Abstract
We have quantum chemically studied activation of HnA–AHn bonds (AHn = CH3, NH2, OH, F) by PdLn catalysts with Ln = no ligand, PH3, (PH3)2, using relativistic density functional theory at ZORA-BLYP/TZ2P. The activation energy associated with the oxidative addition step decreases from H3C–CH3 to H2N–NH2 to HO–OH to F–F, where the activation of the F–F bond is barrierless. Activation strain and Kohn–Sham molecular orbital analyses reveal that the enhanced reactivity along this series of substrates originates from a combination of (i) reduced activation strain due to a weaker HnA–AHn bond; (ii) decreased Pauli repulsion as a result of a difference in steric shielding of the HnA–AHn bond; and (iii) enhanced backbonding interaction between the occupied 4d atomic orbitals of the palladium catalyst and σ* acceptor orbital of the substrate.
Introduction
Catalysis is intertwined with synthetic chemistry and is of central importance to the chemical industry, as approximately 90 % of all industrial processes make use of a catalyst [1]. One of the most prominent catalytic transformations is the palladium-catalyzed cross-coupling reaction, where activated molecular fragments are reacting to form a new bond, typically, between carbon or other main group elements, such as C, N, O, yielding C–C, C–N or C–O bonds, respectively (Scheme 1) [2]. The catalytic cycle of the palladium-catalyzed cross-coupling begins with the activation of the bond between two cojoined main group elements (HnA–AHn, AHn = CH3, NH2, OH, F) by oxidative addition to the palladium center. The insertion of Pd into the HnA–AHn bond is commonly observed as the rate-determining step [3] and has been subjected to extensive experimental [4] and theoretical studies [5, 6]. The next step involves transmetalation where one of the AHn groups bound to the palladium center is replaced by a hydrocarbon R. In the last step, the original palladium catalyst is regenerated by a reductive elimination step (the reverse of oxidative addition), resulting in a new HnA–R bond. To enable the design of better performing catalysts in terms of rate and selectivity, a detailed physical understanding of the oxidative addition step is required related to the difference of the electronic and steric properties of the catalyst as well as the HnA–AHn bond.

Schematic catalytic cycle of a palladium-catalyzed cross-coupling reaction.
To delineate the effect of varying the nature of the HnA–AHn bond on the bond activation process, we have quantum chemically explored the potential energy surface (PES) of the oxidative insertion of PdLn, with Ln = no ligand, PH3, (PH3)2, into the HnA–AHn bond (AHn = CH3, NH2, OH, F), using relativistic density functional theory at ZORA-BLYP/TZ2P (Scheme 2). The activation strain model (ASM) [7] in combination with quantitative Kohn–Sham molecular orbital (KS-MO) theory [8] and a matching energy decomposition analysis (EDA) scheme [9] were employed to unravel the trends in reactivity and provide quantitative insights into the effect of varying AHn on the HnA–AHn bond activation. This computational methodology provides deep physical insight into the factors controlling reactivity and has proven useful for the understanding of, among others, related oxidative addition reactions [6].

Model oxidative addition reaction between PdLn and HnA–AHn, where Ln = no ligand, PH3, (PH3)2; and AHn = CH3, NH2, OH, F.
Computational methods
Computational details
All calculations were executed with the Amsterdam Density Functional (ADF) program [10]. The generalized gradient approximation (GGA) functional BLYP [11] was used for the optimizations of all stationary points and subsequent analyses (spin-unrestricted formalism was used for computing the bond dissociation enthalpies). The basis set used, denoted TZ2P [12], is of triple-ζ quality and is augmented with two sets of polarization functions on each atom. Scalar relativistic effects were taken into account using the zeroth-order regular approximation (ZORA) [13]. This level of theory is denoted as ZORA-BLYP/TZ2P and has been widely tested with several ab initio reference benchmarks up until the coupled cluster CCSD(T) [14]. The accuracies of the fit scheme (Zlm fit) [15a] and the integration grid (Becke grid) [15b] were set to VERYGOOD. Through vibrational analysis, all stationary points were confirmed to be either equilibrium structures (zero imaginary frequencies) or transition states (one single imaginary frequency) [16]. Furthermore, the normal mode character associated with the imaginary frequency was analyzed to ensure that the correct transition state was found. The potential energy surfaces (PESs) of the studied oxidative addition reactions were obtained by utilizing intrinsic reaction coordinate (IRC) calculations [17]. The acquired PESs were further analyzed using the PyFrag 2019 program [18]. Additional calculations including implicit solvation (THF and water) are performed using the conductor-like screening model (COSMO) [19], as implemented in the ADF program. All stationary-point structures were illustrated using CYLview [20].
Activation strain model and energy decomposition analysis
The activation strain model (ASM [7], also known as the distortion/interaction model [21]) is a fragment-based approach to understand the energy profile of a chemical process in terms of the original reactants, which are the palladium catalyst and the substrate HnA–AHn. It considers their rigidity and the extent to which the reactants must deform during the reaction plus their ability to interact as the reaction proceeds. In this model, we decompose the total energy, ΔE(ζ), into the strain and interaction energy, ΔE strain(ζ) and ΔE int(ζ), respectively, along the IRC which is projected onto a reaction coordinate ζ that is critically involved in the reaction [Eq. (1)].
In this equation, the strain energy, ∆E strain(ζ), is the energy required to deform the reactants from their equilibrium structure to the geometry they acquire during the reaction at an arbitrary point ζ of the reaction coordinate. On the other hand, the interaction energy, ∆E int(ζ), accounts for all the mutual interactions that occur between the deformed fragments along the reaction coordinate.
