Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Jul 26.
Published in final edited form as: Organometallics. 2021 Mar 30;40(14):2290–2294. doi: 10.1021/acs.organomet.1c00011

Inherent Selectivity of Pd C-H Activation from Different Metal Oxidation States

Peter Amadeo , Bangaru Bhaskararao , Yun-Fang Yang §, Marisa C Kozlowski ‡,*
PMCID: PMC8528202  NIHMSID: NIHMS1693410  PMID: 34690402

Abstract

In investigating potential control factors that would permit a palladium-catalyzed benzylic vs arene C–H activation as previously reported by our group, it was discovered that the oxidation state of the homogenous palladium species influences the selectivity of C–H activation. DFT calculations show that Pd0 and PdI preferentially activate the sp3 C–H bond in toluene, whereas PdII and PdIII preferentially activate the sp2 C–H bond. This selectivity appears to originate from the steric environment created by the ligand framework on the palladium. As the palladium oxidation state increases, the number of ligand sites increases, which decreases the energetic favorability for activation of the weaker, yet more hindered sp3 C–H bond.

Keywords: C-H activation, palladium catalysis, concerted metalation deprotonation, DFT

Graphical Abstract

graphic file with name nihms-1693410-f0001.jpg

INTRODUCTION

As C–H activation grows as a field, the nuance of selective activation is of increasing interest. A large number of C–H bonds have successfully undergone Pd–mediated activation through a variety of processes. The vast majority of these processes involve Pd(II) concerted metalation deprotonation (CMD) which inherently favors sp2 hybridized carbons vs. sp3 hybridized carbons.1,2,3 This selectivity can be reversed by means of a directing group4,5,6 in CMD processes (Scheme 1ab) or by utilizing radical processes to first abstract a hydrogen and then intercepting the carbon centered radical at a metal center (Scheme 1c).7,8

Scheme 1.

Scheme 1.

Examples of C–H activation by directing group of an arene C–H bond (a) and a benzylic C–H bond (b) as well as by free radical hydrogen atom abstraction (c).

Recent research has shown that oxidation states of Pd asides from Pd0 and PdII are accessible including PdI,9,10 PdIII,11,12,13 and PdIV (Scheme 2).14 This article investigates the inherent activation selectivity for sp2 vs. sp3 hybridized carbons computationally using four oxidation states of Pd and sets the stage for further study of differentially selective processes with these systems.

Scheme 2.

Scheme 2.

Examples of C–H activation with palladium in oxidation states I (a), III (b), and IV (c).

We were motivated to study these processes based on findings from our laboratory that Pd(OAc)2 would selectively activate the benzylic C-H of alkyl arenes (Scheme 3).15,16 Since this selectivity is diametrically opposite of that reported for CMD with PdII, we became interested in determining if any homogeneous Pd catalysts would give rise to the observed selectivity. Thus, this study focuses on two C–H bonds in toluene: the benzylic sp3 center and the least hindered arene sp2 center which is para to the methyl group.

Scheme 3.

Scheme 3.

Our previously reported palladium-catalyzed benzylic C–H activation.

RESULTS AND DISCUSSION

Concerted metalation deprotonation, a variant of σ-bond metathesis, is a well-known mode for PdII C–H activation where a basic ligand abstracts a proton while the Pd–C bond forms. This process is redox neutral and requires an anionic ligand. As such, the oxidation states PdI, PdII, and PdIII were considered with this modality.17 Another, C–H activation pathway is oxidative addition, which is rare in Pd chemistry.18,19,20,21 In this process, the oxidation state of the palladium increases by two as the palladium breaks the C–H bond to form a Pd–C and a Pd–H bond. Oxidative addition was only considered here for Pd0.

It is important to note that the activation energies cannot be compared between the oxidation states without specifying reaction conditions. The relative energies of the different oxidation states of palladium are dependent on the oxidant or reductant present, so the absolute barriers from a precursor common to all eight pathways will not be discussed.

