Next Article in Journal
Exploring the Frontiers of Cathode Catalysts in Lithium–Carbon Dioxide Batteries: A Mini Review
Previous Article in Journal
Tetrahydroxidohexaoxidopentaborate(1-) Salts of C6-Linked Substituted Diimidazolium and Dipyrrolidinium Cations: Synthesis, Characterization and XRD Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

C–H Activation via Group 8–10 Pincer Complexes: A Mechanistic Approach

by
Juan S. Serrano-García
1,
Andrés Amaya-Flórez
1,
Jordi R.-Galindo
1,
Lucero González-Sebastián
2,
Luis Humberto Delgado-Rangel
1 and
David Morales-Morales
1,*
1
Instituto de Química, Universidad Nacional Autónoma de México, Ciudad Universitaria, Circuito Exterior s/n, Ciudad de México C.P. 04510, Mexico
2
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, Av. San Rafael Atlixco No. 186, Ciudad de México C.P. 09340, Mexico
*
Author to whom correspondence should be addressed.
Inorganics 2024, 12(8), 221; https://doi.org/10.3390/inorganics12080221
Submission received: 9 July 2024 / Revised: 9 August 2024 / Accepted: 12 August 2024 / Published: 15 August 2024
(This article belongs to the Special Issue C–H Bond Activation by Transition Metal Complexes)

Abstract

:
C–H bond activation is a crucial synthetic strategy widely utilized in both academic and industrial settings. Due to the strong and kinetically inert nature of the C–H bond, its functionalization typically requires metal-based catalysts. This review highlights the most significant advancements in homogeneously catalyzed reactions using pincer complexes with metals from groups 8–10, capable of promoting challenging C–H activation, published since 2010. In particular, it focuses on C–H bond activation for borylation, isomerization, and dehydrogenation, among other processes, discussing their scope and mechanistic insights.

1. Introduction

Functionalizing C–H bonds is one of the most crucial steps in chemical processes for synthesizing chemicals and intermediates essential for manufacturing functional materials [1,2,3]. Additionally, this strategy can reduce waste and energy consumption in synthetic processes. However, this process is challenging because C–H bonds are much less reactive than other bonds, such as C–O or C–Hal. Furthermore, due to the nature of C–H bonds, controlling selectivity is complicated, necessitating the use of directing groups [4,5].
An alternative for activating C–H bond involves the use of transition metals, which represents an attractive strategy both environmentally and economically due to savings in synthesis steps [2,6,7,8,9,10]. To date, C–H activation has been explored using various complexes, including metals such as palladium, rhodium, ruthenium, iridium, and others. The general mechanism typically includes an organometallic step where C–H bond cleavage and the formation of a C–M bond occur; these processes can proceed through mechanisms such as oxidative addition, electrophilic activation, or σ-bond metathesis [11]. Many instances of C–H bond activation exemplify stoichiometric activation at transition metal centers, which is crucial given the challenge of controlling chemoselectivity in reactions mediated by C–H activations [12,13]. Additionally, C–H activation plays a pivotal role in various catalytic processes (e.g., C–H borylation, dehydrogenation, and alkene isomerization), influencing both chemical and steric selectivity in substrates entering the catalytic cycle [14].
Among the different catalysts for C–H activation, those containing cyclometallated species have shown a remarkable role in this process, particularly highlighting the use of metallacycles employing pincer-type ligands [10,12].
Pincer systems are characterized by bearing a tridentate ligand in a meridional conformation [15,16,17,18,19,20]. These pincer ligands can be either neutral or ionic and form two metallacycles, stabilizing the metal through the chelate effect. A general representation of pincer compounds is shown in Figure 1. Pincers are typically named based on their donor atoms (D, D’: C, Si, N, P, O, S, etc.), and sometimes the connectors (Y: CH2, O, N, S, etc.) are also specified [21,22,23]. For example, an NCN pincer contains donor atoms N, C, and N, while a PSCOP pincer contains donor atoms P, C, and P with connectors O and S. Since their first report by Moulton and Shaw in 1976 [24,25,26], pincers have been extensively studied due to their high thermal stability and remarkable catalytic activity in a wide variety of reactions, including C–H activation [18,27,28,29,30,31]. According to the literature, C–H activation is favored using late transition metals (group 8–10), owing to their high catalytic activity, stability, and controlled stereochemistry [4,32,33].
In this review, we summarize the advances in C–H activation catalyzed by group 8–10 metal pincer complexes published since 2010. The scope of this review is limited to intermolecular C–H activation catalyzed with metal complexes. Additionally, we focus on mechanistic studies of C–H activation, which are crucial for clarifying the catalytic mechanisms involved.

2. Group 8 Pincer Complexes

2.1. Iron

2.1.1. C–H Borylation

C–H borylation is a crucial synthetic tool for introducing complex functional groups into simple and abundant starting materials. Borylated compounds can further undergo C–C couplings, oxidation, and other reactions. Catalysis for C–H borylation has been achieved with high yields using metal complexes [34,35,36]. In recent years, efforts have focused on employing more abundant metals, such as iron, for this catalytic process [2,37].
In 2019, Kato et al. [38] reported the borylation of C–H bonds catalyzed with PNP–iron complexes based on 4,5,6,7-tetrahydroisoindole-2-ide. The pincer complexes achieved C–H borylation with various substrates, including arenes and heteroarenes such as pyrrole, furan, and thiophene. Mechanistic studies involved stoichiometric reactions and reactions with substrates in excess to obtain the corresponding PNP–Fe complexes (Scheme 1). Stoichiometric reaction of 1-Ph with B2Pin2 (bis(pinacolato)diboron) yielded the corresponding BPin (pinacolborane) derivative (1-BPin), which reacted with excess benzene to produce 1-H2 and the borylated product (PhBPin). This complex is a dimer from the hydride iron pincer (1-H), which could be the active catalyst during C–H borylation catalysis. Moreover, 1-H2 was able to activate the C–H bond of 2-methylfuran, generating the corresponding furanyl derivative (1-Furanyl).
Similar mechanistic studies were reported by Kamitani et al. in 2019 [39], who investigated the non-targeted borylation of various aromatic compounds using PNN pincer-type iron catalysts. In this study, stoichiometric reactions of the chloride (2-Cl) or hydride (2-H) iron pincer complexes with B2Pin2 produced the BPin derivative (2-BPin), which reacted with an arene to yield the borylated product. A mechanism for this reaction was proposed and is outlined in Scheme 2. The complex enters the catalytic cycle by reacting with B2Pin2 (step 1). Then, C–H activation could occur via a σ-metathesis or oxidative addition pathway (step 2), releasing the borylated product (step 3). Complex 10 would then react with B2Pin2 to continue the catalytic cycle. Kinetic isotope effect (KIE) determination led to a value of 2.0, indicating that the C–H activation process would be the rate-determining step in this catalytic process.

