79 −0.03 −0.21 0.17 0.34 0.38 0.43

a

The MD (kcal/mol), MAD (kcal/mol), and MA%D are employed to evaluate the performances of various basis sets.

Table 3. MADs (kcal/mol) of the B2NC-PLYP and mPW2NC-PLYP Functionals with Several Selected Basis Sets on the Subsets of the S22B and S66 Databasesa basis set (database)

HB

DD

B2NC-PLYP 0.21 0.18 0.30 0.20 0.31 0.17 0.18 0.27 0.89 0.33 0.32 0.46 0.24 0.46 0.24 0.64 0.24 0.64 mPW2NC-PLYP 6-311+G(2df,2pd) (S22B) 0.26 0.19 aug-cc-pVTZ (S22B) 0.18 0.19 aug-cc-pVTZ (S66) 0.16 0.28 cc-pVTZ (S22B) 0.97 0.35 def2-TZVPP (S22B) 0.36 0.52 ma-TZVPP (S22B) 0.35 0.56 def2-QZVP (S22B) 0.28 0.72 ma-QZVP (S22B) 0.27 0.74 6-311+G(2df,2p) (S22B) 6-311+G(2df,2pd) (S22B) aug-cc-pVTZ (S22B) aug-cc-pVTZ (S66) cc-pVTZ (S22B) def2-TZVPP (S22B) ma-TZVPP (S22B) def2-QZVP (S22B) ma-QZVP (S22B)

Figure 1. Contour maps for the MADs (kcal/mol) of the tested (cHF, cMP2) combinations for the frameworks of the B2PLYP (a) and mPW2-PLYP (b) functionals against the S22B database. The 6311+G(2df,2p) (a) and 6-311+G(2df,2pd) (b) basis sets are used for single point energy computations, respectively.

basis sets, respectively. These abnormal results will be discussed in section 3.5. The S22 database contains three different types of complexes. They are hydrogen-bonded (HB) complexes, dispersion-dominated (DD) complexes, and complexes with mixed (MX) influence. The detailed categories used in this work accord with the original ref 66. The performances of the B2NC-PLYP and mPW2NC-PLYP functionals with the 6311+G(2df,2pd), aug-cc-pVTZ, cc-pVTZ, def2-TZVPP, maTZVPP, def2-QZVP, and ma-QZVP basis sets on these three subsets are presented in Table 3. With the 6-311+G(2df,2pd) and aug-cc-pVTZ basis sets, these two functionals perform relatively well on each subset of the S22B database. However, these two functionals perform worse on the HB subset than other two subsets with the cc-pVTZ basis set, whereas they perform worse on the DD subset with the def2-QZVP and maQZVP basis sets. As further validations, the B2NC-PLYP/augcc-pVTZ and mPW2NC-PLYP/aug-cc-pVTZ levels of theory perform very well on the S66 database and corresponding subsets.

MX

total

0.07 0.12 0.21 0.16 0.17 0.13 0.15 0.25 0.28

0.15 0.21 0.23 0.21 0.46 0.31 0.29 0.39 0.40

0.05 0.11 0.15 0.24 0.17 0.21 0.29 0.34

0.17 0.16 0.20 0.51 0.36 0.38 0.44 0.46

a

The corresponding subsets are hydrogen-bonded (HB) complexes, dispersion-dominated (DD) complexes, and complexes with mixed (MX) influence.

3.3. Assessment of the B2NC-PLYP and mPW2NCPLYP functionals with the ADIM6 database and six rare gas dimers. As to the methane dimer in the S22B database, we have found that the absolute percentage deviations of the 3177

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

B2NC-PLYP functional with the 6-311+G(2df,2p), 6-311+G(2df,2pd), and aug-cc-pVTZ basis sets are 53.9%, 54.6%, and 34.1%, respectively. However, the corresponding absolute deviations are relatively small, because the reference total interaction energy of the methane dimer is only −0.527 kcal/ mol in the S22B database. On the basis of the SAPT computation,74 the dispersion component is extremely significant for the total interaction energy of the methane dimer. Similarly, the dispersion energies are very important for the total interaction energies of the dimers in the ADIM6 database. Therefore, we have assessed the B2N-PLYP, B2NCPLYP, and mPW2NC-PLYP functionals with the aug-cc-pVTZ basis set against the ADIM6 database. As listed in Table 4, the Table 4. Performances of the B2N-PLYP, B2NC-PLYP, and mPW2NC-PLYP Functionals with the aug-cc-pVTZ Basis Set against the ADIM6 Databasea functional

MD

MAD

MA%D

B2N-PLYP B2NC-PLYP mPW2NC-PLYP

−0.75 −0.65 −0.57

0.75 0.65 0.57

22.9 19.9 16.4

Figure 2. Interaction energy curves for the Ar···Ar dimer at the B2NCPLYP/aug-cc-pVTZ and mPW2NC-PLYP/aug-cc-pVTZ levels of theory, respectively. The reference interaction energies obtained at the CCSD(T)/aug-cc-pV5Z+(3s3p2d2f1g) level of theory are taken from ref 77.