All DFT calculations were carried out using Gaussian 16C.01.22 The B3LYP23,24 and then the M0625 functional were used to explore conformational space and perform geometry optimizations, transition state optimizations, and frequency calculations. The 6–31G* basis set was used for atoms C, H, N, and O while the LANL2DZ ECP was used for Pd.26,27 Single point calculations were carried out on optimized structures with the unrestricted M06 functional and the SMD solvation model28 with toluene as the solvent. The 6–311+G** basis set was used for atoms C, H, N, and O while the LANL2DZ ECP was used for Pd. Conformational space was explored thoroughly with each structure by first determining the correct number of ligands and then performing successive dihedral scans to find the lowest energy. All geometry optimizations were confirmed by the presence of no negative frequencies, and all transition states were confirmed by the presence of one and only one negative frequency. Intrinsic reaction coordinate calculations were also undertaken to confirm each transition state.

The results of the transition state calculations are shown in Figure 1. For each transition state, a number of neutral AcOH ligands were coordinated to the palladium center to determine the most energetically favorable ligand sphere. Pd0and PdI are stabilized by one AcOH ligand, but PdII and PdIII are most stable with the number of anionic acetate ligands corresponding to the respective oxidation states.

Figure 1.

Figure 1.

Transition states and relative barriers for toluene activation with different Pd species (sp3 shown in blue, sp2 shown in red).

There is a notable trend moving progressively from an endergonic process to a more and more exergonic process as the oxidation state increases. In every case, formation of the benzylic adduct (sp3 activation) is more thermodynamically favorable due to π-coordination. The transition state barriers of the C-H activation process also increase as the oxidation states increase.

With respect to selectivity, Pd0 exhibited a barrier 2.9 kcal/mol lower for sp3 activation than sp2 activation. These results indicate a direct oxidative addition mechanism would afford selectivity complementary to PdII CMD. Overall, Pd0 favors oxidative addition to the weakest bond (benzylic CH BDE = 85 kcal/mol, arene CH BDE = 103 kcal/mol).29 For PdI, CMD also favors the sp3 activation by 3.4 kcal/mol. However, for PdII sp2 activation becomes more preferable by 3.4 kcal/mol in accord with prior calculations.30 For PdIII, the same trend is observed, but the gap increases to 5.7 kcal/mol.

An examination of the bond lengths (Table 1) indicates the largest degree of C–H bond cleavage in the Pd0 transition states (1.92/1.71 Å) relative to the toluene C-H bonds (1.08/1.09 Å). As such, these transitions states are the latest. There is a trend where the transition state C–H bond lengths progressively shorten moving from Pd0 to PdIII, with the latter having the earliest transition states. These observations are in accord with the overall thermodynamics; namely, the Hammond postulate would indicate that more endergonic reactions would have later transitions states, as is the case for Pd0, while more exergonic reactions would have earlier transition states, as is the case for PdIII. The PdI and PdII transitions states are intermediate to these extremes.

Table 1.

Transition State C–H Bond Lengthsa

Oxidation state TS sp2 C–H (Å) TS sp3 C–H(Å)
Pd0 1.92 1.71
PdI 1.41 1.41
PdII 1.30 1.32
PdIII 1.26 1.27
a

M06/6–31G*/LANL2DZ

There are numerous studies supporting that “C-H activating systems generally exhibit thermodynamic as well as kinetic preferences for aromatic over benzylic activation,”31,32 However, the data in Figure 1 show a potentially more complex scenario within the Pd manifold. To gain better insight into the reasons behind the computed selectivity values, a distortion interaction energy analysis33 was performed by comparing the transition states to the ground states of the separated components (Table 2, Figure 2). The overall electronic energies confirm that sp3 activation is more favorable for Pd0 oxidation addition and PdI CMD, while sp2 activation is favored for the PdII and PdIII CMD. In general, the interaction energies are strongest for sp2 activation indicating that the potential stabilization from π-coordination of the benzylic portion is not dominant in the transition states. Rather, the interaction of Pd with sp2 carbons is stronger than sp3 carbons, which is at least partly a reflection of the stronger Pd-C(sp2) bond strength.34 Interaction energies are strongest for Pd0 which likely reflects greater back-bonding from the low valent metal center.