2.1.2. Hydrogen Isotope Exchange (HIE)

Tritium- and deuterium-labeling systems are very important in medicinal chemistry, with several applications such as the identification of receptor dynamics and the localization of potential drugs in vivo [40,41,42,43]. The use of direct hydrogen isotope exchange (HIE) with metal complexes is a strategic method to study mechanisms of action, avoiding conventional multistep synthesis [44,45]. This information allows the tuning of metal complexes to induce selectivity [46].
In this context, Pony Yu et al., in 2016 [47], reported an HIE catalysis with a CNC–Fe(0) pincer (3) for the deuteration and tritiation of pharmaceutical products such as Varenicline (Chantix and Pfizer), Papaverine, Paroxetine (GSK), Loratadine (Claritin and Schering/Merck), MK-6096 (Merck), Cinacalcet (Sensipar and Amgen), Flumazenil (Roche), and Suvorexant (Belsomra and Merck) (Figure 2), employing D2 and T2 as isotope sources. C–H activation selectivity was identified for electron-poor substrates with sterically accessible C–H bonds.
Mechanistic studies on the catalytic activity of complex 3 were carried out by the Chirik group. A stoichiometric reaction with H2 (1 atm) showed a stable dihydride with a σ-H2 ligand (4-H2), which partially incorporated deuterium from a deuterated solvent (benzene-d6) to yield the isotopomer with a σ-H-D ligand (4-HD) [48]. In 2020, Corpas et al. [49] carried out mechanistic studies with 3 in C–H activation and identified that H2 acted as an activator of the catalyst. Given this activation, benzene-d6 could be used as a deuterium source since H2 could facilitate C6HnD6-n (n = 1–6) elimination, yielding the active catalyst 5 (Scheme 3). Although dinitrogen coordination in 5 would be plausible, it is readily displaced under catalytic conditions.
Based on these results, the authors proposed the mechanism shown in Scheme 4. The CNC–Fe(0) is activated with H2 and C6D6 producing the corresponding dideuteride Fe(II) pincer (step 1). The iron dideuteride compound causes C–H activation via σ-C–H bond coordination (step 2), which incorporates a deuterium atom into the aryl ligand via either a sigma metathesis or a Fe(II/IV) cycle (step 3). Then, the deuterated product is eliminated, and the catalyst is again activated with H2 and C6D6 (step 4).
This mechanism is similar to that proposed by the de Rutier group using a PCP–Fe(II) pincer (6) as catalyst and benzene as substrate [50]. In addition, PCP–Fe(0) pincers exhibited the expected selectivity for the most sterically accessible position in reactions with different arene substrates. Computational studies supported a σ-complex-assisted metathesis (σ-CAM), as the critical step in C–H activation (Scheme 5). This leads to the nonclassical hydride isotopomer, which rotates to position the deuterium close to the aryl carbon. Subsequently, the deuterium is incorporated into the aryl ring, forming the σ-C-D adduct of the deuterated product, which is then released.

2.2. Ruthenium

2.2.1. Hydrogen Isotope Exchange (HIE)

Deuterium oxide is an inexpensive source of deuterium for HIE catalysis, but its use is limited by solubility problems that restrict the number of potential catalysts. Therefore, new strategies are required to overcome this challenge. The Periana group implemented strongly basic conditions as a strategy to analyze an NNN–Ru(III) pincer (7) [51]. This complex was reduced in situ with Zn to generate the Ru(II) active catalyst. In the catalytic cycle, a nucleophilic C–H activation pathway is proposed, facilitated by the OD ligand (Scheme 6). Subsequently, HOD is substituted with deuterium oxide (D2O), followed by the deuterium transfer to the arene.

2.2.2. Alkane Dehydrogenation

Alkenes are very important organic compounds for both industrial and academic purposes. They can be selectively obtained from more abundant sources such as alkanes [52,53]. The main disadvantages of industrial dehydrogenation are the high temperatures required (approx. 400–600 °C) and the use of heterogeneous catalysts, which often do not provide optimal regioselectivity. Therefore, alternative options, such as homogeneous catalysts with metal complexes, have emerged in recent years, allowing these processes to be carried out with higher regioselectivity and at lower reaction temperatures [54,55,56]. For instance, Zhou et al. [57] performed the catalytic alkane dehydrogenation of cyclooctane (COA) using a PSP–Ru pincer (8, 1 mM), with t-butylethylene (TBE) (300 mM) as a hydrogen acceptor (Scheme 7). Turnover frequency (TOF) values of 0.1–0.8 s−1 were obtained at 120–180 °C, with the reaction nearly complete at 180 °C after 5 min. DFT calculations highlighted the C–H activation via oxidative addition, followed by β-H elimination, as determining steps in the catalytic cycle. Furthermore, acceptorless dehydrogenation assays were conducted with complex 8 (1 mM) in a COA solution at 151 °C, showing a low yield of cyclooctene (COE, 14 mM) after 48 h.

2.2.3. Direct Alkenylation

C–C cross coupling catalysis is a powerful synthetic tool in organic chemistry for designing complex molecules or inducing specific stereochemistry, often requiring pre-functionalized substrates [58]. However, direct C–C cross-coupling can be achieved from transition-metal-catalyzed C–H activation products [59,60]. For instance, Cai et al. [61] reported the direct alkenylation of several heteroarenes and 2-olefinpyridines using a CCC–Ru(IV) pincer (9). A kinetic isotope effect (KIE) value of 2.3 implied that the C–H activation step was rate-determining, with 2-(o-tolyl)pyridine as the substrate. Additionally, only trace amounts of the alkenylation product were obtained when the radical scavenger TEMPO (2,2,6,6-tetramethylpiperidine-1-oxyl) was added. Based on these results, a mechanism is proposed and is shown in Scheme 8. Complex 9 is reduced in situ to produce the active Ru(III) catalyst (step 1), which undergoes a single electron transfer (SET) (step 2). This intermediate was detected via real-time high-resolution mass spectrometry (HRMS). Then, coordination of the directing group pyridine favors the C–H activation (step 3), followed by the release of the alkenylated product (step 4) [62].

2.2.4. Dehydrogenative Silylation

Arylsilanes are compounds with prominent applications in organic synthesis, cross-coupling, optoelectronics, and more. These compounds can be tuned via silylation catalysis to yield chiral compounds like spirocyclic scaffolds. For example, the Huang group carried out dehydrogenative silylation catalysis [63] using a PCP–Ru(II) pincer (10). Silicon-centered spirocycles with different backbones were obtained in moderate to high yields (Scheme 9). Although mechanistic studies have not yet been reported, silicon likely acts as a directing group during the C(sp3)–H activation step, according to computational studies with similar Ru complexes [64].

2.3. Osmium

Alkane Dehydrogenation

The Roddick group [65] evaluated the catalytic activity of the PCP–Os(II) pincer (11) in alkane dehydrogenation. An initial turnover frequency (TOF) value of 1520 h−1 was obtained, utilizing COA and TBE (1:1, 0.033 mol%, 200 °C, 10 min). Furthermore, activity inhibition of complex 11 was detected when the concentration of COE or H2 increases. H2 inhibition produced a dimeric compound (12) (Scheme 10). Acceptorless dehydrogenation assays of COA were made and produced trace amount of the alkynated product.

