Molecules 26 04072 v2
Molecules 26 04072 v2
Article
Hydricity of 3d Transition Metal Complexes from Density
Functional Theory: A Benchmarking Study
Alister S. Goodfellow and Michael Bühl *
Abstract: A range of modern density functional theory (DFT) functionals have been benchmarked
against experimentally determined metal hydride bond strengths for three first-row TM hydride com-
plexes. Geometries were found to be produced sufficiently accurately with RI-BP86-D3(PCM)/def2-
SVP and further single-point calculations with PBE0-D3(PCM)/def2-TZVP were found to reproduce
the experimental hydricity accurately, with a mean absolute deviation of 1.4 kcal/mol for the com-
plexes studied.
1. Introduction
At the forefront of modern chemistry is sustainability, with a drive towards ‘greener’
processes and development. Catalysis has always been a tool used to reduce energetic
Citation: Goodfellow, A.S.; Bühl, M.
costs and to promote specific reactions, improving selectivity and the efficacy of various
Hydricity of 3d Transition Metal
transformations. Traditionally, homogeneous catalysts have been based upon expensive
Complexes from Density Functional
and unsustainable metals such as platinum, palladium and rhodium [1] and while these 4d
Theory: A Benchmarking Study.
and 5d transition metals (TMs) have been very successful in this application, their future use
Molecules 2021, 26, 4072. http://
is limited by concerns over sustainability and price. The scarcity of the metals has resulted
doi.org/10.3390/molecules26134072
in a high economic and social cost in their extraction. The metals themselves are also toxic,
Academic Editors: William T. A.
which may lead to issues of contamination in extraction, chemical transformations, or in the
Harrison, R. Alan Aitken and Paul application of these catalysts in industrial processes. To alleviate these issues, development
Waddell has moved towards the use of 3d TMs, which are largely more abundant, less toxic and
more sustainable [2–5].
Received: 3 June 2021 First-row TM catalysts have undergone huge development in the past few years and
Accepted: 30 June 2021 are being optimised to ultimately become competitive against their unsustainable heavy
Published: 3 July 2021 metal congeners. Iron and manganese are attractive due to their low toxicity and high
abundance, and work on catalytic hydrogenation reactions using these metals is very
Publisher’s Note: MDPI stays neutral topical (see Scheme 1). For a review on the development of manganese based pincer
with regard to jurisdictional claims in catalysts, see Garbe et al. [6].
published maps and institutional affil- Density functional theory is a powerful tool for the elucidation of reaction mechanisms
iations. and for the rational design of these catalytic systems. To accurately study such systems,
the DFT functional used must be suitable for the system, accurately predicting bond
energies that are crucial to the functionality of the molecule. 3d TMs are notoriously
tricky to study with DFT, especially bare TMs where the electronic structure is not always
Copyright: © 2021 by the authors. predicted correctly [7,8]. For investigation in the field of homogeneous catalysis, especially
Licensee MDPI, Basel, Switzerland. in hydrogenation reactions as shown in Scheme 1, the metal hydride bond will be of
This article is an open access article prime importance for the catalytic activity, and the functional used must be accurate in the
distributed under the terms and description of these bond strengths.
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
Scheme 1. A representation of the catalytic application of Mn-based systems from the group of
Beller [9].
A number of papers have used DFT computations to examine the reactivity of these
catalytic systems and a variety of methodology has been employed (Table 1). The choice
of methodology may be based upon previous work in related catalysis, perhaps 4d metal
systems, or simply by using methods that have already been applied in the literature.
For example, early computations from the Beller group used B3PW91, based upon the
performance of this hybrid functional in previous work (including benchmarking studies)
across 3d, 4d and 5d transition metal complexes though more recent work from this group
has used the hybrid PBE0 functional [9–15].
An alternate approach was adopted by Ge et al. [16], where the methodology was
selected based upon the testing of a range of functionals for a multi-step transformation.
M06 was selected for their work as the computed barrier heights were found to be closest to
the average value across the functionals tested. A benchmarking study was performed by
Gamez et al. [17] on a number of high-spin Mn-based complexes using a range of different
functionals. It was found that pure functionals (i.e., non-hybrid) favoured low-spin forms
of these Mn complexes, with weak-field nitrogen and oxygen based ligands, experimentally
known to be high-spin complexes. B3LYP was found to produce poor geometries and PBE0
was recommended for use on these systems. Dispersion corrections were also found to be
of importance in optimisation. The accurate description of spin states is important in the
study of 3d metal complexes, where the energetic splitting of metal-based 3d orbitals may
be small and multi-state reactivity may be possible.