The interaction energy between the deformed reactants is further analyzed with the help of our canonical energy decomposition analysis (EDA) scheme [9]. The EDA decomposes the ∆E int(ζ) into the following three energy terms [Eq. (2)]:
From this equation, ∆V elstat(ζ) is the quasi-classical electrostatic interaction between the unperturbed charge distributions of the deformed reactants. The Pauli repulsion, ∆E Pauli(ζ), emerges from the destabilizing interaction between occupied orbitals (more precisely, electrons of same spin) on either of the fragments due to Pauli’s exclusion principle. Lastly, the orbital interaction energy, ∆E oi(ζ), accounts for charge transfer (e.g., HOMO–LUMO interactions) and polarization between the fragments. A detailed, step-by-step guide on how to perform and interpret the ASM and EDA can be found in Ref. [7c].
In this work, the activation strain and energy decomposition analyses were carried out along the intrinsic reaction coordinate (IRC) projected onto the stretch of the activated HnA⋯AHn bond, which is a critical geometric parameter of the reaction [22]. This particular geometric parameter undergoes a well-defined change during the reaction going from the reactant complex via the transition state to the product complex and has been shown to be a useful reaction coordinate for studying oxidative addition reactions [6].
Thermochemistry
The bond dissociation energies (BDE), also known as bond enthalpies (ΔH BDE), are calculated at normal temperature and pressure (NTP, i.e., 298.15 K and 1 atm) from electronic bond energies (ΔE) and vibrational frequencies using the canonical thermochemistry relations for an ideal gas [Eq. (3)] [23].
Herein, ΔE trans,298.15, ΔE rot,298.15, and ΔE vib,0 are the differences in translational, rotational, and zero-point vibrational energy between the HnA–AHn substrate and the homolytically dissociated HnA• radicals. The last term, Δ(ΔE vib,0)298.15, is the vibrational correction energy to bring the system from 0 K to 298.15 K.
Results and discussion
We have analyzed the single C–C, N–N, O–O and F–F bonds in the series HnA–AHn (AHn = CH3, NH2, OH, and F). First, we compare the lengths and strengths of these bonds in the different substrates, which are important factors for the overall bond activation (vide infra). Table 1 contains the computed bond lengths and bond dissociation enthalpies at ZORA-(U)BLYP/TZ2P. We find that the HnA–AHn bond becomes systematically weaker and shorter following the trend from A = C to N to O to F, namely, from 83.2 kcal mol−1 and 1.539 Å for H3C–CH3 to 47.8 kcal mol−1 and 1.440 Å for F–F. There is one exception, however, namely, the fact that the bond length does not become shorter from H2N–NH2 (1.457 Å) to HO–OH (1.495 Å), but longer. This is in line with a previous report where this anomaly in bond length is ascribed to the differing degrees of hyperconjugative stabilization between H2N–NH2 and HO–OH [24].
HnA–AHn (AHn = CH3, N2, OH, F) bond lengths (in Å) and bond dissociation enthalpies (ΔH BDE; in kcal mol−1).a
Substrate | HnA–AHn | ΔH BDE |
---|---|---|
H3C–CH3 | 1.539 | 83.2 |
H2N–NH2 | 1.457 | 61.4 |
HO–OH | 1.495 | 52.1 |
F–F | 1.440 | 47.8 |
-
aComputed at ZORA-(U)BLYP/TZ2P (enthalpies at 298.15 K and 1 atm).
The results of our ZORA-BLYP/TZ2P exploration are collected in Table 2 and Fig. 1. Full details and additional data can be found in Tables S1 and S2 and Figs. S1–S3. Note that the overall activation energy ΔE ‡, that is, the energy difference between the TS and the infinitely separated reactants (PdLn and HnA–AHn), can be negative if a substantially stabilized reactant complex is formed. For a more detailed discussion on the various types of reaction potential energy surfaces, see Reference [25]. The oxidative addition reactions of Pd + HnA–AHn (AHn = CH3, NH2, OH, F) generally proceeds via a reactant complex (RC) and a transition state (TS) towards the product complex (PC). Three reactivity trends can be discerned from the computed reaction profiles. First, the corresponding overall activation energy decreases from +18.7 kcal mol−1 for H3C–CH3 to +7.7 kcal mol−1 for H2N–NH2 to −8.5 kcal mol−1 for HO–OH to barrierless for F–F. This latter barrierless process has been previously studied in detail by our group [26]. Second, the coordination of one phosphine ligand to the palladium metal center, i.e., going from Pd to PdPH3, raises the oxidative-addition barrier of the HnA–AHn bonds that are more sterically shielded by A–H bonds (i.e., H3C–CH3 and H2N–NH2), but it does not affect the reactivity trends along the various bonds compared to the situation for the bare model Pd catalyst. Thus, the highest activation energy is still found for the activation of H3C–CH3 (+26.4 kcal mol−1), then consistently lowers upon going to HO–OH (−10.1 kcal mol−1) and remains barrierless for the activation of F–F. Third, coordinating two phosphine ligands to the Pd catalyst, i.e., Pd(PH3)2, further raises the oxidative-addition barrier for the H3C–CH3 and H2N–NH2 bonds. Along H3C–CH3 and H2N–NH2, the barrier continues to decrease, just as for the model catalysts PdPH3 and Pd. For the activation of the other two substrates, that is, HO–OH and F–F, Pd(PH3)2 has a major influence on the reaction mode, because (i) it changes the reaction mode of the former towards a stepwise process in which the phosphine ligands play an active role (Fig. S3); and (ii) the latter substrate reacts via a concerted pathway with an activation energy of −37.0 kcal mol−1.