a

The MD (kcal/mol), MAD (kcal/mol), and MA%D are utilized to evaluate the performances.

performances of the B2N-PLYP, B2NC-PLYP, and mPW2NCPLYP functionals with the aug-cc-pVTZ basis set are less satisfactory. And, thus, we define the noncovalent interactions between alkanes as problematic cases for these functionals. As listed in Table 5, the B2NC-PLYP and mPW2NC-PLYP functionals with the aug-cc-pVTZ basis set are able to locate the

Table 6. MADs (kcal/mol) of Various Computational Levels against the S22 and S66 Databases level of theory S22 B2NC-PLYP/6-311+G(2df,2p) B2NC-PLYP/6-311+G(2df,2pd) B2NC-PLYP/aug-cc-pVTZ mPW2NC-PLYP/6-311+G(2df,2pd) mPW2NC-PLYP/aug-cc-pVTZ B2PLYP-D3/aug-cc-pVTZa ωB97X-D/aug-cc-pVTZa S66 B2NC-PLYP/aug-cc-pVTZ mPW2NC-PLYP/aug-cc-pVTZ PW6B95-D3/def2-QZVPc B2PLYP-D3/def2-QZVPc B2GP-PLYP-D3/def2-QZVPc DSD-BLYP-D3/def2-QZVPc PWPB95-D3/def2-QZVPc MP2/CBSf MP2.5/CBSf SCS-MI-CCSD/CBSf

Table 5. Interaction Potential Wells (kcal/mol) for Six Rare Gas Dimersa rare gas dimer

referenceb

B2NC-PLYP

mPW2NC-PLYP

He2 Ne2 Ar2 He···Ne He···Ar Ne···Ar

−0.02 −0.08 −0.28 −0.04 −0.06 −0.13

−0.01 −0.09 −0.19 −0.03 −0.03 −0.11

−0.07 −0.18 −0.27 −0.12 −0.10 −0.21

a

These interaction potential wells are determined by potential energy scans at the B2NC-PLYP/aug-cc-pVTZ and mPW2NC-PLYP/aug-ccpVTZ levels of theory. bThe reference interaction potential wells obtained at the CCSD(T)/aug-cc-pV5Z+(3s3p2d2f1g) level of theory are taken from ref 77.

MAD 0.15 0.21 0.23 0.17 0.16 0.38 0.33, 0.18b 0.21 0.20 0.23,d 0.26,d 0.26,d 0.21,d 0.19,d 0.45 0.12 0.06

0.18e 0.21e 0.24e 0.23e 0.22e

a

Reference 10. bCP corrected value. cReference 61. dDFT-D3. eDFTD3(BJ). fReference 70.

interaction potential wells for the six tested rare gas dimers. In addition, the interaction energy curves for the Ar···Ar dimer are shown in Figure 2. The interaction energies of the rare gas dimers are rather small, and it is very challenging to compute them accurately with density functionals. 3.4. Comparisons with Previous Benchmark Results. In this section, we compare the performances of various computational levels on the S22 and S66 databases. The corresponding results are mainly cited from refs 10, 61, and 70. In fact, the reference interaction energies for the S22 and S66 databases are slightly different in different revisions. However, these differences will not affect the assessment of various computational methods, especially the density functionals. As shown in Table 6, the performances of the B2NC-PLYP and mPW2NC-PLYP functionals combined with the 6-

311+G(2df,2pd) and aug-cc-pVTZ basis sets are close to those of the computational levels involving the density functionals with dispersion corrections. The B2NC-PLYP and mPW2NC-PLYP functionals are free of dispersion correction, and moreover, the CP correction is not needed with the triple-ζ quality basis sets including adequate polarization and diffuse functions. 3.5. Comments on the B2NC-PLYP and mPW2NCPLYP Functionals. The cHF and cMP2 parameters for the B2NC-PLYP functional are 0.81 and 0.55, respectively. This PT2 implies that the amounts of EHF are very large in the x and Ec exchange−correlation energy Exc. According to our pervious 3178