Table 1. A summary of the literature methodology used for DFT calculations on 3d metal catalytic systems.
B3PW91/TZVP [9,10,15]
PBE0/def2-TZVP [17]
Table 1. Cont.
B3PW91/TZVP [22]
PBE0/6-311+G(d,p)// [26]
PBE0/6-31G(d,p)
Table 1. Cont.
a For simplicity, the use of any pseudopotential has been indicated with ECP. b M denotes that there is implicit dispersion included, due
to the parametrisation of the Minnesota family of functionals. GD is used to indicate the inclusion of Grimme’s empirical dispersion
correction, with BJ denoting the use of Becke-Johnson dampening. c ‘sp’ and ‘opt’ denote the use of dispersion corrections or solvation
models in either the single-point or geometry optimisation calculation, respectively.
We now present a validation study on the performance of these and other functionals
for the calculation of hydricity for a few select complexes. Hydricity is the heterolytic
bond dissociation energy of a metal hydride to a metal cation and hydride (Equation (1)),
predicted to be an important property for the study of hydrogenation catalysis.
−−
MH )−− M+ + H−
* ◦
∆GH − (1)
A sizeable amount of ∆GH ◦ data is available from experiment [39], the ultimate
−
benchmark. Arguably, this quantity is key in determining the ease of elementary steps
such as transfer of the hydride to an electrophilic substrate (cf. upper arrow in Scheme 1).
As it turns out, there is a notable variability in the performance of different functionals
in their ability to reproduce these hydricity values, and some clear recommendations for
functionals to choose can be made.
2. Methodology
2.1. Computational Details
Geometry optimisations were performed primarily at the GGA level of DFT using the
BP86 functional, employing a variety of basis sets from the redefinitions of the Ahlrichs
family of bases [40–43]. Structures obtained by X-ray crystallography for closely related
systems were used as starting points in geometry optimisations [44–46]. This is akin to
the methodology of Neale [47] used in the investigation of Fe(II) and Co(III) catalysis
and of Bühl and Kabrede [48] in the benchmarking of geometries for a wide range of
first-row transition metal complexes. Def2-TZVP was used for the investigations of proton
and metal cation solvation and throughout further benchmarking, a variety of basis from
this family were used; def2-SVP, with def2-TZVP describing the metal atom of Fe or
Co, and the addition of a diffuse function on the hydridic hydrogen, taking 1/3 of the
exponent from the diffuse ‘s’ function in the def2-TZVP basis for hydrogen. Dispersion
Molecules 2021, 26, 4072 5 of 13
MH + H+ −
)−
−− M + + H2
* ∆G1◦ (2)
H2 −
)−
−− H+ + H−
* ◦
∆GH
2
= 76.0 kcal/mol [39] (3)
Molecules 2021, 26, 4072 6 of 13
Table 2. Three 3d transition metal complexes with experimentally determined thermodynamic hydricity values (∆GH− ).
aHydrogen atoms have been removed for clarity. The viewing angle is looking towards the vacant site on the metal centre where the
hydridic hydrogen atom was residing previously, emphasising the accessibility of the now vacant site by a solvent molecule. Electron
density (isodensity value of 0.001 a.u.) mapped with electrostatic potential [colour-coded between 0.00 a.u. (red) and +0.35 a.u. (blue)].
as such, the reaction shown in Equation (2) may be better represented by the reaction in
Equation (5), where both cations are solvated by explicit solvent molecules.
◦
[M(NCMe]+ −
)−
−− M+ + MeCN
* ∆GBDE (4)
MH + [H(NCMe)2 ]+ −
)−
−− [M(NCMe)]+ + MeCN + H2
* ∆G2◦ (5)
FeCp(CO)2 H + [H(NCMe)2 ]+ − )−
−− [FeCp(CO)2 (NCMe)]+ + MeCN + H2
* (6)
Co(P4 N2 )H + [H(NCMe)2 ] )−− [Co(P4 N2 )]+ + 2 MeCN + H2
+ −−*
(7)
Co(dppe)2 H + [H(NCMe)2 ]+ −)−
−− [Co(dppe)2 ]+ + 2 MeCN + H2
* (8)
1
MAE = Σ| Ecalc − Eexp | (9)
n
Clearly, the Minnesota functionals of M06-2X and M06-HF, both highly parameterised
for main group chemistry, are poor in their description of these 3d TM complexes. The
lowest MAEs were obtained from RI-BP86, PBE0 and revPBE0 (1.7, 1.3 and 0.9 kcal/mol).