Relative energies (kcal mol−1) for the oxidative addition reactions of Pd, PdPH3, and Pd(PH3)2 with HnA–AHn (AHn = CH3, NH2, OH, F) computed in the gas phase.a, b, c
Catalyst | Substrate | RC | TS | PC |
---|---|---|---|---|
Pd | H3C–CH3 | −6.7 | 18.7 | −8.7 |
H2N–NH2 | −21.4 | 7.7 | −42.8 | |
HO–OH | −14.6 | −8.5 | −62.3 | |
F–F | −44.1 | d | −100.9 | |
PdPH3 | H3C–CH3 | −7.9 | 26.4 | 14.1 |
H2N–NH2 | −22.4 | 8.8 | −18.7 | |
HO–OH | −14.5 | −10.1 | −51.9 | |
F–F | −42.4 | d | −101.0 | |
Pd(PH3)2 | H3C–CH3 | 0.0 | 51.8 | 27.1 |
H2N–NH2 | −1.9 | 38.2 | −6.2 | |
HO–OH | e | e | e | |
F–F | −39.8 | −37.0 | −111.3 |
-
aComputed at ZORA-BLYP/TZ2P. bSee Figs. 1 and S1–S3 for stationary-point structures. cRC, reactant complex ΔE RC; TS, activation barrier ΔE ‡; PC, product complex ΔE rxn. dNon-existent: oxidative addition is a barrierless process, see Reference [26]. eFollows a different mechanism, see Table S2 and Fig. S3 for full details.
![Fig. 1:
Stationary point structures (in Å) for the oxidative addition of Pd + HnA–AHn (AHn = CH3, NH2, OH, F) computed in the gas phase at ZORA-BLYP/TZ2P. [a] = Non-existent: oxidative addition into F–F is a barrierless process, see Reference [26]. Atom colors: H = white, C = gray, N = blue, O = red, F = green, Pd = orange.](/document/doi/10.1515/pac-2022-1004/asset/graphic/j_pac-2022-1004_fig_001.jpg)
Stationary point structures (in Å) for the oxidative addition of Pd + HnA–AHn (AHn = CH3, NH2, OH, F) computed in the gas phase at ZORA-BLYP/TZ2P. [a] = Non-existent: oxidative addition into F–F is a barrierless process, see Reference [26]. Atom colors: H = white, C = gray, N = blue, O = red, F = green, Pd = orange.
The computed trends in reactivity are also recovered for bulk solvation computed in THF and water, simulated using COSMO(THF)-ZORA-BLYP/TZ2P and COSMO(water)-ZORA-BLYP/TZ2P. Tables S1 and S2 report the effect of solvation on the reaction profile of the studied systems. The activation barriers for Pd + HnA–AHn (AHn = CH3, NH2, OH, F) are all stabilized by solvation, whereas the activation barriers for both PdPH3 + HnA–AHn and Pd(PH3)2 + HnA–AHn are all slightly destabilized. Note that, in solution, the activation of the F–F bond is a barrierless process for all studied catalysts, including Pd(PH3)2.
To gain quantitative insight into the physical factors governing the oxidative addition reactivity trend, we applied the activation strain model (ASM) of reactivity [7]. Fig. 2 displays the activation strain diagrams (ASDs) of the HnA–AHn bond activation (AHn = CH3, NH2, OH) by the model bare Pd catalyst along the IRC projected on the HnA⋯AHn bond stretch. The activation of F–F follows a barrierless process due to the crossing of the singlet and triplet potential energy surfaces. The underlying physics thereof has previously been studied by our group; we refer the interested reader to Reference [26] for complete details. The bare Pd catalyst is chosen for this detailed analysis, because the simplicity of this model catalyst allows for a lucid picture to emerge from the analyses and because, already for this model catalyst, the universal trend (only Pd(PH3)2 + HO–OH and F–F constitute exceptions, vide supra) of a decreasing activation energy occurs. As shown in Table 2, the activation energy goes down from H3C–CH3 to H2N–NH2 to HO–OH (Fig. 2a), which can be traced back to both the strain and interaction energy (Fig. 2b and 2c). The high activation energy of H3C–CH3 is the result of the highly destabilizing strain energy for a twofold reason: (i) The H3C–CH3 is the strongest bond along the investigated series (see Table 1): the stronger the bond, i.e., larger ΔH BDE, the harder it is to break, i.e., the more activation strain it generates during the oxidative addition reaction; (ii) the methyl C–H bonds shield the H3C–CH3 bond, which requires the methyl groups to tilt away from the incoming Pd catalyst [6]. The lowering of the activation energy upon going from H2N–NH2 to HO–OH is exclusively the result of the more stabilizing interaction energy, since the strain energies around the transition state for these two substrates are nearly identical. Based on the bond dissociation energies discussed earlier (Table 1), one would expect that the stronger H2N–NH2 bond experiences a more destabilizing strain energy than the HO–OH during the oxidative addition reaction. This is, however, not the case, because the latter experiences additional deformation of the strong O–H bonds, which, in turn, are stronger than N–H bonds [27], resulting in similar strain energies around the transition states for HO–OH and H2N–NH2 bond activation.

Activation strain analyses: a) total energy, b) strain energy, c) interaction energy; and energy decomposition analyses: d) electrostatic interaction, e) Pauli repulsion, and f) orbital interactions, for the oxidative addition reactions of Pd + HnA–AHn (AHn = CH3, NH2, OH), where the transition states are indicated with a dot and the energy terms along the IRC are projected on the HnA⋯AHn bond stretch. Computed at ZORA-BLYP/TZ2P.
Next, we turn to the energy decomposition analysis (EDA) [9] to obtain a better understanding of why the interaction energy becomes more stabilizing when the substrate changes from H2N–NH2 to HO–OH. The energy decomposition analysis diagrams reveal that the enhanced interaction energy upon going from H2N–NH2 to HO–OH is a result of two factors, namely, (i) a less destabilizing Pauli repulsion (Fig. 2e); and (ii) more stabilizing orbital interactions (Fig. 2f). The electrostatic interactions (Fig. 2d), on the other hand, are not responsible for the trend in reactivity, because they run counter to the observed reactivity trend.