Table 2.

Distortion-Interaction Electronic Energiesa

Oxidation state TS sp2 TS sp3
Ed(Pd) Ed(PhCH3) Ed(tot) Ei Eact Ed(Pd) Ed(PhCH3) Ed(tot) Ei Eact
Pd0 1.0 88.5 89.5 −86.4 3.1 0.1 58.2 58.3 −58.4 −0.1a
PdI 3.4 34.7 38.1 −41.7 −3.6a 0.8 28.7 29.5 −35.3 −5.8a
PdII 27.5 28.5 56.0 −48.2 7.8 46.3 20.6 66.9 −51.4 15.5
PdIII 36.9 25.7 62.6 −48.8 13.8 32.6 18.3 50.9 −29.9 21.0
a

M06/6–31G*/LANL2DZ, kcal/mol.

b

Negative values arise from lack of entropic and thermal corrections and comparison to separated starting materials rather than coordination adducts (c.f. Figure 1).

Figure 2.

Figure 2.

Distortion-interaction diagram for toluene activation.

For Pd0 and PdI, there is almost no distortion of the Pd portions in any of the transition states indicating the largely unhindered nature of these low coordinate species. However, the distortion energies of the overall transition states are high (29.5–89.5 kcal/mol) due to the toluene portion. For late transition states, significant distortion of the toluene portion from the ground state aligns with this finding. The distortion energy of the sp2 center of toluene is far higher (34.7–88.5 kcal/mol) than that of the sp3 center (28.7–58.2 kcal/mol), which likely reflects both the stronger sp2 C—H bond and a reduction of aromaticity in the reaction of the sp2 carbon that would arise from a p-orbital of the aromatic π-system participating in the formation of the new Pd–C bond. This effect can be seen in the distortion of the sp2 C—H bond out the aromatic plane in the transition state. Such an interaction does not occur with the sp3 carbon. It is striking the these higher sp2distortion energies are so large that they both offset the higher sp2 interaction energy and destabilize the transition states relative to the sp3 activation. All told, the selective activation of the sp3 centers by Pd0 and PdI is both kinetic and thermodynamically favored due to lesser distortion of the sp3 center upon coordination and reaction.

For PdII and PdIII, there is a large amount of distortion in the Pd portions (27.5–46.3) indicating that significant rearrangement of the coordination sphere is needed from the reactants to the transition state. The overall distortion energies from PdII and PdIII are of similar magnitudes to those from Pd0 and PdI. As a consequence, the toluene distortion energies are smaller for PdII and PdIII (18.3–28.5 kcal/mol). As these transition states are earlier, there is less perturbation of the toluene from its ground state and accordingly lower distortion energies. Although there is less distortion for the toluene portion in the sp3 activation, the interaction energies and Pd distortion energies combine to disfavor sp3 activation relative to sp2 activation. Because the transition states are earlier there is no longer as strong a penalty from loss of aromaticity in the toluene portion upon sp2 activation. Notably, the thermodynamic products appear to be the sp3 activation adducts which are stabilized by benzylic π-coordination; however, the early nature of the transition states precludes this factor from being decisive with respect to the transition state energies. All told, these PdII and PdIII species are more hindered than their Pd0 and PdI analogs resulting in more favorable reaction with the less hindered trigonal carbon that also benefits from forming a stronger bond to the metal.