The percentage of HF exchange has been shown to be important by Moltved et al. [83] and
this effect can be seen in the comparison of PBE0 with PBE0-1/3, with 25% and 33% HF
exchange, respectively. Both forms of the hybrid PBE functional performed well in terms of
overall performance, yet PBE0-1/3 had over twice the MAE of PBE0 (1.3 and 2.9 kcal/mol,
respectively).
Based purely upon the MAEs obtained here, revPBE0 would be the functional of
choice; however, we recommend to use PBE0 (based upon an improved performance
over revPBE0 on geometries produced with the smaller def2-SVP basis, MAEs of 1.4 and
Molecules 2021, 26, 4072 8 of 13
1.8 kcal/mol, see ESI Figure S7 and Section 3.3.2), which has been shown to perform well
in many previous studies of TM systems [17,85] and in the differentiation of spin state
energetics [17].
Figure 1. MAE in the calculation of hydricity across each of the three complexes with a range of
single-point functionals. Performed with implicit solvation (PCMMeCN ) and a def2-TZVP basis on
optimised geometries from RI-BP86-D3(PCMMeCN )/def2-TZVP.
geometry optimisation, for Co(dppe)2 H of approximately 12.5 days compared to 2 days for
def2-TZVP and def2-SVP basis sets, respectively). Additionally, there was little change in
the optimised geometry between the use of def2-SVP and def2-TZVP, with central metal lig-
and bond distances reproduced to within 0.006 Å (see Table S3 in the ESI for a comparison
to X-Ray crystal structures from Ciancanelli et al. [81] and Ariyaratne et al. [86]).
Theoretically, for the most accurate description of the system, both dispersion and
solvation should be considered and indeed we found that the inclusion of both minimised
the MAEs in the calculation of hydricity across the three complexes (Figure S8 in the ESI).
4. Conclusions
We have accurately reproduced experimentally measured values of hydricity for three
3d TM complexes. While a mixture of functionals have been used in the literature for
studies on 3d metal homogeneous catalysis, we propose a methodology that has been
shown to accurately reproduce a key M−H bond strength, central to the reactivity of
these compounds.
While low on the Jacob’s ladder of functionals, the pure GGA, BP86, has been shown
to produce accurate energetics for the hydricity of 3d TM hydrides. The hybrid functional,
PBE0 has also been shown to perform well and is recommended for energy calculations
over BP86 due to the improved ability to more reliably differentiate between spin states of
3d TM complexes [17]. The lowest mean absolute errors were found with the inclusion of
both dispersion corrections and implicit solvation.
A double-ζ basis, def2-SVP, was used in geometry optimisation with the RI-BP86-
D3(PCM) and led to a MAE of 1.4 kcal/mol after evaluation of subsequent single-point
at the level of PBE0-D3(PCM)/def2-TZVP. A larger triple-ζ basis, def2-TZVP, used at the
stage of the geometry optimisation led to a lower MAE of 1.3 kcal/mol which, was not
shown to offer any significant improvement for the additional cost.
For a balance between accuracy and expense, we recommend the methodology of
PBE0-D3(PCM)/def2-TZVP//RI-BP86-D3(PCM)/def2-SVP for use on systems involving
3d TM hydride complexes.
Supplementary Materials: The following are available online at, additional computational details
and graphical and tabular material.
Author Contributions: Conceptualisation, M.B.; methodology, A.S.G. and M.B.; investigation, A.S.G.;
writing—original draft, A.S.G.; Supervision, M.B. All authors have read and agreed to the published
version of the manuscript.
Funding: This research received no external funding. The publication of this work received support
from the St Andrews Institutional Open Access Fund.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The research data supporting this publication can be accessed at
https://doi.org/10.17630/7cf7b7e9-3edc-4ad4-b016-3b9b5548f9ac (accessed on 2 July 2021).
Acknowledgments: We thank EastCHEM and the School of Chemistry for support through the EaSI-
CAT program. Calculations were performed on a local compute cluster maintained by H. Früchtl.