The origin of the less destabilizing Pauli repulsion for the oxidative addition reaction between Pd and HO–OH compared to H2N–NH2 was further investigated by performing a Kohn-Sham molecular orbital (KS-MO) [8] analysis (Fig. 3). The overlap between occupied orbitals of Pd, H2N–NH2, and HO–OH that engage in a repulsive closed-shell–closed-shell orbital interaction was quantified at consistent geometries obtained from the IRC with a HnA⋯AHn bond stretch of 0.441 Å. Performing this analysis at a consistent point along the reaction coordinate (near all transition state structures), rather than on the individual transition state structures alone, ensures that the results are not skewed by the position, earlier or later, of the transition state [7b]. For the model bare palladium catalyst, we consider all five occupied 4d atomic orbitals (HOMOPd). The participating occupied orbitals of H2N–NH2 and HO–OH are the π*-HOMOHnA–AHn, where there is a nodal plane between the two AHn fragments. The closed-shell–closed-shell orbital overlap between the 4d AOs of Pd and HnA–AHn is the largest for H2N–NH2 (S = 0.19) and the smallest and, therefore, less destabilizing for HO–OH (S = 0.07) (Fig. 3a).

Origin into the trends in Pauli repulsive interactions. a) Schematic molecular orbital diagram of the most important occupied-occupied orbital overlaps, and b) the key occupied molecular orbitals of HnA–AHn (isovalue = 0.03 Bohr−3/2) of the oxidative addition reaction between Pd and HnA–AHn (AHn = NH2, OH). Computed at ZORA-BLYP/TZ2P along the IRC projected with a HnA⋯AHn bond stretch of 0.441 Å.
The differences in Pauli-orbital overlap originate from the difference in orientation between Pd and HnA–AHn (Fig. 1), which, in turn, is a direct consequence of the degree of steric shielding of the HnA–AHn bond by the A–H bonds. The NH2 groups of H2N–NH2 have a conformation that shields the direct attack of the Pd catalyst. This differs from the OH groups of HO–OH and hence allows the direct attack of the Pd catalyst for this molecular species. Thus, upon bond activation, the NH2 groups need to rotate to accommodate the Pd catalyst, thereby forcing the Pd + H2N–NH2 structure to be symmetric, such that both newly formed Pd⋯NH2 bonds are identical in length. Consequently, there are two contact points between the occupied orbitals of Pd and H2N–NH2, at both NH2 sides, yielding a large closed-shell–closed-shell orbital overlap and hence Pauli repulsion. For HO–OH, on the other hand, the Pd catalyst can approach the HO–OH bond directly without any need for the OH groups to rotate, due to the absence of steric shielding of the O–H bonds. This gives rise to an asymmetric structure where one newly formed Pd⋯OH bond is shorter than the other. As a result, the number of contact points between the occupied orbitals of Pd and HO–OH is reduced, leading to a less destabilizing closed-shell–closed-shell orbital overlap and thus a lower Pauli repulsion.
Finally, we examine why the oxidative addition reaction involving HO–OH proceeds with more stabilizing orbital interactions than H2N–NH2. Changing the substrate from H2N–NH2 to HO–OH results in a strengthening of the HOMOPd–LUMOHnA–AHn backbonding interaction, whereas the LUMOPd–HOMOHnA–AHn bonding interaction remains nearly constant, as explained in the following (Fig. 4). By performing a Kohn–Sham molecular orbital (KS-MO) analysis on consistent geometries obtained from the IRC with a HnA⋯AHn bond stretch of 0.441 Å, we found that the LUMOHnA–AHn drops in energy from −2.4 eV for H2N–NH2 to −6.0 eV for HO–OH, thereby reducing the HOMOPd–LUMOHnA–AHn orbital energy gap (Fig. 4a). The reduction in orbital energy gap, together with the enhanced HOMOPd–LUMOHnA–AHn orbital overlap due to the previously discussed difference in orientation between Pd and HnA–AHn, results in a more stabilizing backbonding interaction for HO–OH compared to H2N–NH2.

Schematic molecular orbital diagram with the key orbital energies and overlaps of a) the HOMOPd–LUMOHnA–AHn backbonding interaction and c) the LUMOPd–HOMOHnA–AHn bonding interaction of the oxidative addition reaction between Pd and HnA–AHn (AHn = NH2, OH). The key molecular orbitals of HnA–AHn contributing to b) the backbonding interaction and d) bonding interaction (isovalue = 0.03 Bohr−3/2). Computed at ZORA-BLYP/TZ2P along the IRC projected with a HnA⋯AHn bond stretch of 0.441 Å.
The LUMOPd–HOMOHnA–AHn bonding interaction, on the other hand, remains nearly constant when going from H2N–NH2 to HO–OH, even though the σ-HOMOHO–OH is lower in energy than HOMOH2N–NH2, resulting in a larger, less stabilizing, orbital energy gap for the former. HO–OH, however, benefits, due to the approach of the reactants, from an additional stabilizing bonding interaction between LUMOPd–π*-HOMOHO–OH, thereby recovering the bonding orbital interactions (Fig. 4c).
Conclusions
The activation barrier for palladium-mediated HnA–AHn bond activation decreases systematically along the series H3C–CH3 to H2N–NH2 to HO–OH to F–F. The activation of the F–F bond is even barrierless. This follows from our quantum chemical exploration of the A–A oxidative addition to palladium in the model reactions of PdLn + HnA–AHn (Ln = no ligand, PH3, (PH3)2; AHn = CH3, NH2, OH, F), using relativistic density functional theory.