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

work,56 the large amount of EHF in the Exc will reduce the x exchange-repulsion components for the total interaction energies of the HCOOH···C6H6 complexes. Therefore, we consider that the large amount of EHF contributes to the x performance of the B2NC-PLYP functional on noncovalent interactions. Furthermore, the large amount of EPT2 well c describes the dispersion components for noncovalent interactions and also contributes to the performance of B2NC-PLYP functional, especially for the DD and MX complexes. As to the mPW2NC-PLYP functional, the cHF and cMP2 parameters are 0.67 and 0.49, respectively. Combined with the LYP correlation,98 the mPW exchange99 can reduce the amounts PT2 of EHF compared with the B88 exchange.100 This is x and Ec due to the improved long-range behavior of the mPW exchange.99 As mentioned above, the performances of the B2NC-PLYP and mPW2NC-PLYP functionals with the aug-cc-pVTZ basis set against the ADIM6 database are less satisfactory. Furthermore, some radical−molecule noncovalent interactions may also be problematic cases for these two functionals because of the large amount of EHF x . Various density functionals should be assessed against the open-shell noncovalent interactions in the future.101 Additionally, we have assessed the performances of the B2NC-PLYP and mPW2NC-PLYP functionals with the 6311+G(2df,2p) basis set on other subsets of the GMTKN30 database.40,73 The W4-08,22,40,73 BH76, and BH76RC40,73,102,103 subsets are employed. The corresponding computations using the B2NC-PLYP and mPW2NC-PLYP functionals were performed with the GAMESS93,94 and Firefly93,104 programs, respectively. The MDs and MADs of these two levels against the W4-08, BH76, and BH76RC subsets are listed in Table 7. As a result, the mPW2NC-PLYP

another issue. In this work, the basis set incompleteness is absorbed by the optimized parameters; i.e., the corresponding BSSEs are incorporated in the B2NC-PLYP and mPW2NCPLYP functionals. Therefore, these two functionals perform relatively well with the triple-ζ quality basis sets. And, moreover, the MADs using the nearly saturated def2-QZVP and ma-QZVP basis sets are larger than those with the def2TZVPP and ma-TZVPP basis sets, respectively. In subsequent work, the basis set such as def2-QZVP and/or extrapolation to the complete basis set (CBS) limit26 will be used to address this issue. We have also examined the CP correction for the basis sets of aug-cc-pVDZ and def2-TZVPP with the B2NC-PLYP and mPW2NC-PLYP functionals. As shown in Table 8, the half Table 8. Comparisons of CP and Half CP Corrected Results for the aug-cc-pVDZ and def2-TZVPP Basis Setsa basis set aug-cc-pVDZ (CP) aug-cc-pVDZ (half CP) def2-TZVPP (CP) def2-TZVPP (half CP) aug-cc-pVDZ (CP) aug-cc-pVDZ (half CP) def2-TZVPP (CP) def2-TZVPP (half CP)

B2NC-PLYP MAD

MD

MAD

W4-08 BH76 BH76RC

−7.91 1.79 −0.15

8.03 1.87 1.11

0.03 −1.93 0.27

3.97 2.69 1.57

MA%D

1.07 0.23 0.99 0.59

22.5 5.4 22.0 14.8

1.13 0.24 1.05 0.66

21.8 3.9 21.3 14.4

The MD (kcal/mol), MAD (kcal/mol), and MA%D are used to assess the performances.

CP22,36,108 correction for the aug-cc-pVDZ basis set performs relatively well against the S22B database. As mentioned above, the BSSEs of the triple-ζ quality 6-311+G(2df,2p) and 6311+G(2df,2pd) basis sets have been incorporated in the B2NC-PLYP and mPW2NC-PLYP functionals, respectively. And thus, the CP and half CP corrections degrade the corresponding performances with the def2-TZVPP basis set.

mPW2NC-PLYP

MD

MAD

a

Table 7. Performances of the B2NC-PLYP and mPW2NCPLYP Functionals with the 6-311+G(2df,2p) Basis Set against the W4-08, BH76, and BH76RC Subsets of the GMTKN30 Databasea subset

MD B2NC-PLYP 1.07 0.05 0.99 0.56 mPW2NC-PLYP 1.13 0.17 1.05 0.61

4. CONCLUSIONS Two DHDFs of B2NC-PLYP and mPW2NC-PLYP have been obtained for noncovalent interactions in this work. These two functionals are derived free of dispersion and CP corrections against the S22B database. On the basis of our benchmark computations, we recommend the B2NC-PLYP and mPW2NC-PLYP functionals with the triple-ζ quality basis sets including adequate polarization and diffuse functions to study noncovalent interactions. The B2PLYP-D3 and mPW2PLYP-D3 functionals may be utilized for further validation. However, different exchange and correlation functionals will be tested as the B2NC-PLYP and mPW2NC-PLYP functionals in the future. We hope to reduce the amount of the Fockexchange further and improve the corresponding performance on the ADIM6 database.

a

The MD (in kcal/mol) and MAD (in kcal/mol) are utilized to evaluate the performances.

functional performs better than the B2NC-PLYP functional on the W4-08 subset. On the contrary, the B2NC-PLYP functional performs better on the BH76 and BH76RC subsets. Comparisons with other computational levels are presented in the Supporting Information (Table S2). However, the B2TPLYP,21 B2K-PLYP,21 mPW2K-PLYP,21 and B2GP-PLYP22 functionals are reasonable choices for thermochemistry and/or thermochemical kinetics. Moreover, the performances of the B2NC-PLYP/aug-cc-pVTZ level of theory against the small representative atomization energies (AE6) and barrier heights (BH6) databases105−107 are given in the Supporting Information (Table S3). Both the MAD per bond with the AE6 database and MAD against the BH6 database are close to 1 kcal/mol. The incompleteness of the Pople’s split-valence triple-ζ basis set in the optimization of the cHF and cMP2 parameters is