CONCLUSION

Ligand design Pd0 product structures better for ligand design; ligand can be more distal since Pd-C is largely formed and shorter in TS, however need to have a larger ligand due to the openness of the Pd center. For PdII and PdIII, precordination complexes more relevant, ligands need to intimately interact with incoming substrate, can adjust the electrophilicity with cationic complexes

The selectivity trends discovered herein set the stage for designing future catalysts. For example, ligands could be incorporated to shift the transition states from early to late or vice versa to favor selective activation of different centers. With PdII and PdIII, selection of ligands that destabilize the product could potentially shift selectivity. This work also how C-H activation from Pd0 and PdI could lead to novel reactivity profiles. Overall, ligand frameworks and reaction conditions (e.g. electrochemical) to access these different redox couples with Pd and related elements will allow the fundamental drivers of selectivity to be modified. Such studies would allow even greater control over selectivity in C-H activation, considered one of the Holy Grails in synthetic methods development.35,36

Supplementary Material

Supporting Information 02
Supporting Information 01

ACKNOWLEDGMENT

We are grateful to the NIH (R35 GM131902) and NSF (CHE1764298) for financial support of this research. We acknowledge XSEDE (TGCHEM120052) for computational resources. We thank Prof. Xin Hong (Zhejiang Univ.) for helpful discussions.

Footnotes

The authors declare no competing financial interest.

Computational methods and thermochemical data. This material is available free of charge via the Internet at http://pubs.acs.org.