Conflicts of Interest: The authors declare no conflict of interest.
Molecules 2021, 26, 4072 10 of 13
References
1. Herrmann, W.A.; Cornils, B. Organometallic Homogeneous Catalysis - Quo vadis? Angew. Chem. (Int. Ed. Engl.) 1997,
36, 1048–1067. [CrossRef]
2. Reed-Berendt, B.G.; Polidano, K.; Morrill, L.C. Recent advances in homogeneous borrowing hydrogen catalysis using earth-
abundant first row transition metals. Org. Biomol. Chem. 2019, 17, 1595–1607. [CrossRef]
3. Carney, J.R.; Dillon, B.R.; Thomas, S.P. Recent Advances of Manganese Catalysis for Organic Synthesis. Eur. J. Org. Chem. 2016,
2016, 3912–3929. [CrossRef]
4. Mukherjee, A.; Milstein, D. Homogeneous Catalysis by Cobalt and Manganese Pincer Complexes. ACS Catal. 2018, 8, 11435–11469.
[CrossRef]
5. Bender, T.A.; Dabrowski, J.A.; Gagné, M.R. Homogeneous catalysis for the production of low-volume, high-value chemicals from
biomass. Nat. Rev. Chem. 2018, 2, 35–46. [CrossRef]
6. Garbe, M.; Junge, K.; Beller, M. Homogeneous Catalysis by Manganese-Based Pincer Complexes. Eur. J. Org. Chem. 2017,
2017, 4344–4362. [CrossRef]
7. Cramer, C.J.; Truhlar, D.G. Density functional theory for transition metals and transition metal chemistry. Phys. Chem. Chem.
Phys. 2009, 11, 10757–10816. [CrossRef]
8. Vogiatzis, K.D.; Polynski, M.V.; Kirkland, J.K.; Townsend, J.; Hashemi, A.; Liu, C.; Pidko, E.A. Computational Approach to
Molecular Catalysis by 3d Transition Metals: Challenges and Opportunities. Chem. Rev. 2019, 119, 2453–2523. [CrossRef]
[PubMed]
9. Elangovan, S.; Topf, C.; Fischer, S.; Jiao, H.; Spannenberg, A.; Baumann, W.; Ludwig, R.; Junge, K.; Beller, M. Selective Catalytic
Hydrogenations of Nitriles, Ketones, and Aldehydes by Well-Defined Manganese Pincer Complexes. J. Am. Chem. Soc. 2016,
138, 8809–8814. [CrossRef]
10. Wei, Z.; De Aguirre, A.; Junge, K.; Beller, M.; Jiao, H. Exploring the mechanisms of aqueous methanol dehydrogenation catalyzed
by defined PNP Mn and Re pincer complexes under base-free as well as strong base conditions. Catal. Sci. Technol. 2018,
8, 3649–3665. [CrossRef]
11. Junge, K.; Wendt, B.; Jiao, H.; Beller, M. Iridium-Catalyzed Hydrogenation of Carboxylic Acid Esters. ChemCatChem 2014,
6, 2810–2814. [CrossRef]
12. Wei, Z.; Junge, K.; Beller, M.; Jiao, H. Hydrogenation of phenyl-substituted CN, CN, CC, CC and CO functional groups by Cr, Mo
and W PNP pincer complexes–A DFT study. Catal. Sci. Technol. 2017, 7, 2298–2307. [CrossRef]
13. Alberico, E.; Lennox, A.J.; Vogt, L.K.; Jiao, H.; Baumann, W.; Drexler, H.J.; Nielsen, M.; Spannenberg, A.; Checinski, M.P.; Junge, H.;
et al. Unravelling the Mechanism of Basic Aqueous Methanol Dehydrogenation Catalyzed by Ru-PNP Pincer Complexes. J. Am.