ASSOCIATED CONTENT

S Supporting Information *

Input files for the B2NC-PLYP and mPW2NC-PLYP functionals; total energies of three examples in the S22B database computed at the B2NC-PLYP/aug-cc-pVTZ level of theory with the GAMESS and ORCA programs (Table S1); MADs of various computational levels against the W4-08, BH76, and 3179

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

(17) Zhao, Y.; Lynch, B. J.; Truhlar, D. G. Multi-Coefficient Extrapolated Density Functional Theory for Thermochemistry and Thermochemical Kinetics. Phys. Chem. Chem. Phys. 2005, 7, 43−52. (18) Grimme, S. Semiempirical Hybrid Density Functional with Perturbative Second-Order Correlation. J. Chem. Phys. 2006, 124, 034108. (19) Schwabe, T.; Grimme, S. Towards Chemical Accuracy for the Thermodynamics of Large Molecules: New Hybrid Density Functionals Including Non-Local Correlation Effects. Phys. Chem. Chem. Phys. 2006, 8, 4398−4401. (20) Schwabe, T.; Grimme, S. Double-Hybrid Density Functionals with Long-Range Dispersion Corrections: Higher Accuracy and Extended Applicability. Phys. Chem. Chem. Phys. 2007, 9, 3397−3406. (21) Tarnopolsky, A.; Karton, A.; Sertchook, R.; Vuzman, D.; Martin, J. M. L. Double-Hybrid Functionals for Thermochemical Kinetics. J. Phys. Chem. A 2008, 112, 3−8. (22) Karton, A.; Tarnopolsky, A.; Lamère, J.-F.; Schatz, G. C.; Martin, J. M. L. Highly Accurate First-Principles Benchmark Data Sets for the Parametrization and Validation of Density Functional and Other Approximate Methods. Derivation of a Robust, Generally Applicable, Double-Hybrid Functional for Thermochemistry and Thermochemical Kinetics. J. Phys. Chem. A 2008, 112, 12868−12886. (23) Sancho-García, J. C.; Pérez-Jiménez, A. J. Assessment of DoubleHybrid Energy Functionals for π-Conjugated Systems. J. Chem. Phys. 2009, 131, 084108. (24) Graham, D. C.; Menon, A. S.; Goerigk, L.; Grimme, S.; Radom, L. Optimization and Basis-Set Dependence of a Restricted-Open-Shell Form of B2-PLYP Double-Hybrid Density Functional Theory. J. Phys. Chem. A 2009, 113, 9861−9873. (25) Benighaus, T.; DiStasio, R. A., Jr.; Lochan, R. C.; Chai, J.-D.; Head-Gordon, M. Semiempirical Double-Hybrid Density Functional with Improved Description of Long-Range Correlation. J. Phys. Chem. A 2008, 112, 2702−2712. (26) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected DoubleHybrid Density Functionals. J. Chem. Phys. 2009, 131, 174105. (27) Zhang, Y.; Xu, X.; Goddard, W. A., III. Doubly Hybrid Density Functional for Accurate Descriptions of Nonbond Interactions, Thermochemistry, and Thermochemical Kinetics. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 4963−4968. (28) Zhang, I. Y.; Xu, X.; Jung, Y.; Goddard, W. A., III. A Fast Doubly Hybrid Density Functional Method Close to Chemical Accuracy Using a Local Opposite Spin Ansatz. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 19896−19900. (29) Zhang, I. Y.; Wu, J.; Xu, X. Extending the Reliability and Applicability of B3LYP. Chem. Commun. 2010, 46, 3057−3070. (30) Zhang, I. Y.; Luo, Y.; Xu, X. XYG3s: Speedup of the XYG3 Fifth-Rung Density Functional with Scaling-All-Correlation Method. J. Chem. Phys. 2010, 132, 194105. (31) Zhang, I. Y.; Luo, Y.; Xu, X. Basis Set Dependence of the Doubly Hybrid XYG3 Functional. J. Chem. Phys. 2010, 133, 104105. (32) Zhang, I. Y.; Xu, X. Doubly Hybrid Density Functional for Accurate Description of Thermochemistry, Thermochemical Kinetics and Nonbonded Interactions. Int. Rev. Phys. Chem. 2011, 30, 115−160. (33) Shamov, G. A.; Schreckenbach, G.; Budzelaar, P. H. M. Stability of Hydrocarbons of the Polyhedrane Family Containing Bridged CH Groups: A Case of Failure of the Colle−Salvetti Correlation Density Functionals. J. Chem. Theory Comput. 2010, 6, 3442−3455. (34) Kozuch, S.; Gruzman, D.; Martin, J. M. L. DSD-BLYP: A General Purpose Double Hybrid Density Functional Including Spin Component Scaling and Dispersion Correction. J. Phys. Chem. C 2010, 114, 20801−20808. (35) Kozuch, S.; Martin, J. M. L. DSD-PBEP86: In Search of the Best Double-Hybrid DFT with Spin-Component Scaled MP2 and Dispersion Corrections. Phys. Chem. Chem. Phys. 2011, 13, 20104− 20107. (36) Kozuch, S.; Martin, J. M. L. Spin-Component-Scaled Double Hybrids: An Extensive Search for the Best Fifth-Rung Functionals Blending DFT and Perturbation Theory. J. Comput. Chem. 2013, 34, 2327−2344.