REFERENCES

  • 1.Fagnou K; Lapointe D Overview of the Mechanistic Work on the Concerted Metallation-Deprotonation Pathway Chem. Lett 2010, 39, 1118–1126. [Google Scholar]
  • 2.Yeung CS; Dong VM Catalytic Dehydrogenative Cross-Coupling: Forming Carbon-Carbon Bonds by Oxidizing Two Carbon-Hydrogen Bonds Chem. Rev 2011, 111, 1215–1292. [DOI] [PubMed] [Google Scholar]
  • 3.He J; Wasa M; Chan KSL; Shao Q; Yu J Palladium-Catalyzed Transformations of Alkyl C–H Bonds Chem. Rev 2017, 117, 8754–8786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Hull KL; Sanford MS Catalytic and Highly Regioselective Cross-Coupling of Aromatic C-H Substrates J. Am. Chem. Soc 2007, 129, 11904–11905. [DOI] [PubMed] [Google Scholar]
  • 5.Lyons TW; Sanford MS Palladium-Catalyzed Ligand-Directed C-H Functionalization Reactions Chem. Rev 2010, 110, 1147–1169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Neufeldt SR; Sanford MS Controlling Site Selectivity in Palladium-Catalyzed C–H Bond Functionalization Acc. Chem. Res 2012, 45, 936–946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Xie P; Xia C; Huang H Palladium-Catalyzed Oxidative Aminocarbonylation: A New Entry to Amides via C–H Activation Org. Lett 2013, 15, 3370–3373. [DOI] [PubMed] [Google Scholar]
  • 8.Xie P; Xie Y; Qian B; Zhou H; Xia C; Huang H Palladium-Catalyzed Oxidative Carbonylation of Benzylic C–H Bonds via Nondirected C(sp3)–H Activation J. Am. Chem. Soc 2012, 134, 9902–9905. [DOI] [PubMed] [Google Scholar]
  • 9.Proutiere F; Aufiero M; Schoenebeck F Reactivity and Stability of Dinuclear Pd(I) Complexes: Studies on the Active Catalytic Species, Insights into Precatalyst Activation and Deactivation, and Application in Highly Selective Cross-Coupling Reactions J. Am. Chem. Soc 2012, 134, 606–612. [DOI] [PubMed] [Google Scholar]
  • 10.Stirner CK; Schoenebeck F; Sperger T Bench-Stable and Recoverable Palladium(I) Dimer as an Efficient Catalyst for Heck Cross-Coupling Synthesis 2017, 49, 115–120. [Google Scholar]
  • 11.Powers DC; Ritter T A Transition State Analogue for the Oxidation of Binuclear Palladium(II) to Binuclear Palladium(III) Complexes Organometallics 2013, 32, 2042–2045. [Google Scholar]
  • 12.Powers DC; Benitez D; Tkatchouk E; III WAG; Ritter T Bimetallic Reductive Elimination from Dinuclear Pd(III) Complexes J. Am. Chem. Soc 2010, 132, 14092–14103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Powers DC; Ritter T Bimetallic Pd(III) complexes in palladium-catalysed carbon–heteroatom bond formation Nat. Chem 2009, 1, 302–309. [DOI] [PubMed] [Google Scholar]
  • 14.Park H; Verma P; Hong K; Yu J Controlling Pd(IV) reductive elimination pathways enables Pd(II)-catalysed enantioselective C(sp3)−H fluorination Nat. Chem 2018, 10, 755–762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Curto JM; Kozlowski MC Chemoselective Activation of sp3 vs sp2 C–H Bonds with Pd(II) J. Am. Chem. Soc 2015, 137, 18–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Hong G; Nahide PD; Neelam UK; Amadeo P; Vijeta A; Curto JM; Hendrick CE; Vangelder KF; Kozlowski MC Palladium-Catalyzed Chemoselective Activation of sp3 vs sp2 C–H Bonds: Oxidative Coupling To Form Quaternary Centers ACS Catal. 2019, 9, 3716–3724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Only acetate assisted intramolecular variants of the CMD transistion states were considered as they have been found to be lowest in energy. See reference 1.
  • 18.Gilbert TM; Hristov I; Ziegler T Comparison between Oxidative Addition and σ-Bond Metathesis as Possible Mechanisms for the Catalytica Methane Activation Process by Platinum(II) Complexes: A Density Functional Theory Study Organometallics 2001, 20, 4633–4639. [Google Scholar]
  • 19.Nag NK A Study on the Formation of Palladium Hydride in a Carbon-Supported Palladium Catalyst J. Phys. Chem 2001, 105, 5945–5949. [Google Scholar]
  • 20.Chin YC; Buda C; Neurock M; Iglesia E Consequences of Metal–Oxide Interconversion for C–H Bond Activation during CH4 Reactions on Pd Catalysts J. Am. Chem. Soc 2013, 135, 15425–15442. [DOI] [PubMed] [Google Scholar]
  • 21.Diefenbach A; Bickelhaupt FM Activation of H-H, C-H, C-C, and C-Cl Bonds by Pd(0). Insight from the Activation Strain Model J. Phys. Chem. A 2004, 108, 8460–8466. [Google Scholar]