Our detailed activation strain and Kohn-Sham molecular orbital analyses reveal that the decreased oxidative addition reaction barrier upon going from H3C–CH3 to H2N–NH2 to HO–OH to F–F mainly originates from the difference in bond strength along this series which decreases along this series. The H3C–CH3 bond is the strongest bond along the investigated series. The stronger the bond, the more energy it costs to break, and hence the more activation strain it generates during the oxidative addition reaction. Furthermore, the methyl C–H bonds shield the H3C–CH3 bond, which requires the methyl groups to bend away from the incoming Pd catalyst, generating additional activation strain.
The activation barrier lowers further from H2N–NH2 to HO–OH due to two additional barrier lowering phenomena, namely, (i) a less destabilizing steric (Pauli) repulsion; and (ii) more stabilizing orbital interactions. The former effect is a result of the difference in steric shielding of the HnA–AHn bond by the AHn groups. The NH2 groups of H2N–NH2 have a conformation that shields the N–N bond, forcing the Pd catalyst to attack the bond more side-on and hence yields a large closed-shell–closed-shell orbital overlap and Pauli repulsion. In the case of HO–OH, the absence of steric shielding of the OH groups allows the direct attack of the Pd catalyst, resulting in less contact points between the occupied orbitals of the reactants and hence decreased Pauli repulsion. The more stabilizing orbital interactions, on the other hand, originate from the energetically lower-lying σ* acceptor orbital of HO–OH compared to H2N–NH2 that can engage in a strong backbonding interaction with the occupied 4d atomic orbitals of the palladium catalyst. Finally, the activation of the F–F bond is a barrierless process due to the crossing of the singlet and triplet potential energy surfaces, as has been described in detail by de Jong et al. [26].
Article note:
A collection of invited papers based on presentations at the Virtual Conference on Chemistry and its Applications 2022 (VCCA-2022) held on-line, 8–12 August 2022.
Acknowledgments
We thank the National Research Foundation of South Africa (NRF, UID grant no. 115979 and CPRR grant no. 141992), Nuffic’s Netherlands Education Support Office (NESO), and Netherlands Organization for Scientific Research (NWO) for financial support. This work was carried out on the Dutch national e-infrastructure with the support of SURF cooperative.
-
Conflict of interest: All authors declare no conflict of interest.
Supporting Information
Additional computational results; Cartesian coordinates, energies, and the number of imaginary frequencies of all stationary points.
References
[1a)] J. Collman, L. Hegedus, J. Norton, R. Finke. Principles and Applications of Organotransition Metal Chemistry, University Science Books, Herndon, USA (1987).Search in Google Scholar
b) F. Diederich, P. J. Stang. Metal-Catalyzed Cross-Coupling Reactions, Wiley-VCH, Weinheim (1998).10.1002/9783527612222Search in Google Scholar
c) J. F. Hartwig. Organotransition Metal Chemistry: From Bonding to Catalysis, University Science Books, Sausalito (2010).Search in Google Scholar
d) P. W. N. M. van Leeuwen. Homogeneous Catalysis: Understanding the Art, Kluwer Academic Publishers, Dordrecht (2004).10.1007/1-4020-2000-7Search in Google Scholar
e) L. Souillart, N. Cramer. Chem. Rev. 115, 9410 (2015), https://doi.org/10.1021/acs.chemrev.5b00138.Search in Google Scholar PubMed
f) P. A. Dub, J. C. Gordon. Nat. Rev. Chem 2, 396 (2018), https://doi.org/10.1038/s41570-018-0049-z.Search in Google Scholar
[2a)] C. C. C. Johansson Seechurn, M. O. Kitching, T. J. Colacot, V. Snieckus. Angew. Chem. Int. Ed. 51, 5062 (2012), https://doi.org/10.1002/anie.201107017.Search in Google Scholar PubMed
b) E.-I. Negishi. Angew. Chem. Int. Ed. 50, 6738 (2011), https://doi.org/10.1002/anie.201101380.Search in Google Scholar PubMed
c) A. Suzuki. Angew. Chem. Int. Ed. 50, 6722 (2011), https://doi.org/10.1002/anie.201101379.Search in Google Scholar PubMed
[3a)] G. B. Smith, G. C. Dezeny, D. L. Hughes, A. O. King, T. R. Verhoeven. J. Org. Chem. 59, 8151 (1994), https://doi.org/10.1021/jo00105a036.Search in Google Scholar
b) Hydrogenation. In Homogeneous Catalysis, pp. 75–100. Springer, Netherlands, Dordrecht (2004).10.1007/1-4020-2000-7_4Search in Google Scholar
c) B. Crociani, S. Antonaroli, L. Canovese, P. Uguagliati, F. Visentin. Eur. J. Inorg. Chem. 732 (2004).10.1002/ejic.200300376Search in Google Scholar
d) R. van Asselt, K. Vrieze, C. J. Elsevier. J. Org. Chem. 480, 27 (1994), https://doi.org/10.1016/0022-328x(94)87099-3.Search in Google Scholar
[4a)] J. C. Weisshaar. Acc. Chem. Res. 26, 213 (1993), https://doi.org/10.1021/ar00028a012.Search in Google Scholar