BH76RC databases (Table S2); and assessment of the B2NCPLYP functional with the aug-cc-pVTZ basis set against the AE6 and BH6 databases (Table S3). This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Feng Yu gratefully acknowledges the Xi’an Technological University for the financial support (No. XAGDXJJ1030). Partial computations were performed at the Supercomputing Center of Qingdao Institute of Bioenergy and Bioprocess Technology.



REFERENCES

(1) Müller-Dethlefs, K.; Hobza, P. Noncovalent Interactions: A Challenge for Experiment and Theory. Chem. Rev. 2000, 100, 143− 167. (2) Hobza, P.; Müller-Dethlefs, K. Non-covalent Interactions: Theory and Experiment; Royal Society of Chemistry: Cambridge, U.K., 2010. (3) Riley, K. E.; Pitoňaḱ , M.; Jurečka, P.; Hobza, P. Stabilization and Structure Calculations for Noncovalent Interactions in Extended Molecular Systems Based on Wave Function and Density Functional Theories. Chem. Rev. 2010, 110, 5023−5063. (4) Riley, K. E.; Hobza, P. Noncovalent Interactions in Biochemistry. WIREs Comput. Mol. Sci. 2011, 1, 3−17. (5) Hobza, P. Calculations on Noncovalent Interactions and Databases of Benchmark Interaction Energies. Acc. Chem. Res. 2012, 45, 663−672. (6) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157−167. (7) Zhao, Y.; Truhlar, D. G. Applications and Validations of the Minnesota Density Functionals. Chem. Phys. Lett. 2011, 502, 1−13. (8) Peverati, R.; Truhlar, D. G. The Quest for a Universal Density Functional: The Accuracy of Density Functionals across a Broad Spectrum of Databases in Chemistry and Physics. arXiv:1212.0944 [physics.chem-ph]. (9) Grimme, S. Density Functional Theory with London Dispersion Corrections. WIREs Comput. Mol. Sci. 2011, 1, 211−228. (10) Burns, L. A.; Vázquez-Mayagoitia, Á .; Sumpter, B. G.; Sherrill, C. D. Density-Functional Approaches to Noncovalent Interactions: A Comparison of Dispersion Corrections (DFT-D), Exchange-Hole Dipole Moment (XDM) Theory, and Specialized Functionals. J. Chem. Phys. 2011, 134, 084107. (11) Hohenstein, E. G.; Sherrill, C. D. Wavefunction Methods for Noncovalent Interactions. WIREs Comput. Mol. Sci. 2012, 2, 304−326. (12) Klimeš, J.; Michaelides, A. Advances and Challenges in Treating van der Waals Dispersion Forces in Density Functional Theory. J. Chem. Phys. 2012, 137, 120901. (13) Podeszwa, R.; Szalewicz, K. Density Functional Theory Overcomes the Failure of Predicting Intermolecular Interaction Energies. J. Chem. Phys. 2012, 136, 161102. (14) Johnson, E. R.; Mackie, I. D.; DiLabio, G. A. Dispersion Interactions in Density-Functional Theory. J. Phys. Org. Chem. 2009, 22, 1127−1135. (15) Swart, M.; Solà, M.; Bickelhaupt, F. M. Inter- and Intramolecular Dispersion Interactions. J. Comput. Chem. 2011, 32, 1117− 1127. (16) Zhao, Y.; Lynch, B. J.; Truhlar, D. G. Doubly Hybrid Meta DFT: New Multi-Coefficient Correlation and Density Functional Methods for Thermochemistry and Thermochemical Kinetics. J. Phys. Chem. A 2004, 108, 4786−4791. 3180