3. Group 9 Pincer Complexes

3.1. Cobalt

3.1.1. C–H Borylation

PNP–Co pincer compounds have been evaluated as catalysts in C(sp2)–H borylation catalysis by the Chirik group, contributing to several mechanistic studies. For instance, Obligacion and colleagues assessed the catalytic activity of complex 13 using substituted arenes and pyridines (Scheme 11) [66]. For pyridines, catalysis was carried out under two different conditions (conditions A and B) with 0.5 equiv and 1 equiv of B2Pin2. A higher yield was observed under condition B compared to condition A. Additionally, the byproduct obtained under condition A was H2, whereas under condition B it was HBPin. This study demonstrated that several meta-borylation products were favored by the Co(I) pincer catalyst.
Kinetics studies, including Kinetic Isotope Effect (KIE) measurements and rate law determinations, were conducted for complexes 13, 14-pyrr, and 14-Me using different substrates and borane sources (Scheme 12, Table 1) [67,68,69]. The rate laws showed no apparent dependency on the borane source. However, Pabst and colleagues identified that the byproduct HBPin exhibited a half-order behavior for complex 14-Me. In addition, complexes 14-Me and 14-pyrr showed a KIE close to one, indicating a different turnover step from C–H activation. [69]. Complex 13, on the other hand, exhibited higher KIEs with 2,6-lutidine and benzofuran, suggesting that C–H activation is likely the turnover-determining step in the catalysis. For benzofuran, a zero-order behavior of the benzofuran concentration eliminates this proposal [67,68].
To determine the catalyst resting state, C–H borylation was examined via 1H and 31P NMR spectroscopies in THF-d8 at 23 °C. The catalyses were performed with 2,6-lutidine, B2Pin2, or HBPin, employing a high loading of 13 (10 mol%) at 80 °C. Higher catalyst loading facilitated compound identification in solution. C–H borylation yields were 85% after 10 h and 19% after 32 h with B2Pin2 and HBPin, respectively. Complexes 15 and 14-BPin were identified as resting states when B2Pin2 was employed as the borane source. Stoichiometric reactions with B2Pin2 and HBPin confirmed these compounds, with no inhibitory effect detected for dinitrogen (N2) [67]. In contrast, 14-H and 14-BPin were obtained when HBPin was utilized (Scheme 13). Interestingly, a catalysis with 2-methylfuran (instead of 2,6-lutidine) only presented 14-H as the resting state. This catalysis showed a C–H borylation yield of 62% after 24 h at 23 °C in C6D6 [68]. Thus, the catalyst resting state appears to depend on the concentration of the BPin unit during the catalysis.
Based on a Co(I)–Co(III) cycle, a general mechanism is proposed for C(sp2)–H borylation with PNP–Co pincers, considering the experimental data as presented in Scheme 14. The cobalt pincer precatalyst initiates the catalytic cycle by reacting with either HBPin or B2Pin2, yielding the corresponding Co(I) hydride pincer or its BPin derivative, respectively (steps 1 or 1′). In cases where the Co(I) complex lacked para-substituents, BPin would be added at its para position [68]. The Co(I) hydride complex undergoes oxidative addition with XBPin (X = H, BPin) to form a Co(III) complex (step 2), which then eliminates HX via reductive elimination, yielding the BPin derivative (step 3). Subsequently, C(sp2)-H oxidative addition of the corresponding arene to the BPin derivative is kinetically favored, leading to an isomer where the boryl and aryl ligands are in the trans position (step 4) [70]. Following this, isomerization of the Co(III) pincer to the cis stereoisomer relative to the boryl and aryl ligands occurs through reductive elimination and oxidative addition of HBPin (step 5), ultimately resulting in the formation of the aryl boronate product via reductive elimination and regeneration of the Co(I) hydride pincer complex (step 6). A higher concentration of HBPin during catalysis directs the Co(I) hydride pincer to the resting state (step 7) or, alternatively, via rearrangement when X = H (step 8).
Selectivity for a specific borylation position in arenes is determined in the C(sp2)–H oxidative addition (Scheme 14, step 4) and depends on several factors such as steric accessibility, M–C/C–H bond thermodynamics, and chelation assistance [71]. For example, the Chirik group identified a high ortho-to-fluorine selectivity in C–H borylation with different functional groups on the arenes. In addition, they reported a four-step synthesis of the anti-inflammatory drug flurbiprofen employing the ortho-to-fluorine selective C–H borylation, which reduced the previously reported eight-step route [72]. Likewise, Pabst and coworkers compared the cobalt–carbon bond dissociation free energies (BDFEs) of nine fluoroaryl complexes (Co-ArFn, n = 0–2) versus their C–H BDFEs [69]. Their computational study showed a higher M–C stabilization as the ortho fluorine substitution increased. For 1-fluoro-3-(trifluoromethyl)benzene, a ortho/meta selectivity of 99:1 was predicted since the ortho isomer (10b) ground energy was calculated to be 2.6 kcal/mol more stable than that of the meta isomer (10a). This calculation was in good agreement with the experimental selectivity of 95:5 (ortho/meta) (Scheme 15).
The connectors between the phosphine and the pyridine ring in PNP–Co pincers were observed to be key points for their catalytic activity. For example, the Hall group demonstrated through computational studies that conformational rearrangements of these methylene connectors in PNP–Co were relevant throughout the catalytic cycle [70]. Moreover, methylation of the methylene connectors led to a decrease in the catalytic activity and selectivity. This modification increases the steric profile of the connector, thereby reducing the flexibility of the pincer [73,74]. Substituting the methylene connector with an oxygen atom has not yet yielded a stable Co pincer complex [75].
Terpyridine (NNN)–cobalt complexes were also evaluated as ligands in C(sp2)-H borylation; however, they exhibited lower yields compared to PNP pincers. To enhance catalytic activity, Léonard and colleagues added LiOMe as a base with an NNN–Co(II) pincer to control the excess HBPin concentration, forming a methoxide–borohydride adduct and avoiding an off-cycle resting state [76]. Interestingly, Pabst and coworkers employed an NNN–Co(I) pincer to favor meta-to-fluorine selective C–H borylation. For complex 18, a competition deuterium kinetic isotope experiment (KIE) was carried out with 3-fluoro-N,N-dimethylbenzenesulfonamide as the substrate (Scheme 16). The result of a KIE of 2.9(2) suggests that the C–H activation step is turnover limiting. Isomerization following C–H oxidative addition to the complex was disregarded, as borylation of fluoroaryl Co(I) pincers exclusively yielded the corresponding borylated product. Given that the ortho position is stabilized by fluorine substituents, meta selectivity would result from the kinetic product of irreversible C–H activation [77].
Bis(carbene)- and bis(silylene)pyridines Co pincer compounds have been applied as electron-rich catalysts in C–H borylation. For example, Co(I) carbenes synthesized by Roque and coworkers exhibited high catalytic activity in C–H borylation, displaying meta-to-fluorine selectivity at room temperature. However, reaction of 19 with an excess of 1,3-difluorobenzene demonstrated isomerization of the fluoroaryl complex to the ortho-isomer after 16 h (Scheme 17a), yielding the corresponding borylated product. They designed two different experimental conditions to favor either the meta-isomer (2 equiv B2Pin2, rt) or the ortho-isomer (1 equiv HBPin, 50 °C) (Scheme 17b) [78].
On the other hand, N-Heterocyclic silylene–pyridine Co pincer compounds were first used by the Cui group in C–H borylation. This Co(II) pincer catalyzed selectively meta-to-fluorine C–H borylation with a catalyst loading of 4 mol% and 8 mol% of NaBHEt3 as a reductant. In addition, 0.5 mmol of cyclohexene was added to control excess concentrations of HBPin [79]. Subsequently, Arevalo and coworkers evaluated the catalytic activity of a series of SiNSi–Co(III) pincers (20, 20-Ph). High catalyst loading was required to identify any resting state during the catalysis. For 2-methylfurane (Scheme 18a), two resting states were observed, involving the BPin–Co(III) derivative (21-Ph) and the complex bearing a σ-H2 ligand (22). For 2,6-lutidine, the resting states resembled 1-Ph and depended on the aryl substituent of the silylene fragment (Scheme 18b). With a non-substituted aryl fragment, borylation at the meta position of the catalyst was observed (21-BPin) [80]. These resting states are similar to those of the PNP pincer and, therefore, could follow the same mechanism.