Chem. Soc. 2016, 138, 14890–14904. [CrossRef]
14. Jiao, H.; Junge, K.; Alberico, E.; Beller, M. A Comparative Computationally Study About the Defined M(II) Pincer Hydrogenation
Catalysts (M = Fe, Ru, Os). J. Comput. Chem. 2016, 37, 168–176. [CrossRef] [PubMed]
15. Elangovan, S.; Garbe, M.; Jiao, H.; Spannenberg, A.; Junge, K.; Beller, M. Hydrogenation of Esters to Alcohols Catalyzed by
Defined Manganese Pincer Complexes. Angew. Chem. 2016, 128, 15590–15594. [CrossRef]
16. Ge, H.; Chen, X.; Yang, X. A mechanistic study and computational prediction of iron, cobalt and manganese cyclopentadienone
complexes for hydrogenation of carbon dioxide. Chem. Commun. 2016, 52, 12422–12425. [CrossRef]
17. Gámez, J.A.; Hölscher, M.; Leitner, W. On the applicability of density functional theory to manganese-based complexes with
catalytic activity toward water oxidation. J. Comput. Chem. 2017, 38, 1747–1751. [CrossRef] [PubMed]
18. Qian, F.; Chen, X.; Yang, X. DFT and AIMD prediction of a SNS manganese pincer complex for hydrogenation of acetophenone.
Chem. Phys. Lett. 2019, 714, 37–44. [CrossRef]
19. Chen, X.; Ge, H.; Yang, X. Newly designed manganese and cobalt complexes with pendant amines for the hydrogenation of CO2
to methanol: A DFT study. Catal. Sci. Technol. 2017, 7, 348–355. [CrossRef]
20. Passera, A.; Mezzetti, A. Mn(I) and Fe(II)/PN(H)P Catalysts for the Hydrogenation of Ketones: A Comparison by Experiment
and Calculation. Adv. Synth. Catal. 2019, 361, 4691–4706. [CrossRef]
21. Zubar, V.; Lebedev, Y.; Azofra, L.M.; Cavallo, L.; El-Sepelgy, O.; Rueping, M. Hydrogenation of CO2 -Derived Carbonates
and Polycarbonates to Methanol and Diols by Metal–Ligand Cooperative Manganese Catalysis. Angew. Chem. Int. Ed. 2018,
57, 13439–13443. [CrossRef]
22. Garbe, M.; Junge, K.; Walker, S.; Wei, Z.; Jiao, H.; Spannenberg, A.; Bachmann, S.; Scalone, M.; Beller, M. Manganese(I)-Catalyzed
Enantioselective Hydrogenation of Ketones Using a Defined Chiral PNP Pincer Ligand. Angew. Chem. 2017, 129, 11389–11393.
[CrossRef]
23. Zhang, L.; Wang, Z.; Han, Z.; Ding, K. Manganese-Catalyzed anti -Selective Asymmetric Hydrogenation of α-Substituted
β-Ketoamides. Angew. Chem. 2020, 132, 15695–15699. [CrossRef]
24. Wang, Y.; Zhu, L.; Shao, Z.; Li, G.; Lan, Y.; Liu, Q. Unmasking the Ligand Effect in Manganese-Catalyzed Hydrogenation:
Mechanistic Insight and Catalytic Application. J. Am. Chem. Soc. 2019, 141, 17337–17349. [CrossRef]
25. Glatz, M.; Stöger, B.; Himmelbauer, D.; Veiros, L.F.; Kirchner, K. Chemoselective Hydrogenation of Aldehydes under Mild,
Base-Free Conditions: Manganese Outperforms Rhenium. ACS Catal. 2018, 8, 4009–4016. [CrossRef]
Molecules 2021, 26, 4072 11 of 13
26. Van Putten, R.; Uslamin, E.A.; Garbe, M.; Liu, C.; Gonzalez-de Castro, A.; Lutz, M.; Junge, K.; Hensen, E.J.; Beller, M.; Lefort, L.; et
al. Non-Pincer-Type Manganese Complexes as Efficient Catalysts for the Hydrogenation of Esters. Angew. Chem. Int. Ed. 2017,
56, 7531–7534. [CrossRef]
27. Ryabchuk, P.; Stier, K.; Junge, K.; Checinski, M.P.; Beller, M. Molecularly Defined Manganese Catalyst for Low-Temperature
Hydrogenation of Carbon Monoxide to Methanol. J. Am. Chem. Soc. 2019, 141, 16923–16929. [CrossRef] [PubMed]
28. Zeng, L.; Yang, H.; Zhao, M.; Wen, J.; Tucker, J.H.; Zhang, X. C1-Symmetric PNP Ligands for Manganese-Catalyzed Enantioselec-
tive Hydrogenation of Ketones: Reaction Scope and Enantioinduction Model. ACS Catal. 2020, 10, 13794–13799. [CrossRef]
29. Liu, C.; van Putten, R.; Kulyaev, P.O.; Filonenko, G.A.; Pidko, E.A. Computational insights into the catalytic role of the base
promoters in ester hydrogenation with homogeneous non-pincer-based Mn-P,N catalyst. J. Catal. 2018, 363, 136–143. [CrossRef]
30. Van Putten, R.; Filonenko, G.A.; Gonzalez De Castro, A.; Liu, C.; Weber, M.; Müller, C.; Lefort, L.; Pidko, E. Mechanistic
Complexity of Asymmetric Transfer Hydrogenation with Simple Mn-Diamine Catalysts. Organometallics 2019, 38, 3187–3196.