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

(37) Karton, A.; Martin, J. M. L. Basis Set Convergence of Explicitly Correlated Double-Hybrid Density Functional Theory Calculations. J. Chem. Phys. 2011, 135, 144119. (38) Sharkas, K.; Toulouse, J.; Savin, A. Double-Hybrid DensityFunctional Theory Made Rigorous. J. Chem. Phys. 2011, 134, 064113. (39) Souvi, S. O.; Sharkas, K.; Toulouse, J. Double-Hybrid DensityFunctional Theory with Meta-Generalized-Gradient Approximations. J. Chem. Phys. 2014, 140, 084107. (40) Goerigk, L.; Grimme, S. Efficient and Accurate Double-HybridMeta-GGA Density FunctionalsEvaluation with the Extended GMTKN30 Database for General Main Group Thermochemistry, Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2011, 7, 291−309. (41) Chan, B.; Radom, L. Obtaining Good Performance with Tripleζ-Type Basis Sets in Double-Hybrid Density Functional Theory Procedures. J. Chem. Theory Comput. 2011, 7, 2852−2863. (42) Brémond, E.; Adamo, C. Seeking for Parameter-Free DoubleHybrid Functionals: The PBE0-DH Model. J. Chem. Phys. 2011, 135, 024106. (43) Fromager, E. Rigorous Formulation of Two-Parameter DoubleHybrid Density-Functionals. J. Chem. Phys. 2011, 135, 244106. (44) 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. (45) Mohajeri, A.; Alipour, M. B2-PPW91: A Promising DoubleHybrid Density Functional for the Electric Response Properties. J. Chem. Phys. 2012, 136, 124111. (46) Zhang, I. Y.; Su, N. Q.; Brémond, É. A. G.; Adamo, C.; Xu, X. Doubly Hybrid Density Functional xDH-PBE0 from a Parameter-Free Global Hybrid Model PBE0. J. Chem. Phys. 2012, 136, 174103. (47) Zhang, I. Y.; Xu, X. XYG3 and XYGJ-OS Performances for Noncovalent Binding Energies Relevant to Biomolecular Structures. Phys. Chem. Chem. Phys. 2012, 14, 12554−12570. (48) Bousquet, D.; Brémond, E.; Sancho-García, J. C.; Ciofini, I.; Adamo, C. Is There Still Room for Parameter Free Double Hybrids? Performances of PBE0-DH and B2PLYP over Extended Benchmark Sets. J. Chem. Theory Comput. 2013, 9, 3444−3452. (49) Cornaton, Y.; Franck, O.; Teale, A. M.; Fromager, E. Analysis of Double-Hybrid Density Functionals along the Adiabatic Connection. Mol. Phys. 2013, 111, 1275−1294. (50) Peverati, R.; Head-Gordon, M. Orbital Optimized DoubleHybrid Density Functionals. J. Chem. Phys. 2013, 139, 024110. (51) Cornaton, Y.; Fromager, E. Double Hybrid Density-Functional Theory Using the Coulomb-Attenuating Method. Int. J. Quantum Chem. 2014, DOI: 10.1002/qua.24682. (52) Ji, H.; Shao, Y.; Goddard, W. A.; Jung, Y. Analytic Derivatives of Quartic-Scaling Doubly Hybrid XYGJ-OS Functional: Theory, Implementation, and Benchmark Comparison with M06-2X and MP2 Geometries for Nonbonded Complexes. J. Chem. Theory Comput. 2013, 9, 1971−1976. (53) Sancho-García, J. C.; Adamo, C. Double-Hybrid Density Functionals: Merging Wavefunction and Density Approaches to Get the Best of Both Worlds. Phys. Chem. Chem. Phys. 2013, 15, 14581− 14594. (54) Zhang, I. Y.; Xu, X. Reaching a Uniform Accuracy for Complex Molecular Systems: Long-Range-Corrected XYG3 Doubly Hybrid Density Functional. J. Phys. Chem. Lett. 2013, 4, 1669−1675. (55) Aragó, J.; Ortí, E.; Sancho-García, J. C. Nonlocal van der Waals Approach Merged with Double-Hybrid Density Functionals: Toward the Accurate Treatment of Noncovalent Interactions. J. Chem. Theory Comput. 2013, 9, 3437−3443. (56) Yu, F. Intermolecular Interactions of Formic Acid with Benzene: Energy Decomposition Analyses with Ab Initio MP2 and DoubleHybrid Density Functional Computations. Int. J. Quantum Chem. 2013, 113, 2355−2360. (57) Su, N. Q.; Xu, X. Construction of a Parameter-Free Doubly Hybrid Density Functional from Adiabatic Connection. J. Chem. Phys. 2014, 140, 18A512.