3.1.2. Hydrogen Isotope Exchange (HIE)

Roque and coworkers reported a catalytic hydrogen isotope exchange (HIE) via C(sp2)–H activation employing a bis(silylene)pyridine–cobalt(III) pincer complex (21) [81]. Deuterium incorporation was achieved with benzene-d6 and a low loading of cobalt precursor 21 (0.5–5 mol %). Additionally, sterically hindered positions in arenes with methyl substituents (23a-f) were highly deuterated (>97%) (Scheme 19).
A slower rate of deuterium incorporation was observed when HBPin was added to the reaction, which suggested a resting state off-cycle. Therefore, loss of HBPin from the complex 21 was proposed to generate the cobalt(I) hydride complex, which would react via oxidative addition with benzene-d6 in the HIE catalysis [82]. In addition, different Co(III) pincer precursors were evaluated in HIE to identify the active catalyst; 3-fluoro-N,N-dimethylbenzenesulfonamide was employed as substrate (Scheme 20). Complete deuteration (>98%) was observed with complex 21 after 60 min and with complex 20 after 12 min. When benzofuranyl derivatives were utilized, a higher deuterium incorporation was detected with the monobenzofuranyl complex (24a, <5%) than that of the bisbenzofuranyl complex (24b, 78%). A lower yield with complex 24b could be because this complex would not generate the cobalt(I) hydride.

3.2. Rhodium

3.2.1. C–H Borylation

The Esteruelas group evaluated C–H borylation catalysis using a POP–Rh(I) complex (25) [83]. Stoichiometric reactions with arenes and HBPin produced the corresponding aryl and boryl derivatives, respectively. Borylation of arenes exhibited a preference for the meta position (Scheme 21). An exception was observed with fluorobenzene, where ortho position borylation was favored. This ortho-to-fluorine selectivity in Rh complexes is attributed to thermodynamic and electronic factors [84,85,86]. Catalysis was efficient with either B2Pin2 or HBPin, suggesting a catalytic cycle similar to that of Co pincers (vide supra).

3.2.2. Direct Arylation

Lewis et al. reported a catalytic direct arylation with a PNP–Rh(I) pincer (26) [87]. C–H activation in benzene readily yielded the aryl derivative at room temperature using tBuOK as the base after 16 h, which reacted with iodoarene to form the corresponding C–C coupling product. Arylation of several arenes showed functionalization at both C(sp2) and C(sp3) positions (Scheme 22). Selectivity was identified for C(sp2) and at ortho positions. This selectivity was similar to that observed in the reaction without a catalyst, which could indicate a radical pathway. Therefore, the Rh(I) would accelerate the reaction via radical intermediates.

3.3. Iridium

3.3.1. C–H Borylation

8-Hydroxyquinoline-2-carboxylic acid-derived complexes 27, 27-py, 28, and 28-py were found to be active catalysts in the borylation of arenes [88]. The catalytic activity of these complexes was evaluated using HBPin as the boron source under an argon atmosphere in neat benzene, with a catalyst loading of 5.0 mol% (Scheme 23). Conversion percentages ranged from 20% to 72%, with the highest percentages observed for the pyridine derivatives (27-py and 28-py). Complex 28-py, for instance, achieved a 71% yield of the aryl-borane product from the C–H borylation of anisole, with selectivity favoring the meta-isomer. Although mechanistic studies have not yet been conducted, the preference for the meta-position is likely associated with steric factors.
The Ozerov group [89,90] has developed catalytic C–H borylation of arenes with POCOP–Ir(III) pincers (29a-c), with the addition of hydrogen acceptors (HAs). Hydrogen acceptors were required to reduce the metal complex, increasing the borylation percentage on the arene from 4% to ca. 90%. For complex 29c, catalyst loading was reduced to 0.004 mol%, yielding a conversion percentage of 83% and a turnover frequency (TOF) of 20,750 h−1 with benzene-d6 as the substrate and ethylene as the H2 acceptor. Conversely, side reactions of HBPin with an HA mainly produced the corresponding alkane. On the other hand, side reactions of B2Pin2 and HAs led mainly to the corresponding alkane. Stoichiometric reaction of 29b with fluorobenzene resulted in the fluoroaryl derivative (31), which next reacted with 1 equiv HBPin to yield the borylated product and the dihydride iridium complex (32). A proposed mechanism (Scheme 24) begins with the formation of the 14-electron pincer intermediate via a dehydrochlorination reaction of complex 29 using a base (step 1). Subsequently, oxidative addition of the arene occurs (step 2), followed by formation of the borylated product (step 3). And the 14-electrons pincer intermediate is regenerated via alkene reduction (step 4). Reaction of 32 with H2 formed complex 33, which could act as an off-cycle step. Hence, HAs are necessary to prevent this by consuming H2 (step 5). Additionally, auxiliary off-cycle hydride/boryl redistribution equilibria from 32 were observed when reacting with a stoichiometric amount of HBPin.
Ozerov and coworkers carried out a dehydrogenative borylation of terminal alkynes (DHBTA) catalyzed with a SiNN–Ir complex (34) [91]. This reaction typically yields borylated alkanes as major products, with borylated olefins as byproducts. However, they discovered a method to selectively obtain borylated olefins by degassing and introducing 1 atm of CO into the reaction mixture after 10 min (using 4-ethynyltoluene and five equivalents of HBPin at 55 °C for 18 h). Under these conditions, a triborylated olefin was obtained with a higher yield (74%), allowing for fine-tuning of olefins through subsequent cross-coupling reactions. Additionally, substrate scope demonstrated triborylation yields up to 80% with substituted aryl and alkyl alkynes (Scheme 25). An exception was observed for Me3SiC≡CH, which stopped at the alkynylboronate stage, likely due to steric hindrance. Stoichiometric reaction of CO with 34 generated the mono- and dicarbonyl–iridium pincer, but individual catalysis with the monocarbonyl complex resulted in a lower conversion percentage (48%).