[CrossRef]
31. Zirakzadeh, A.; de Aguiar, S.R.; Stöger, B.; Widhalm, M.; Kirchner, K. Enantioselective Transfer Hydrogenation of Ketones
Catalyzed by a Manganese Complex Containing an Unsymmetrical Chiral PNP’ Tridentate Ligand. ChemCatChem 2017,
9, 1744–1748. [CrossRef]
32. Rawat, K.S.; Mahata, A.; Choudhuri, I.; Pathak, B. Catalytic hydrogenation of CO2 by manganese complexes: Role of π-acceptor
ligands. J. Phys. Chem. C 2016, 120, 16478–16488. [CrossRef]
33. Yang, X. Hydrogenation of carbon dioxide catalyzed by PNP pincer iridium, iron, and cobalt complexes: A computational design
of base metal catalysts. ACS Catal. 2011, 1, 849–854. [CrossRef]
34. Mondal, B.; Sengupta, K.; Rana, A.; Mahammed, A.; Botoshansky, M.; Dey, S.G.; Gross, Z.; Dey, A. Cobalt corrole catalyst for
efficient hydrogen evolution reaction from H2 O under ambient conditions: Reactivity, spectroscopy, and density functional
theory calculations. Inorg. Chem. 2013, 52, 3381–3387. [CrossRef] [PubMed]
35. Ganguly, G.; Malakar, T.; Paul, A. Theoretical studies on the mechanism of homogeneous catalytic olefin hydrogenation and
amine-borane dehydrogenation by a versatile boryl-ligand-based cobalt catalyst. ACS Catal. 2015, 5, 2754–2769. [CrossRef]
36. Della Monica, F.; Vummaleti, S.V.; Buonerba, A.; Nisi, A.D.; Monari, M.; Milione, S.; Grassi, A.; Cavallo, L.; Capacchione, C.
Coupling of Carbon Dioxide with Epoxides Efficiently Catalyzed by Thioether-Triphenolate Bimetallic Iron(III) Complexes:
Catalyst Structure–Reactivity Relationship and Mechanistic DFT Study. Adv. Synth. Catal. 2016, 358, 3231–3243. [CrossRef]
37. Murugesan, K.; Wei, Z.; Chandrashekhar, V.G.; Neumann, H.; Spannenberg, A.; Jiao, H.; Beller, M.; Jagadeesh, R.V. Homogeneous
cobalt-catalyzed reductive amination for synthesis of functionalized primary amines. Nat. Commun. 2019, 10, 1–9. [CrossRef]
[PubMed]
38. Li, Y.Y.; Tong, L.P.; Liao, R.Z. Mechanism of Water Oxidation Catalyzed by a Mononuclear Iron Complex with a Square
Polypyridine Ligand: A DFT Study. Inorg. Chem. 2018, 57, 4590–4601. [CrossRef] [PubMed]
39. Wiedner, E.S.; Chambers, M.B.; Pitman, C.L.; Bullock, R.M.; Miller, A.J.; Appel, A.M. Thermodynamic Hydricity of Transition
Metal Hydrides. Chem. Rev. 2016, 116, 8655–8692. [CrossRef]
40. Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992,
97, 2571–2577. [CrossRef]
41. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr.