(58) Grimme, S. Accurate Description of van der Waals Complexes by Density Functional Theory Including Empirical Corrections. J. Comput. Chem. 2004, 25, 1463−1473. (59) Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787−1799. (60) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (61) Goerigk, L.; Kruse, H.; Grimme, S. Benchmarking Density Functional Methods against the S66 and S66 × 8 Datasets for NonCovalent Interactions. ChemPhysChem 2011, 12, 3421−3433. (62) Su, P.; Li, H. Energy Decomposition Analysis of Covalent Bonds and Intermolecular Interactions. J. Chem. Phys. 2009, 131, 014102. (63) Zhao, Y.; Ng, H. T.; Hanson, E. Benchmark Data for Noncovalent Interactions in HCOOH···Benzene Complexes and Their Use for Validation of Density Functionals. J. Chem. Theory Comput. 2009, 5, 2726−2733. (64) Zhao, Y.; Truhlar, D. G. Design of Density Functionals That Are Broadly Accurate for Thermochemistry, Thermochemical Kinetics, and Nonbonded Interactions. J. Phys. Chem. A 2005, 109, 5656−5667. (65) Zhao, Y.; Truhlar, D. G. Benchmark Databases for Nonbonded Interactions and Their Use to Test Density Functional Theory. J. Chem. Theory Comput. 2005, 1, 415−432. (66) Jurečka, P.; Šponer, J.; Č erný, J.; Hobza, P. Benchmark Database of Accurate (MP2 and CCSD(T) Complete Basis Set Limit) Interaction Energies of Small Model Complexes, DNA Base Pairs, and Amino Acid Pairs. Phys. Chem. Chem. Phys. 2006, 8, 1985−1993. (67) Takatani, T.; Hohenstein, E. G.; Malagoli, M.; Marshall, M. S.; Sherrill, C. D. Basis Set Consistent Revision of the S22 Test Set of Noncovalent Interaction Energies. J. Chem. Phys. 2010, 132, 144104. (68) Podeszwa, R.; Patkowski, K.; Szalewicz, K. Improved Interaction Energy Benchmarks for Dimers of Biological Relevance. Phys. Chem. Chem. Phys. 2010, 12, 5974−5979. (69) Marshall, M. S.; Burns, L. A.; Sherrill, C. D. Basis Set CCSD(T) : Best Convergence of the Coupled-Cluster Correction, δMP2 Practices for Benchmarking Non-Covalent Interactions and the Attendant Revision of the S22, NBC10, HBC6, and HSG Databases. J. Chem. Phys. 2011, 135, 194102. (70) Ř ezác,̌ J.; Riley, K. E.; Hobza, P. S66: A Well-balanced Database of Benchmark Interaction Energies Relevant to Biomolecular Structures. J. Chem. Theory Comput. 2011, 7, 2427−2438. (71) Ř ezác,̌ J.; Riley, K. E.; Hobza, P. Extensions of the S66 Data Set: More Accurate Interaction Energies and Angular-Displaced Nonequilibrium Geometries. J. Chem. Theory Comput. 2011, 7, 3466−3470. (72) Tsuzuki, S.; Honda, K.; Uchimaru, T.; Mikami, M. Estimated MP2 and CCSD(T) Interaction Energies of n-Alkane Dimers at the Basis Set Limit: Comparison of the Methods of Helgaker et al. and Feller. J. Chem. Phys. 2006, 124, 114304. (73) Goerigk, L.; Grimme, S. A Thorough Benchmark of Density Functional Methods for General Main Group Thermochemistry, Kinetics, and Noncovalent Interactions. Phys. Chem. Chem. Phys. 2011, 13, 6670−6688. (74) Flick, J. C.; Kosenkov, D.; Hohenstein, E. G.; Sherrill, C. D.; Slipchenko, L. V. Accurate Prediction of Noncovalent Interaction Energies with the Effective Fragment Potential Method: Comparison of Energy Components to Symmetry-Adapted Perturbation Theory for the S22 Test Set. J. Chem. Theory Comput. 2012, 8, 2835−2843. (75) Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van der Waals Complexes. Chem. Rev. 1994, 94, 1887−1930. (76) Hohenstein, E. G.; Sherrill, C. D. Density Fitting of Intramonomer Correlation Effects in Symmetry-Adapted Perturbation Theory. J. Chem. Phys. 2010, 133, 014101. (77) Cybulski, S. M.; Toczyłowski, R. R. Ground State Potential Energy Curves for He2, Ne2, Ar2, He−Ne, He−Ar, and Ne−Ar: A Coupled-Cluster Study. J. Chem. Phys. 1999, 111, 10520−10528. 3181