3.3.2. Alkane Dehydrogenation

Alkane dehydrogenation catalyzed with iridium complexes has been extensively studied since the first report of alkane dehydrogenation with an iridium complex by Crabtree in 1979 [92]. Since then, efforts have focused on designing new ligands to either tune or enhance their catalytic activity [93,94,95]. For instance, the Huang group synthesized various PCN–Ir pincer complexes (35–46) [96] for use as homogeneous catalysts in alkane dehydrogenation (AD) (Figure 3). These complexes feature a bulky substituent (e.g., tBu) on the P arm to prevent dimer formation, and an N-type arm that could provide favorable steric effects, aiming for asymmetric reactions [97]. Moreover, the complexes are stabilized by electronic effects arising from the push–pull interaction between N and P donor atoms.
Complexes 35–48 were evaluated as precatalysts in the transfer dehydrogenation (TD) reactions of simple alkanes (Scheme 26, Table 2). The highly effective precursors 47 and 48 (PCP-IrHCl) served as catalytic controls using cyclooctane (COA) and 2-tert-butyl-1-ethylene (TBE) as substrates (entries 8 and 9) [98,99]. The catalytically active 14-electron species PCP–Ir, along with NaCl and tert-butanol as byproducts, were generated through the dehydrochlorination of PCP–IrHCl complexes using NaOtBu as the base [100,101]. Iridium complexes 38 and 41 showed the highest catalytic activities with TOF values of 333 h−1 and 191 h−1, respectively. These complexes feature a phosphinite donor arm, characterized by its rigid structure similar to that of 47 and 48. Under the given experimental conditions, the productivity of catalyst 41 was comparable to that of the control catalysts in the catalytic assay.
On the other hand, complex 41 was also used in the transfer dehydrogenation (TD) of simple heterocycles, exhibiting outstanding activity. Heteroarenes like 2,3-dihydrobenzofuran and indole were produced with a high conversion percentage (>95%) and a low catalyst loading of 0.1–0.2 mol% (Scheme 27). On the contrary, pyridines production required a higher temperature (200 °C) and higher catalyst loading (5 mol%) to yield a conversion percentage up to 88%. Those percentages were higher than those of the catalyst of control 48 (up to 63%).
The Huang group proposed that the lower conversion percentage observed with pyridine-containing ligands could be attributed to inhibition by pyridine coordination to the metal. To investigate this, they utilized the agostic complex (46) to identify these coordination species, such as the hexacoordinated Ir complex (49). This complex showed an ortho-C(sp2)-H cyclometallation of the pyridine ring of the PCN ligand, caused by N dissociation and a subsequent C(sp2)–C(sp2) rotation, producing intermediate 50 (Scheme 28). Thus, the formation of these Ir(III) species outside the catalytic cycle could deactivate the catalyst. Therefore, the enhanced activity of complex 46 is attributed to its rigid structure, which precludes C–H activation promoted by the metal center and thereby eliminates the potential for cyclometallation.
Goldman and coworkers [102] conducted comparative studies of regioselectivity in AD using iridium pincer complexes. They combined experimental and computational approaches to elucidate the fundamental factors governing regioselectivity (Scheme 29). Experimental competition studies of AD were performed using n-alkanes and cycloalkanes, such as n-pentane and cyclododecane (CDA), with complexes 30 and 51–56. All complexes, except for 53, showed high selectivity for the dehydrogenation of n-pentane over CDA with selectivity ratios reaching up to 25:1. The main product formed was predominantly 1-pentene.
Computational comparative studies were conducted through DFT calculations based on the catalytic cycle proposed for catalysts 51 and 53 (Scheme 30). For complex 51, the calculations obtained for the electronic structure with n-hexane indicated that the rate-determining step and selectivity are governed by the β-H transfer (step 2). In addition, C–H activation (step 1) at the terminal carbon (C1) is favored over addition at the internal carbon (C2), both kinetically (ΔG‡) and thermodynamically (ΔG°) by ca. 3 kcal/mol. For complex 53, replacing the connector CH2 with O does not change the preference for C–H activation at C1. However, the rate-determining step for this complex is the olefin dissociation (step 3), instead of β-H transfer (step 2). In step 3, the loss of the olefin is favored by 2 kcal/mol to the internal olefin than the terminal olefin, which is consistent with the inversion of regioselectivity for complex 53. The 14-electron iridium species is regenerated via alkene reduction (step 4).
Goldman group [103], have also investigated the mechanism and regioselectivity of catalytic dehydrogenation of alkyl chains attached to molecules other than alkanes. For example, the formation of enamines from tertiary amines, using PCP–Ir pincer complexes. A competitive reaction was realized with COA, N,N-diisopropylethylamine (DIPEA), and PCP–Ir pincer 57 (51 precursor), showing a selectivity of 2:1 DIPEA:COA (Scheme 31a).
Kinetic isotope effect (KIE) determination was performed employing DIPEA (A: iPr2NCH2CH3) and its isotopomers (B: iPr2NCD2CD3, C: iPr2NCD2CH3, D: iPr2NCH2CD3). A kA/kB value of 7.0 revealed a strong rate-limiting step originated in the ethyl chain. KIE of the partially deuterated isotopomers showed values of 2.0 and 3.7 for kA/kC and kA/kD, respectively. The value of 2.0 for kA/kC suggests an equilibrium in the C–H activation in which there is a transfer of H from the carbon atom to the metal center, while the value of 3.7 for kA/kD would indicate the β-H elimination as the rate-determining step (r.d.s.) (Scheme 31b).
Due to the high catalytic activity presented by the Ir pincer compounds in alkane dehydrogenation, the Goldman group successfully synthesized benzene from hexane via catalysis [104]. Mechanistic studies through DFT calculations of this reaction were reported by the Sunoj group, utilizing a PCP–Ir catalyst [105]. These studies revealed that the most plausible route for benzene formation could occur through the formation of hex-1-ene by activating the C1-H bond via a σ-agostic interaction between iridium and the terminal methylene group, followed by the formation of hexa-1,3-diene and hexa-1,3,5-triene. Once these species are formed, the final step involves electrocyclization of the triene to cyclohexadiene, followed by benzene releasing a dehydroaromatization (Scheme 32).

3.3.3. Direct Alkylation

Inactive C–H bonds are predominant in organic compounds, where the activation of such bonds becomes desirable to carry out the formation of different high-value compounds. Unlike C(sp2)–H bonds, C(sp3)–H bonds are difficult to activate due to their non-acidic nature, resulting in a smaller dissociation energy. To address this issue, the use of transition metal complexes has been an efficient way to activate such bonds [106,107,108] that could be part of the metal complex as a ligand or be further functionalized [109].
The C(sp3)–H activation has also been achieved in enantioselective catalysis through functionalization via carbenoids according to the Song group [110,111] by using chiral NCN–Ir(III) catalysts (Scheme 33a). From a series of complexes, they determined that the compounds carrying out better enantioselective C–H insertion using α-aryl-α-diazoacetates with protected indoles (Scheme 33b) and various 3-diazo-oxindoles substituted with 1,4-cyclohexadiene (Scheme 33c). Catalyst 58 carried out an insertion at the C3 position of the indole through C–H activation, resulting in the formation of optically active indole derivatives with yields up to 98% and enantioselectivities (ee) ranging from moderate to good (up to 86% ee). On the other hand, catalyst 59 was also employed to carry out insertion reactions, but this time for the formation of oxindoles (Scheme 33b). The insertion of 1,4-cyclohexadiene occurs at the 3-position of the oxindole, resulting in the formation of optically active compounds with yields of 42–95% and enantiomeric excesses ranging from moderate to good (51–99% ee).
DFT calculations were employed to elucidate the mechanism of the catalyst, and based on these results, a proposed catalytic cycle is presented in Scheme 34. The reaction initiates with the dissociation of water from the catalyst, forming a 16-electron species. The energy required for water dissociation to form the trans intermediate is minimal, at 1.4 kcal/mol. This trans intermediate then interacts with 3-diazo-oxindole, resulting in the formation of an iridium carbene complex (−39.1 kcal/mol) and the expulsion of nitrogen. Further calculations indicated that the iridium carbene complex with the trans intermediate exhibits an energy gap of 25.9 kcal/mol, whereas formation via the cis intermediate encounters an energy barrier of 17.5 kcal/mol, suggesting a rearrangement favoring the trans intermediate. The intermediate carbene complex undergoes C–H insertion with 1,4-cyclohexadiene, yielding the desired organic product. Based on computational results, the authors proposed that the enantioselectivity would be originated by a hydride transfer and a C–C formation step via the Si-face approach of the cyclohexadiene to the intermediate complex [110].
Blakey and colleagues [112] synthesized a series of Ir(III) pincer complexes from phebox-type ligands, with the aim of conducting enantioselective functionalization reactions via C(sp3)–H activation (Scheme 35). Similar to the studies conducted by Song, reactions were carried out using different diazoesters and 1,4-cyclohexadiene with the Ir(III) catalyst (33). These reactions exhibited yields ranging from 71% to 99% and enantiomeric excess (%ee) values ranging from 83% to 99%. Computational results would indicate that the enantioselectivity is originated by an axial conformation of the intermediate that exposes only the Re-face to the cyclohexadiene.