  • 22.Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Mennucci B; Petersson GA; Nakatsuji H; Caricato M; Li X; Hratchian HP; Izmaylov AF; Bloino J; Zheng G; Sonnenberg JL; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark M; Heyd JJ; Brothers E; Kudin KN; Staroverov VN; Kobayashi R; Normand J; Raghavachari K; Rendell A; Burant JC; Iyengar SS; Tomasi J; Cossi M; Rega N; Millam NJ; Klene M; Knox JE; Cross JB; Bakken V;Adamo C; Jaramillo J; Gomperts R; Stratmann RE; Yazyev O; Austin AJ; Cammi R; Pomelli C; Ochterski JW; Martin RL; Morokuma K; Zakrzewski VG; Voth GA; Salvador P;Dannenberg JJ; Dapprich S; Daniels AD; Farkas Ö;Foresman JB; Ortiz JV; Cioslowski J; Fox DJ Gaussian 16, Revision C.01; Gaussian, Inc., Wallingford CT, 2019. [Google Scholar]
  • 23.Becke AD Density-functional thermochemistry. III. The role of exact exchange J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]
  • 24.Kohn W; Becke AD; Parr RG Density Functional Theory of Electronic Structure J. Phys. Chem 1996, 100, 12974–12980. [Google Scholar]
  • 25.Zhao Y; Truhlar DG 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 functionals Theor. Chem. Acc 2008, 120, 215–241. [Google Scholar]
  • 26.Hay PJ; Wadt WR Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitals J. Chem. Phys 1985, 82, 299–310. [Google Scholar]
  • 27.Wadt WR; Hay PJ Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg J. Chem. Phys 1985, 82, 270–283. [Google Scholar]
  • 28.Marenich A V; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions J. Phys. Chem. B 2009, 113, 6378–6396. [DOI] [PubMed] [Google Scholar]
  • 29.Benson SW Bond Energies J. Chem. Educ 1965, 42, 502–518. [Google Scholar]
  • 30.Zhou Y; Wang M; Fang S; Chen Y; Liu J DFT studies on the mechanism of palladium catalyzed arylthiolation of unactive arene to diaryl sulfide RSC Adv. 2016, 6, 18300–18307. [Google Scholar]
  • 31.Heyduk AF; Driver TG; Labinger JA; Bercaw JE Kinetic and Thermodynamic Preferences in Aryl vs Benzylic C-H Bond Activation with Cationic Pt(II) Complexes J. Am. Chem. Soc 2004, 126, 15034–15035. [DOI] [PubMed] [Google Scholar]
  • 32.(a) Janowicz AH; Bergman RG Activation of C-H Bonds in Saturated Hydrocarbons on Photolysis of (775-C5Me5)(PMe3)IrH2. Relative Rates of Reaction of the Intermediate with Different Types of C-H Bonds and Functionalization of the Metal-Bound Alkyl Groups J. Am. Chem. Soc 1983, 105, 3929–3939. [Google Scholar]; (b) Jones WD; Hessell ET Photolysis of Tp/Rh(CN-neopentyl)(772-PhN=C=N-neopentyl) in Alkanes and Arenes: Kinetic and Thermodynamic Selectivity of [TpzRh(CN-neopentyl)] for Various Types of C-H Bonds J. Am. Chem. Soc 1993, 115, 554–562. [Google Scholar]; (c) Driver TG; Day MW; Labinger JA; Bercaw JE “Mechanism of C-H Bond Activation of Alkyl-Substituted Benzenes by Cationic Platinum(II) Complexes” Organometallics 2005, 24, 3644–3654. [Google Scholar]; (d) Zhao SB; Song D; Jia WL; Wang S “Regioselective C-H Activation of Toluene with a 1,2-Bis(N-7-azaindolyl)benzene Platinum(II) Complex” Organometallics 2005, 24, 3290–3296. [Google Scholar]; (e) Johansson L; Ryan OB; Rømming C; Tilset M “Unexpected Selectivities in C-H Activations of Toluene and p-Xylene at Cationic Platinum(II) Diimine Complexes. New Mechanistic Insight into Product-Determining Factors” J. Am. Chem. Soc 2001, 123, 6579–6590. [DOI] [PubMed] [Google Scholar]
  • 33.(a) Bickelhaupt FM Understanding reactivity with Kohn–Sham molecular orbital theory: E2–SN2 mechanistic spectrum and other concepts J. Comput. Chem 1999, 20, 114–128. [Google Scholar]; (b) Ess DH; Houk KN Distortion/Interaction Energy Control of 1,3-Dipolar Cycloaddition Reactivity J. Am. Chem. Soc 2007, 129, 10646. [DOI] [PubMed] [Google Scholar]; (c) Bickelhaupt FM; Houk KN Analyzing Reaction Rates with the Distortion/Interaction-Activation Strain Model Angew. Chem., Int. Ed 2017, 56, 10070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Siegbahn PEM Trends of Metal-Carbon Bond Strengths in Transition Metal Complexes J. Phys. Chem 1995, 99, 12723–12729. [Google Scholar]
  • 35.Kozlowski MC Oxidative Coupling in Complexity Building Transforms Acc. Chem. Res 2017, 50, 638–643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Yang Y; Hong X; Yu J; Houk KN Experimental-Computational Synergy for Selective Pd(II)-Catalyzed C-H Activation of Aryl and Alkyl Groups Acc. Chem. Res 2017, 50, 2853–2860. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information 02
Supporting Information 01

RESOURCES