b) A. E. Shilov, G. B. Shul’pin. Chem. Rev. 97, 2879 (1997), https://doi.org/10.1021/cr9411886.Search in Google Scholar PubMed
c) M. E. van der Boom, D. Milstein. Chem. Rev. 103, 1759 (2003), https://doi.org/10.1021/cr960118r.Search in Google Scholar PubMed
d) R. H. Crabtree. J. Organomet. Chem. 689, 4083 (2004), https://doi.org/10.1016/j.jorganchem.2004.07.034.Search in Google Scholar
e) C.-H. Jun. Chem. Soc. Rev. 33, 610 (2004), https://doi.org/10.1039/b308864m.Search in Google Scholar PubMed
f) M. Lersch, M. Tilset. Chem. Rev. 105, 2471 (2005), https://doi.org/10.1021/cr030710y.Search in Google Scholar PubMed
g) J. R. Hummel, J. A. Boerth, J. A. Ellman. Chem. Rev. 117, 9163 (2016), https://doi.org/10.1021/acs.chemrev.6b00661.Search in Google Scholar PubMed PubMed Central
[5a)] S. Niu, M. B. Hall. Chem. Rev. 100, 353 (2000), https://doi.org/10.1021/cr980404y.Search in Google Scholar PubMed
b) M. Torrent, M. Solà, G. Frenking. Chem. Rev. 100, 439 (2000), https://doi.org/10.1021/cr980452i.Search in Google Scholar PubMed
c) A. Dedieu. Chem. Rev. 100, 543 (2000), https://doi.org/10.1021/cr980407a.Search in Google Scholar PubMed
d) S. Kozuch, C. Amatore, A. Jutand, S. Shaik. Organometallics 24, 2319 (2005), https://doi.org/10.1021/om050160p.Search in Google Scholar
e) L. Xue, Z. Lin. Chem. Soc. Rev. 39, 1692 (2010), https://doi.org/10.1039/b814973a.Search in Google Scholar PubMed
f) D. Balcells, E. Clot, O. Eisenstein. Chem. Rev. 110, 749 (2010), https://doi.org/10.1021/cr900315k.Search in Google Scholar PubMed
g) M. Besora, C. Gourlaouen, B. Yates, F. Maseras. Dalton Trans. 40, 11089 (2011), https://doi.org/10.1039/c1dt10983a.Search in Google Scholar PubMed
h) M. Garcia-Melchor, A. A. Braga, A. Lledos, G. Ujaque, F. Maseras. Acc. Chem. Res. 46, 2626 (2013), https://doi.org/10.1021/ar400080r.Search in Google Scholar PubMed
i) J. Joy, T. Stuyver, S. Shaik. J. Am. Chem. Soc. 142, 3836 (2020), https://doi.org/10.1021/jacs.9b11507.Search in Google Scholar PubMed
j) M. C. D’Alterio, E. Casals-Cruañas, N. V. Tzouras, G. Talarico, S. P. Nolan, A. Poater. Chem. Eur J. 27, 12481 (2021).10.1002/chem.202185461Search in Google Scholar
[6a)] P. Vermeeren, X. Sun, F. M. Bickelhaupt. Sci. Rep. 8, 10729 (2018), https://doi.org/10.1038/s41598-018-28998-3.Search in Google Scholar PubMed PubMed Central
b) L. P. Wolters, W.-J. van Zeist, F. M. Bickelhaupt. Chem. Eur J. 20, 11370 (2014), https://doi.org/10.1002/chem.201403237.Search in Google Scholar PubMed
c) L. P. Wolters, F. M. Bickelhaupt. Chem. Asian J. 10, 2272 (2015), https://doi.org/10.1002/asia.201500368.Search in Google Scholar PubMed
d) T. Hansen, X. Sun, M. Dalla Tiezza, W.-J. van Zeist, J. N. P. van Stralen, D. P. Geerke, L. P. Wolters, T. A. Hamlin, F. M. Bickelhaupt. Chem. Eur. J. 28, e202201093 (2022), https://doi.org/10.1002/chem.202201093.Search in Google Scholar PubMed PubMed Central
e) T. Hansen, X. Sun, M. Dalla Tiezza, W.-J. van Zeist, J. Poater, T. A. Hamlin, F. M. Bickelhaupt. Chem. Eur. J. 28, e202201093 (2022), https://doi.org/10.1002/chem.202103953.Search in Google Scholar PubMed PubMed Central
f) B. P. Moloto, P. Vermeeren, M. Dalla Tiezza, C. Esterhuysen, F. M. Bickelhaupt, T. A. Hamlin. Eur. J. Org. Chem., e202200722 (2022), https://doi.org/10.1002/ejoc.202200722.Search in Google Scholar
[7] For reviews, see:Search in Google Scholar
a) F. M. Bickelhaupt, K. N. Houk, Angew. Chem. 129, 10204 (2017), https://doi.org/10.1002/ange.201701486; Angew. Chem. Int. Ed. 56, 10070 (2017), https://doi.org/10.1002/anie.201701486.Search in Google Scholar
b) P. Vermeeren, T. A. Hamlin, F. M. Bickelhaupt. Chem. Commun. 57, 5880 (2021), https://doi.org/10.1039/d1cc02042k.Search in Google Scholar PubMed PubMed Central
c) For a step-by-step protocol, see: P. Vermeeren, S. C. C. van der Lubbe, C. Fonseca Guerra, F. M. Bickelhaupt, T. A. Hamlin. Nat. Protoc. 15, 649 (2020), https://doi.org/10.1038/s41596-019-0265-0.Search in Google Scholar PubMed
[8a)] R. van Meer, O. V. Gritsenko, E. J. Baerends. J. Chem. Theor. Comput. 10, 4432 (2014), https://doi.org/10.1021/ct500727c.Search in Google Scholar PubMed
b) T. A. Albright, J. K. Burdett, M.-H. Whangbo. Orbital Interactions in Chemistry, John Wiley & Sons, Inc. (2013).10.1002/9781118558409Search in Google Scholar
[9] For an overview of the EDA method, see:Search in Google Scholar
a) T. A. Hamlin, P. Vermeeren, C. Fonseca Guerra, F. M. Bickelhaupt. Complementary Bonding Analysis, S. Grabowsky (Ed.), 199–212, De Gruyter, Berlin (2021); for a detailed overview of the EDA method, see.10.1515/9783110660074-008Search in Google Scholar
b) F. M. Bickelhaupt, E. J. Baerends. Reviews in Computational Chemistry, K. B. Lipkowitz, D. B. Boyd (Eds.), pp. 1–86, Wiley, Hoboken (2000).10.1002/9780470125922.ch1Search in Google Scholar