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

The Journal of Physical Chemistry A

Article

(98) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785−789. (99) Adamo, C.; Barone, V. Exchange Functionals with Improved Long-Range Behavior and Adiabatic Connection Methods without Adjustable Parameters: The mPW and mPW1PW Models. J. Chem. Phys. 1998, 108, 664−675. (100) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098−3100. (101) Soydaş, E.; Bozkaya, U. Accurate Open-Shell Noncovalent Interaction Energies from the Orbital-Optimized Møller−Plesset Perturbation Theory: Achieving CCSD Quality at the MP2 Level by Orbital Optimization. J. Chem. Theory Comput. 2013, 9, 4679−4683. (102) Zhao, Y.; Lynch, B. J.; Truhlar, D. G. Development and Assessment of a New Hybrid Density Functional Model for Thermochemical Kinetics. J. Phys. Chem. A 2004, 108, 2715−2719. (103) Zhao, Y.; González-García, N.; Truhlar, D. G. Benchmark Database of Barrier Heights for Heavy Atom Transfer, Nucleophilic Substitution, Association, and Unimolecular Reactions and Its Use to Test Theoretical Methods. J. Phys. Chem. A 2005, 109, 2012−2018. (104) Granovsky, A. A. Firefly version 8.0, http://classic.chem.msu. su/gran/firefly/index.html (accessed April 2014). (105) Lynch, B. J.; Truhlar, D. G. Small Representative Benchmarks for Thermochemical Calculations. J. Phys. Chem. A 2003, 107, 8996− 8999. (106) Peverati, R.; Truhlar, D. G. A Global Hybrid Generalized Gradient Approximation to the Exchange-Correlation Functional that Satisfies the Second-Order Density-Gradient Constraint and Has Broad Applicability in Chemistry. J. Chem. Phys. 2011, 135, 191102. (107) Zheng, J.; Zhao, Y.; Truhlar, D. G. The DBH24/08 Database and Its Use to Assess Electronic Structure Model Chemistries for Chemical Reaction Barrier Heights. J. Chem. Theory Comput. 2009, 5, 808−821. (108) Kim, K. S.; Tarakeshwar, P.; Lee, J. Y. Molecular Clusters of πSystems: Theoretical Studies of Structures, Spectra, and Origin of Interaction Energies. Chem. Rev. 2000, 100, 4145−4186.

(78) Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724−728. (79) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257−2261. (80) Francl, M. M.; Pietro, W. J.; Hehre, W. J.; Binkley, J. S.; Gordon, M. S.; DeFrees, D. J.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XXIII. A Polarization-Type Basis Set for Second-Row Elements. J. Chem. Phys. 1982, 77, 3654−3665. (81) Hariharan, P. C.; Pople, J. A. The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213−222. (82) Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. v. R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-Row Elements, Li−F. J. Comput. Chem. 1983, 4, 294−301. (83) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. SelfConsistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650−654. (84) Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-Consistent Molecular Orbital Methods 25. Supplementary Functions for Gaussian Basis Sets. J. Chem. Phys. 1984, 80, 3265−3269. (85) Dunning, T. H., Jr. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007−1023. (86) Kendall, R. A.; Dunning, T. H., Jr.; Harrison, R. J. Electron Affinities of the First-Row Atoms Revisited. Systematic Basis Sets and Wave Functions. J. Chem. Phys. 1992, 96, 6796−6806. (87) Woon, D. E.; Dunning, T. H., Jr. Gaussian Basis Sets for Use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon. J. Chem. Phys. 1993, 98, 1358−1371. (88) Woon, D. E.; Dunning, T. H., Jr. Gaussian Basis Sets for Use in Correlated Molecular Calculations. IV. Calculation of Static Electrical Response Properties. J. Chem. Phys. 1994, 100, 2975−2988. (89) 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. (90) Zheng, J.; Xu, X.; Truhlar, D. G. Minimally Augmented Karlsruhe Basis Sets. Theor. Chem. Acc. 2011, 128, 295−305. (91) Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553−566. (92) Lynch, B. J.; Zhao, Y.; Truhlar, D. G. Effectiveness of Diffuse Basis Functions for Calculating Relative Energies by Density Functional Theory. J. Phys. Chem. A 2003, 107, 1384−1388. (93) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S. J.; Windus, T. L.; Dupuis, M.; Montgomery, J. A. General Atomic and Molecular Electronic Structure System. J. Comput. Chem. 1993, 14, 1347−1363. (94) Gordon, M. S.; Schmidt, M. W. Advances in Electronic Structure Theory: GAMESS a Decade Later. In Theory and Applications of Computational Chemistry, the First Forty Years; Dykstra, C. E., Frenking, G., Kim, K. S., Scuseria, G. E., Ed.; Elsevier: Amsterdam, 2005; pp 1167−1189. (95) Neese, F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73−78. (96) Lebedev, V. I.; Laikov, D. N. A Quadrature Formula for the Sphere of the 131st Algebraic Order of Accuracy. Dokl. Math. 1999, 59, 477−481. (97) Yu, F. Assessment of Ab Initio MP2 and Density Functionals for Characterizing the Potential Energy Profiles of the SN2 Reactions at N Center. J. Comput. Chem. 2012, 33, 1347−1352. 3182

dx.doi.org/10.1021/jp5005506 | J. Phys. Chem. A 2014, 118, 3175−3182

Double-hybrid density functionals free of dispersion and counterpoise corrections for non-covalent interactions.

We have optimized two double-hybrid density functionals (DHDFs) within the frameworks of B2PLYP and mPW2-PLYP against the S22B database. These two fun...
514KB Sizes 0 Downloads 4 Views