3.3.4. Alkene Isomerization

Over the past couple of years, there has been a quest for the development of highly active and selective catalysts to conduct isomerization reactions, as they are of vital importance in a wide range of areas. For instance, they play a significant role in Shell higher olefin processes, the formation of aliphatic amines through internal olefins via hydroaminomethylation [113,114,115], or the synthesis of terpenoid [116,117] compounds. Due to this, the group of Chianese and colleagues [118] proposed a potential reaction mechanism in the isomerization of alkenes catalyzed with a CCC–Ir(III) pincer complex (61). Through NMR experiments and H-D cross experiments, as well as isolation and characterization via X-ray diffraction of one of these intermediates (36), they observed different reaction intermediates. The isomerization mechanism (Scheme 36) suggests that the reaction occurs in four steps: (1) Complex 61 undergoes dehydrohalogenation using NaOtBu as the base, generating a 14-electron iridium intermediate which reacts with 1-alkene to form an η2-1-alkene complex. (2) At this point, oxidative addition occurs via an allylic C–H bond, followed by rotation of the η3-allyl ligand to generate the resting state of the catalyst (62). (3) Subsequently, a reductive elimination occurs to generate an η2-2-alkene complex, and finally, (4) the 2-alkene is replaced by 1-alkene, initiating the catalytic cycle anew.

4. Group 10 Pincer Complexes

4.1. Nickel

Direct Alkylation

The use of nickel is attractive due to its comparable catalytic efficiency to other transition metals, coupled with advantages such as economic feasibility, abundance, and lower toxicity. Nickel pincer complexes have proven to be efficient catalysts in C–H alkylation [119]. Arora et al. emphasize the significance of employing pincer-type compounds with nickel in alkylation and arylation reactions (Figure 4), involving the activation of C–H bonds. They analyze various reports of NNN pincer compounds in catalyzing reactions with substrates including indoles, arenes, pyrazolyl, and carbazolyl, among others, featuring diverse functional groups such as halides, ethers, and amines (Complexes 63–66) [120,121,122,123,124,125]. While the authors do not delve into specific reaction mechanisms, they note that direct activation of C–H bonds predominates in most cases, leading to moderate to high yields across the diverse scenarios presented [119].
In 2014, Gartia et al. published their findings on the catalytic capabilities of a nickel(II) pincer NNN complex featuring the ligand N,N-bis(2,6-diisopropylphenyl)-2,6-pyridinedicarboxamido (L). They explored Grignard reagent couplings by activating C–H bonds in heterocyclic compounds such as tetrahydrofuran and furan. The nickel(II) complex demonstrated excellent activity in catalyzing C–H activation and promoting coupling with various Grignard reagents. Notably, C–H activation occurred under ambient reaction conditions with short reaction times (1–2 h) and a high turnover frequency of 4130 h−1, using a catalyst loading as low as 0.01 mol% [124]. Gartia et al. noted that the C–H activation process occurs at the 2-position relative to oxygen, facilitated by oxidative addition where the heterocyclic compound binds to Ni, positioning it for subsequent coupling with an alkyl or aryl group from the Grignard reagent, resulting in substitution at the 2-position of the heterocycle (Scheme 37).

4.2. Palladium

4.2.1. C–H Borylation

Mao et al. describe oxidative borylation catalyzed with palladium species, proposing a concerted metalation–deprotonation-type mechanism. They propose that initial oxidation of the metal species activates the C–H bond of alkenes, facilitating the formation of new C–B bonds [126]. The proposed catalytic cycle for C–H borylation with NCN–Pd is depicted in Scheme 38, outlining the mechanistic steps involved in the activation of allylic C–H bonds of alkenes for subsequent borylation through a single, concerted metalation–deprotonation process.

4.2.2. Direct Alkylation

Arumugan et al. reported the synthesis of two NNO clamp-type palladium compounds, tested as catalysts in the synthesis of carbazole derivatives. The authors proposed a reaction mechanism highlighting the role of these complexes in the different steps of the synthesis. For our focus, we address the latter part of the synthesis, where the complexes activate the N–H and C–H bonds, leading to the formation of a new five-membered ring and a C–N bond. According to the authors, the catalytic cycle continues with the deprotonation of the N–H group, forming a new palladium complex that immediately activates the adjacent phenyl C–H bond, resulting in a six-membered palladacycle. Finally, reductive elimination releases the N-acetylcarbazole and regenerates the active Pd(0) species for the next catalytic cycle (Scheme 39) [127].

5. Conclusions

C–H activation is a crucial tool in synthesis, and numerous studies have focused on optimizing and fine-tuning this process to specific requirements. As highlighted in this review, pincer complexes have been successfully employed to activate and functionalize C–H bonds, effectively targeting substrates such as alkanes, which are abundant organic compounds.
As demonstrated, C–H bond activation is a crucial step in the catalytic cycle of various processes such as C–H borylation, hydrogen isotope exchange (HIE), and alkane dehydrogenation. It determines the reaction rate and influences regio- and stereoselectivity during catalysis. Moreover, C–H activation facilitates direct alkylation or arylation, enabling direct C–C coupling without the need for a functionalized intermediate.
Various mechanistic insights into the C–H activation pathway have been proposed through the identification of resting states, kinetic isotopic effects (KIE), and computational calculations. These insights are influenced by the choice of metal pincer and substrate. However, additional investigation is necessary to validate these mechanisms and, crucially, to predict and optimize catalysis for specific applications. The perspective outlined above clearly reveals a broad range of opportunities for future developments of more efficient and selective catalysts based on pincer complexes.