[10a)] G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra, S. J. A. van Gisbergen, J. G. Snijders, T. Ziegler. J. Comput. Chem. 22, 931 (2001), https://doi.org/10.1002/jcc.1056.Search in Google Scholar
b) C. Fonseca Guerra, J. G. Snijders, G. te Velde, E. J. Baerends. Theor. Chem. Acc. 99, 391 (1998), https://doi.org/10.1007/s002140050021.Search in Google Scholar
c) SCM Theoretical Chemistry, http://www.scm.com, Vrije Universiteit: Amsterdam (Netherlands), (ADF2019.102).Search in Google Scholar
[11a)] A. D. Becke. Phys. Rev. A 38, 3098 (1988), https://doi.org/10.1103/physreva.38.3098.Search in Google Scholar PubMed
b) B. G. Johnson, P. M. W. Gill, J. A. Pople. J. Chem. Phys. 98, 5612 (1993), https://doi.org/10.1063/1.464906.Search in Google Scholar
c) C. Lee, W. Yang, R. G. Parr. Phys. Rev. B 37, 785 (1988), https://doi.org/10.1103/physrevb.37.785.Search in Google Scholar PubMed
d) T. V. Russo, R. L. Martin, P. J. Hay. J. Chem. Phys. 101, 7729 (1994), https://doi.org/10.1063/1.468265.Search in Google Scholar
[12] E. van Lenthe, E. J. Baerends. J. Chem. Phys. 24, 1142 (2003), https://doi.org/10.1002/jcc.10255.Search in Google Scholar PubMed
[13a)] E. van Lenthe, E. J. Baerends, J. G. Snijders. J. Chem. Phys. 99, 4597 (1993), https://doi.org/10.1063/1.466059.Search in Google Scholar
b) E. van Lenthe, E. J. Baerends, J. G. Snijders. J. Chem. Phys. 101, 9783 (1994), https://doi.org/10.1063/1.467943.Search in Google Scholar
c) E. van Lenthe, A. Ehlers, E. J. Baerends. J. Chem. Phys. 110, 8943 (1999), https://doi.org/10.1063/1.478813.Search in Google Scholar
[14a)] G. T. de Jong, M. Solà, L. Visscher, F. M. Bickelhaupt. J. Chem. Phys. 121, 9982 (2004), https://doi.org/10.1063/1.1792151.Search in Google Scholar PubMed
b) G. T. de Jong, D. P. Geerke, A. Diefenbach, F. M. Bickelhaupt. J. Chem. Phys. 313, 261 (2005), https://doi.org/10.1016/j.chemphys.2005.01.017.Search in Google Scholar
c) G. T. De Jong, D. P. Geerke, A. Diefenbach, M. Solà, F. M. Bickelhaupt. J. Comput. Chem. 26, 1006 (2005), https://doi.org/10.1002/jcc.20233.Search in Google Scholar PubMed
d) G. Th. de Jong, F. M. Bickelhaupt. J. Phys. Chem. A 109, 9685 (2005), https://doi.org/10.1021/jp053587i.Search in Google Scholar PubMed
[15a)] M. Franchini, P. H. T. Philipsen, E. van Lenthe, L. Visscher. J. Chem. Theor. Comput. 10, 1994 (2014), https://doi.org/10.1021/ct500172n.Search in Google Scholar PubMed
b) M. Franchini, P. H. T. Philipsen, L. Visscher. J. Comput. Chem. 34, 1819 (2013), https://doi.org/10.1002/jcc.23323.Search in Google Scholar PubMed
[16a)] A. Bérces, R. M. Dickson, L. Fan, H. Jacobsen, D. P. Swerhone, T. Ziegler. Comput. Phys. Commun. 100, 247 (1997), https://doi.org/10.1016/s0010-4655(96)00120-8.Search in Google Scholar
b) H. Jacobsen, A. Bérces, D. P. Swerhone, T. Ziegler. Comput. Phys. Commun. 100, 263 (1997), https://doi.org/10.1016/s0010-4655(96)00119-1.Search in Google Scholar
c) S. K. Wolff. Int. J. Quant. Chem. 104, 645 (2005), https://doi.org/10.1002/qua.20653.Search in Google Scholar
[17a)] K. Fukui. Acc. Chem. Res. 14, 363 (1981), https://doi.org/10.1021/ar00072a001.Search in Google Scholar
b) L. Deng, T. Ziegler, L. Fan. J. Chem. Phys. 99, 3823 (1993), https://doi.org/10.1063/1.466129.Search in Google Scholar
c) L. Deng, T. Ziegler. Int. J. Quant. Chem. 52, 731 (1994), https://doi.org/10.1002/qua.560520406.Search in Google Scholar
[18a)] X. Sun, T. M. Soini, J. Poater, T. A. Hamlin, F. M. Bickelhaupt. J. Comput. Chem. 40, 2227 (2019), https://doi.org/10.1002/jcc.25871.Search in Google Scholar PubMed PubMed Central
b) X. Sun, T. Soini, L. P. Wolters, W.-J. van Zeist, C. Fonseca Guerra, T. A. Hamlin, F. M. Bickelhaupt. Vrije Universiteit, Amsterdam, The Netherlands (2019), PyFrag.Search in Google Scholar
[19a)] A. Klamt, G. Schüürmann. J. Chem. Soc. Perkin Trans. 2, 799 (1993), https://doi.org/10.1039/P29930000799.Search in Google Scholar
b) A. Klamt. J. Phys. Chem. 99, 2224 (1995).10.1021/j100007a062Search in Google Scholar
c) A. Klamt, V. Jonas. J. Chem. Phys. 105, 9972 (1996), https://doi.org/10.1063/1.472829.Search in Google Scholar
d) C. C. Pye, T. Ziegler. Theor. Chem. Acc. 101, 396 (1999), https://doi.org/10.1007/s002140050457.Search in Google Scholar
[20] C. Y. Legault. Université de Sherbrooke, CYLview2.0 (2020), http://www.cylview.org.Search in Google Scholar
[21a)] D. H. Ess, K. Houk. J. Am. Chem. Soc. 129, 10646 (2007), https://doi.org/10.1021/ja0734086.Search in Google Scholar PubMed
b) D. H. Ess, K. Houk. J. Am. Chem. Soc. 130, 10187 (2008), https://doi.org/10.1021/ja800009z.Search in Google Scholar PubMed
[22] W.-J. van Zeist, A. H. Koers, L. P. Wolters, F. M. Bickelhaupt. J. Chem. Theor. Comput. 4, 920 (2008), https://doi.org/10.1021/ct700214v.Search in Google Scholar PubMed