J. Chem. Phys. 1994, 100, 5829–5835. [CrossRef]
42. 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. [CrossRef] [PubMed]
43. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [CrossRef] [PubMed]
44. Chin, B.; Lough, A.J.; Morris, R.H.; Schweitzer, C.T.; D’Agostino, C. Influence of Chloride versus Hydride on H-H Bonding and
Acidity of the Trans Dihydrogen Ligand in the Complexes trans-[Ru(H2 )X(PR2 CH2 CH2 PR2 )2 ]+ , X = Cl, H, R = Ph, Et. Crystal
Structure Determinations of [RuCl(dppe)2 ]PF6 and trans-[Ru(H2 )Cl(dppe)2 ]PF6 . Inorg. Chem. 1994, 33, 6278–6288. [CrossRef]
45. Guerchais, V.; Astruc, D.; Nunn, C.M.; Cowley, A.H. Generation, Characterization, and Chemistry of the Methylene Com-
plexes [Fe(η 5 -C5 Me5 )(CO)(L)(=CH2 )]+ (L = CO, PPh3 ) and the X-ray Crystal Structure of [Fe(η 5 -C5 Me5 )(CO)2 (CH2 PPh3 )]+ BF4 − .
Organometallics 1990, 9, 1036–1041. [CrossRef]
46. Klug, C.M.; Dougherty, W.G.; Kassel, W.S.; Wiedner, E.S. Electrocatalytic Hydrogen Production by a Nickel Complex Containing
a Tetradentate Phosphine Ligand. Organometallics 2019, 38, 1269–1279. [CrossRef]
47. Neale, S.E. Computational Investigations of Homogeneous Catalysis and Spin- State Energetics in Fe ( II ) and Co ( III ) Complexes.
Ph.D. Thesis, Heriot-Watt University, Edinburgh, UK, 2019.
48. Bühl, M.; Kabrede, H. Geometries of transition-metal complexes from density-functional theory. J. Chem. Theory Comput. 2006,
2, 1282–1290. [CrossRef]
49. 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. [CrossRef]
50. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J.
Comput. Chem. 2011, 32, 1457–1465. [CrossRef]
51. Perdew, J.P.; Ruzsinszky, A.; Tao, J.; Staroverov, V.N.; Scuseria, G.E.; Csonka, G.I. Prescription for the design and selection of
density functional approximations: More constraint satisfaction with fewer fits. J. Chem. Phys. 2005, 123, 062201. [CrossRef]
Molecules 2021, 26, 4072 12 of 13
52. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [CrossRef]
53. 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. [CrossRef] [PubMed]
54. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results obtained with the correlation energy density functionals of Becke and Lee,
Yang and Parr. Chem. Phys. Lett. 1989, 157, 200–206. [CrossRef]
55. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system.
Phys. Rev. B 1996, 54, 16533–16539. [CrossRef]
56. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and
surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687.
[CrossRef] [PubMed]
57. Shi, J.M.; Peeters, F.M.; Hai, G.Q.; Devreese, J.T. Erratum: Donor transition energy in GaAs superlattices in a magnetic field along
the growth axis. Phys. Rev. B 1993, 48, 4978–4978. [CrossRef] [PubMed]
58. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behaviour. Phys. Rev. A 1988,
38, 3098–3100. [CrossRef] [PubMed]
59. Perdew, J.P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986,
33, 8822–8824. [CrossRef]
60. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange-correlation functional using the Coulomb-attenuating method
(CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [CrossRef]
61. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics,
noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class
functionals and 12 other function. Theor. Chem. Acc. 2008, 120, 215–241. [CrossRef]
62. Zhao, Y.; Truhlar, D.G. Comparative DFT study of van der Waals complexes: Rare-gas dimers, alkaline-earth dimers, zinc dimer
and zinc-rare-gas dimers. J. Phys. Chem. A 2006, 110, 5121–5129. [CrossRef] [PubMed]
63. Zhao, Y.; Truhlar, D.G. Density functional for spectroscopy: No long-range self-interaction error, good performance for Rydberg
and charge-transfer states, and better performance on average than B3LYP for ground states. J. Phys. Chem. A 2006, 110, 13126–
13130. [CrossRef] [PubMed]
64. Zhao, Y.; Truhlar, D.G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical
kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101. [CrossRef] [PubMed]
65. 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. [CrossRef]
66. Peverati, R.; Truhlar, D.G. M11-L: A local density functional that provides improved accuracy for electronic structure calculations
in chemistry and physics. J. Phys. Chem. Lett. 2012, 3, 117–124. [CrossRef]
67. Peverati, R.; Truhlar, D.G. An improved and broadly accurate local approximation to the exchange-correlation density functional:
The MN12-L functional for electronic structure calculations in chemistry and physics. Phys. Chem. Chem. Phys. 2012, 14, 13171–
13174. [CrossRef]
68. Yu, H.S.; He, X.; Li, S.L.; Truhlar, D.G. MN15: A Kohn-Sham global-hybrid exchange-correlation density functional with broad
accuracy for multi-reference and single-reference systems and noncovalent interactions. Chem. Sci. 2016, 7, 5032–5051. [CrossRef]
69. Yu, H.S.; He, X.; Truhlar, D.G. MN15-L: A New Local Exchange-Correlation Functional for Kohn-Sham Density Functional Theory
with Broad Accuracy for Atoms, Molecules, and Solids. J. Chem. Theory Comput. 2016, 12, 1280–1293. [CrossRef] [PubMed]
70. Perdew, J.P.; Burke, K.; Ernzerhof, M. Errata: Generalized Gradient Approximation Made Simple [Phys. Rev. Lett. 77, 3865
(1996)]. Phys. Rev. Lett. 1997, 78, 1396. [CrossRef]
71. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem.