Author Contributions

Writing—original draft preparation, J.S.S.-G., A.A.-F., J.R.-G., L.H.D.-R. and D.M.-M.; execution and drawing, J.S.S.-G., A.A.-F., J.R.-G., L.H.D.-R. and D.M.-M. writing—review and editing, J.S.S.-G., A.A.-F., L.G.-S. and D.M.-M.; visualization and supervision, J.S.S.-G. and D.M.-M.; funding acquisition, D.M.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the doctoral scholarship provided by Consejo Nacional de Ciencia y Tecnología (CONAHCyT)—CVU 997800, 1099979 and 1032866. CONAHCyT for the Postdoctoral Fellowships awarded under the “Estancias Posdoctorales por México 2023(1), program respectively. D.M-M would like to thank UNAM-DGAPA-PAPIIT IN223323 and CONAHCYT A1-S-033933 for their generous financial support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rogge, T.; Kaplaneris, N.; Chatani, N.; Kim, J.; Chang, S.; Punji, B.; Schafer, L.L.; Musaev, D.G.; Wencel-Delord, J.; Roberts, C.A.; et al. C–H Activation. Nat. Rev. Methods Primers 2021, 1, 43. [Google Scholar] [CrossRef]
  2. Shul’pin, G.B. C–H Functionalization: Thoroughly Tuning Ligands at a Metal Ion, a Chemist Can Greatly Enhance Catalyst’s Activity and Selectivity. Dalton Trans. 2013, 42, 12794–12818. [Google Scholar] [CrossRef]
  3. Chen, D.Y.-K.; Youn, S.W. C-H activation: A Complementary Tool in the Total Synthesis of Complex Natural Products. Chem. Eur. J. 2012, 18, 9452–9474. [Google Scholar] [CrossRef] [PubMed]
  4. Valdés, H.; Rufino-Felipe, E.; Morales-Morales, D. Pincer Complexes, Leading Characters in C–H Bond Activation Processes. Synthesis and Catalytic Applications. J. Organomet. Chem. 2019, 898, 120864. [Google Scholar] [CrossRef]
  5. Landaeta, V.R.; Rodríguez-Lugo, R.E. Insights on the C H Bond Activation by Transition Metal Complexes from Groups 8–10 Bearing (P-N) Chelates. J. Mol. Catal. A Chem. 2017, 426, 316–325. [Google Scholar] [CrossRef]
  6. Guillemard, L.; Kaplaneris, N.; Ackermann, L.; Johansson, M.J. Late-Stage C–H Functionalization Offers New Opportunities in Drug Discovery. Nat. Rev. Chem. 2021, 5, 522–545. [Google Scholar] [CrossRef] [PubMed]
  7. Gensch, T.; Hopkinson, M.N.; Glorius, F.; Wencel-Delord, J. Mild Metal-Catalyzed C–H Activation: Examples and Concepts. Chem. Soc. Rev. 2016, 45, 2900–2936. [Google Scholar] [CrossRef] [PubMed]
  8. Wencel-Delord, J.; Dröge, T.; Liu, F.; Glorius, F. Towards Mild Metal-Catalyzed C–H Bond Activation. Chem. Soc. Rev. 2011, 40, 4740–4761. [Google Scholar] [CrossRef]
  9. Rej, S.; Das, A.; Chatani, N. Strategic Evolution in Transition Metal-Catalyzed Directed C–H Bond Activation and Future Directions. Coord. Chem. Rev. 2021, 431, 213683. [Google Scholar] [CrossRef]
  10. Kasera, A.; Biswas, J.P.; Ali Alshehri, A.; Ahmed Al-Thabaiti, S.; Mokhtar, M.; Maiti, D. Transition Metal Pincer Complexes: A Series of Potential Catalysts in C H Activation Reactions. Coord. Chem. Rev. 2023, 475, 214915. [Google Scholar] [CrossRef]
  11. Ess, D.H.; Goddard, W.A.; Periana, R.A. Electrophilic, Ambiphilic, and Nucleophilic C−H Bond Activation: Understanding the Electronic Continuum of C−H Bond Activation Through Transition-State and Reaction Pathway Interaction Energy Decompositions. Organometallics 2010, 29, 6459–6472. [Google Scholar] [CrossRef]
  12. Zakis, J.M.; Smejkal, T.; Wencel-Delord, J. Cyclometallated Complexes as Catalysts for C–H Activation and Functionalization. Chem. Commun. 2022, 58, 483–490. [Google Scholar] [CrossRef] [PubMed]
  13. Najafian, A.; Cundari, T.R. Methane C–H Activation via 3d Metal Methoxide Complexes with Potentially Redox-Noninnocent Pincer Ligands: A Density Functional Theory Study. Inorg. Chem. 2017, 56, 12282–12290. [Google Scholar] [CrossRef] [PubMed]
  14. Feng, Y.; Wang, C.; Fan, H. Theoretical Studies on the Intra- and Intermolecular C–H Activation by T-Shaped Pincer Complexes. J. Organomet. Chem. 2011, 696, 4064–4069. [Google Scholar] [CrossRef]
  15. van Koten, G. Tuning the Reactivity of Metals Held in a Rigid Ligand Environment. Pure Appl. Chem. 1989, 61, 1681–1694. [Google Scholar] [CrossRef]
  16. Morales-Morales, D. Pincer Compounds: Chemistry and Applications, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 0128129328. [Google Scholar]
  17. Morales-Morales, D.; Cruz Almanza, R. Pincer Complexes. Applications in Catalysis. J. Mex. Chem. Soc. 2004, 48, 338–346. [Google Scholar] [CrossRef]
  18. Asay, M.; Morales-Morales, D. Non-Symmetric Pincer Ligands: Complexes and Applications in Catalysis. Dalton Trans. 2015, 44, 17432–17447. [Google Scholar] [CrossRef]
  19. Omae, I. Organometallic Intramolecular-Coordination Compounds Containing a Phosphorus Donor Ligand. Coord. Chem. Rev. 1980, 32, 235–271. [Google Scholar] [CrossRef]
  20. Parshall, G.W. Intramolecular Aromatic Substitution in Transition Metal Complexes. Acc. Chem. Res. 1970, 3, 139–144. [Google Scholar] [CrossRef]
  21. Peris, E.; Crabtree, R.H. Key Factors in Pincer Ligand Design. Chem. Soc. Rev. 2018, 47, 1959–1968. [Google Scholar] [CrossRef]
  22. Morales-Morales, D.; Jensen, C.G.M. The Chemistry of Pincer Compounds; Elsevier: Amsterdam, The Netherlands, 2011; ISBN 0080545157. [Google Scholar]
  23. van Koten, G. Organometallic Pincer Chemistry. Top Organomet. Chem. 2013, 40. [Google Scholar] [CrossRef]
  24. Moulton, C.J.; Shaw, B.L. Transition Metal–Carbon Bonds. Part XLII. Complexes of Nickel, Palladium, Platinum, Rhodium and Iridium with the Tridentate Ligand 2,6-Bis[(Di-t-Butylphosphino) Methyl] Phenyl. J. Chem. Soc. Dalton Trans. 1976, 1020–1024. [Google Scholar] [CrossRef]
  25. Empsall, H.D.; Hyde, E.M.; Shaw, B.L. Some Unusual Iridium-(I), -(II), and -(III) Complexes Formed from 2-Methoxyphenyl- or 2-Hydroxyphenyl-Di-t-Butylphosphine. J. Chem. Soc. Dalton Trans. 1975, 1690–1696. [Google Scholar] [CrossRef]