[23a)] D. A. McQuarrie, J. D. Simon. Physical Chemistry: A Molecular Approach, University Science Books, Sausalito, CA (1997).Search in Google Scholar
b) C. J. Cramer. Essentials of Computational Chemistry: Theories and Models, Wiley, Chichester, West Sussex, England; Hoboken, NJ (2004).Search in Google Scholar
c) F. Jensen. Introduction to Computational Chemistry, Wiley, Chichester, UK; Hoboken, NJ (2017).Search in Google Scholar
d) P. W. Atkins, J. De Paula, J. Keeler. Atkins’ Physical Chemistry, Oxford University Press, Oxford, United Kingdom; New York, NY (2018).Search in Google Scholar
[24a)] D. J. Mckay, J. S. Wright. J. Am. Chem. Soc. 120, 1003 (1998), https://doi.org/10.1021/ja971534b.Search in Google Scholar
b) L. Song, M. Liu, W. Wu, Q. Zhang, Y. Mo. J. Chem. Theor. Comput. 1, 394 (2005), https://doi.org/10.1021/ct049843x.Search in Google Scholar PubMed
[25] F. M. Bickelhaupt. Mass Spectrom. Rev. 20, 347 (2001), https://doi.org/10.1002/mas.10007.Search in Google Scholar PubMed
[26] For a detailed discussion of Pd + F–F reaction, see:G. T. de Jong, A. Kovács, F. M. Bickelhaupt. J. Phys. Chem. A 110, 7943 (2006), https://doi.org/10.1021/jp061501v.Search in Google Scholar PubMed
[27a)] B. Ruscic, J. Berkowitz. J. Chem. Phys. 95, 4378 (1991), https://doi.org/10.1063/1.461761.Search in Google Scholar
b) S. W. Benson. J. Chem. Educ. 42, 502 (1965), https://doi.org/10.1021/ed042p502.Search in Google Scholar
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/pac-2022-1004).
© 2023 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/
This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Obituary for Professor Hugh Burrows, Scientific Editor of Pure and Applied Chemistry
- Preface
- The virtual conference on chemistry and its applications, VCCA-2022, 8–12 August 2022
- Conference papers
- Production and characterization of a bioflocculant produced by Proteus mirabilis AB 932526.1 and its application in wastewater treatment and dye removal
- Palladium-catalyzed activation of HnA–AHn bonds (AHn = CH3, NH2, OH, F)
- Mechanistic aspect for the atom transfer radical polymerization of itaconimide monomers with methyl methacrylate: a computational study
- A new freely-downloadable hands-on density-functional theory workbook using a freely-downloadable version of deMon2k
- Liquid phase selective oxidation of cyclohexane using gamma alumina doped manganese catalysts and ozone: an insight into reaction mechanism
- Exploring alkali metal cation⋯hydrogen interaction in the formation half sandwich complexes with cycloalkanes: a DFT approach
- Expanding the Australia Group’s chemical weapons precursors control list with a family-based approach
- Effect of solvent inclusion on the structures and solid-state fluorescence of coordination compounds of naphthalimide derivatives and metal halides
- Peripheral inflammation is associated with alterations in brain biochemistry and mood: evidence from in vivo proton magnetic resonance spectroscopy study
- A framework for integrating safety and environmental impact in the conceptual design of chemical processes
- Recent applications of mechanochemistry in synthetic organic chemistry
Articles in the same Issue
- Frontmatter
- In this issue
- Editorial
- Obituary for Professor Hugh Burrows, Scientific Editor of Pure and Applied Chemistry
- Preface
- The virtual conference on chemistry and its applications, VCCA-2022, 8–12 August 2022
- Conference papers
- Production and characterization of a bioflocculant produced by Proteus mirabilis AB 932526.1 and its application in wastewater treatment and dye removal
- Palladium-catalyzed activation of HnA–AHn bonds (AHn = CH3, NH2, OH, F)
- Mechanistic aspect for the atom transfer radical polymerization of itaconimide monomers with methyl methacrylate: a computational study
- A new freely-downloadable hands-on density-functional theory workbook using a freely-downloadable version of deMon2k
- Liquid phase selective oxidation of cyclohexane using gamma alumina doped manganese catalysts and ozone: an insight into reaction mechanism
- Exploring alkali metal cation⋯hydrogen interaction in the formation half sandwich complexes with cycloalkanes: a DFT approach
- Expanding the Australia Group’s chemical weapons precursors control list with a family-based approach
- Effect of solvent inclusion on the structures and solid-state fluorescence of coordination compounds of naphthalimide derivatives and metal halides
- Peripheral inflammation is associated with alterations in brain biochemistry and mood: evidence from in vivo proton magnetic resonance spectroscopy study
- A framework for integrating safety and environmental impact in the conceptual design of chemical processes
- Recent applications of mechanochemistry in synthetic organic chemistry