Phys. 1999, 110, 6158–6170. [CrossRef]
72. Guido, C.A.; Brémond, E.; Adamo, C.; Cortona, P. Communication: One third: A new recipe for the PBE0 paradigm. J. Chem.
Phys. 2013, 138, 021104. [CrossRef]
73. Ernzerhof, M.; Perdew, J.P. Generalized gradient approximation to the angle- and system-averaged exchange hole. J. Chem. Phys.
1998, 109, 3313–3320. [CrossRef]
74. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the density functional ladder: Nonempirical meta–generalized
gradient approximation designed for molecules and solids. Phys. Rev. Lett. 2003, 91, 146401. [CrossRef]
75. Staroverov, V.N.; Scuseria, G.E.; Tao, J.; Perdew, J.P. Comparative assessment of a new nonempirical density functional: Molecules
and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 12129–12137; Erratum in J. Chem. Phys. 2004, 121, 11507. [CrossRef]
76. Chai, J.D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections.
Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [CrossRef]
77. Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput.
Chem. 2006, 27, 1787–1799. [CrossRef]
78. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.;
Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian Inc.: Wallingford, CT, USA, 2013.
79. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.;
Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2019.
Molecules 2021, 26, 4072 13 of 13
80. Estes, D.P.; Vannucci, A.K.; Hall, A.R.; Lichtenberger, D.L.; Norton, J.R. Thermodynamics of the metal-hydrogen bonds in
(η 5 -C5 H5 )M(CO)2 H (M = Fe, Ru, Os). Organometallics 2011, 30, 3444–3447. [CrossRef]
81. Ciancanelli, R.; Noll, B.C.; DuBois, D.L.; Rakowski DuBois, M. Comprehensive thermodynamic characterization of the metal-
hydrogen bond in a series of cobalt-hydride complexes. J. Am. Chem. Soc. 2002, 124, 2984–2992. [CrossRef]
82. Wiedner, E.S.; Appel, A.M.; Dubois, D.L.; Bullock, R.M. Thermochemical and mechanistic studies of electrocatalytic hydrogen
production by cobalt complexes containing pendant amines. Inorg. Chem. 2013, 52, 14391–14403. [CrossRef] [PubMed]
83. Moltved, K.A.; Kepp, K.P. Chemical Bond Energies of 3d Transition Metals Studied by Density Functional Theory. J. Chem.
Theory Comput. 2018, 14, 3479–3492. [CrossRef] [PubMed]
84. Himmel, D.; Goll, S.K.; Leito, I.; Krossing, I. A unified pH scale for all phases. Angew. Chem. Int. Ed. 2010, 49, 6885–6888.
[CrossRef] [PubMed]
85. Bühl, M.; Reimann, C.; Pantazis, D.A.; Bredow, T.; Neese, F. Geometries of third-row transition-metal complexes from density-
functional theory. J. Chem. Theory Comput. 2008, 4, 1449–1459. [CrossRef] [PubMed]
86. Ariyaratne, J.K.P.; Bierrum, A.M.; Green, M.L.H.; Ishaq, M.; Prout, C.K.; Swanwick, M.G. Evidence for near-neighbour
interactions in some substituted methyl derivatives of transition metals including the molecular crystal structure determinations
of (π-C5 H5 )Fe(CO)2 CH2 CO2 H and (π-C5 H5 )Mo(CO)3 CH2 CO2 H. J. Chem. Soc. A 1969, 1309–1321. [CrossRef]