  26. Shaw, B.L.; Uttley, M.F. Dihydro-Platinum(II) and Related Complexes. J. Chem. Soc. Chem. Commun. 1974, 918. [Google Scholar] [CrossRef]
  27. Morales-Morales, D. Recent Applications of Phosphinite POCOP Pincer Complexes Towards Organic Transformations. Mini Rev. Org. Chem. 2008, 5, 141–152. [Google Scholar] [CrossRef]
  28. Asay, M.; Morales-Morales, D. Recent Advances on the Chemistry of POCOP–Nickel Pincer Compounds. In The Privileged Pincer-Metal Platform: Coordination Chemistry & Applications, 1st ed.; van Koten, G., Gosagge, R.A., Eds.; Springer: Cham, Switzerland, 2015; Volume 54, pp. 239–268. [Google Scholar]
  29. Serrano-Becerra, J.M.; Morales-Morales, D. Applications in Catalysis and Organic Transformations Mediated by Platinum Group PCP and PNP Aromatic-Based Pincer Complexes: Recent Advances. Curr. Org. Synth. 2009, 6, 169–192. [Google Scholar] [CrossRef]
  30. Valdés, H.; González-Sebastián, L.; Morales-Morales, D. Aromatic Para-Functionalized NCN Pincer Compounds. J. Organomet. Chem. 2017, 845, 229–257. [Google Scholar] [CrossRef]
  31. Cruz-Navarro, J.A.; Sánchez-Mora, A.; Serrano-García, J.S.; Amaya-Flórez, A.; Colorado-Peralta, R.; Reyes-Márquez, V.; Morales-Morales, D. Advances in Cross-Coupling Reactions Catalyzed by Aromatic Pincer Complexes Based on Earth-Abundant 3d Metals (Mn, Fe, Co, Ni, Cu). Catalysts 2024, 14, 69. [Google Scholar] [CrossRef]
  32. Morales-Morales, D.; Redón, R.; Yung, C.; Jensen, C.M. Dehydrogenation of Alkanes Catalyzed by an Iridium Phosphinito PCP Pincer Complex. Inorganica Chim. Acta 2004, 357, 2953–2956. [Google Scholar] [CrossRef]
  33. Valdés, H.; García-Eleno, M.A.; Canseco-Gonzalez, D.; Morales-Morales, D. Recent Advances in Catalysis with Transition-Metal Pincer Compounds. ChemCatChem 2018, 10, 3136–3172. [Google Scholar] [CrossRef]
  34. Gensch, T.; James, M.J.; Dalton, T.; Glorius, F. Increasing Catalyst Efficiency in C−H Activation Catalysis. Angew. Chem. Int. Ed. 2018, 57, 2296–2306. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, I.F.; Wilson, J.W.; Hartwig, J.F. Transition-Metal-Catalyzed Silylation and Borylation of C–H Bonds for the Synthesis and Functionalization of Complex Molecules. Chem. Rev. 2023, 123, 11619–11663. [Google Scholar] [CrossRef]
  36. Hartwig, J.F. Borylation and Silylation of C–H Bonds: A Platform for Diverse C–H Bond Functionalizations. Acc. Chem. Res. 2012, 45, 864–873. [Google Scholar] [CrossRef]
  37. Arevalo, R.; Chirik, P.J. Enabling Two-Electron Pathways with Iron and Cobalt: From Ligand Design to Catalytic Applications. J. Am. Chem. Soc. 2019, 141, 9106–9123. [Google Scholar] [CrossRef]
  38. Kato, T.; Kuriyama, S.; Nakajima, K.; Nishibayashi, Y. Catalytic C−H Borylation Using Iron Complexes Bearing 4,5,6,7-Tetrahydroisoindol-2-ide-Based PNP-Type Pincer Ligand. Chem. Asian J. 2019, 14, 2097–2101. [Google Scholar] [CrossRef]
  39. Kamitani, M.; Kusaka, H.; Yuge, H. Iron-Catalyzed Versatile and Efficient C(sp2)-H Borylation. Chem. Lett. 2019, 48, 898–901. [Google Scholar] [CrossRef]
  40. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M. Deuterium- and Tritium-Labelled Compounds: Applications in the Life Sciences. Angew. Chem. Int. Ed. 2018, 57, 1758–1784. [Google Scholar] [CrossRef]
  41. Kang, Q.-K.; Shi, H. Catalytic Hydrogen Isotope Exchange Reactions in Late-Stage Functionalization. Synlett 2022, 33, 329–338. [Google Scholar] [CrossRef]
  42. Prakash, G.; Paul, N.; Oliver, G.A.; Werz, D.B.; Maiti, D. C–H Deuteration of Organic Compounds and Potential Drug Candidates. Chem. Soc. Rev. 2022, 51, 3123–3163. [Google Scholar] [CrossRef]
  43. Atzrodt, J.; Derdau, V.; Kerr, W.J.; Reid, M. C−H Functionalisation for Hydrogen Isotope Exchange. Angew. Chem. Int. Ed. 2018, 57, 3022–3047. [Google Scholar] [CrossRef]
  44. Yang, H.; Hesk, D. Base Metal-catalyzed Hydrogen Isotope Exchange. J. Label. Comp. Radiopharm. 2020, 63, 296–307. [Google Scholar] [CrossRef] [PubMed]
  45. Kopf, S.; Bourriquen, F.; Li, W.; Neumann, H.; Junge, K.; Beller, M. Recent Developments for the Deuterium and Tritium Labeling of Organic Molecules. Chem. Rev. 2022, 122, 6634–6718. [Google Scholar] [CrossRef]
  46. Sattler, A. Hydrogen/Deuterium (H/D) Exchange Catalysis in Alkanes. ACS Catal. 2018, 8, 2296–2312. [Google Scholar] [CrossRef]
  47. Pony Yu, R.; Hesk, D.; Rivera, N.; Pelczer, I.; Chirik, P.J. Iron-Catalysed Tritiation of Pharmaceuticals. Nature 2016, 529, 195–199. [Google Scholar] [CrossRef]
  48. Yu, R.P.; Darmon, J.M.; Semproni, S.P.; Turner, Z.R.; Chirik, P.J. Synthesis of Iron Hydride Complexes Relevant to Hydrogen Isotope Exchange in Pharmaceuticals. Organometallics 2017, 36, 4341–4343. [Google Scholar] [CrossRef]
  49. Corpas, J.; Viereck, P.; Chirik, P.J. C(sp2)–H Activation with Pyridine Dicarbene Iron Dialkyl Complexes: Hydrogen Isotope Exchange of Arenes Using Benzene-d6 as a Deuterium Source. ACS Catal. 2020, 10, 8640–8647. [Google Scholar] [CrossRef]
  50. Garhwal, S.; Kaushansky, A.; Fridman, N.; Shimon, L.J.W.; Ruiter, G. De Facile H/D Exchange at (Hetero)Aromatic Hydrocarbons Catalyzed by a Stable Trans-Dihydride N-Heterocyclic Carbene (NHC) Iron Complex. J. Am. Chem. Soc. 2020, 142, 17131–17139. [Google Scholar] [CrossRef]
  51. Hashiguchi, B.G.; Young, K.J.H.; Yousufuddin, M.; Goddard, W.A.; Periana, R.A. Acceleration of Nucleophilic CH Activation by Strongly Basic Solvents. J. Am. Chem. Soc. 2010, 132, 12542–12545. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, Y.; Yao, W.; Fang, H.; Hu, A.; Huang, Z. Catalytic Alkane Dehydrogenations. Sci. Bull. 2015, 60, 1316–1331. [Google Scholar] [CrossRef]
  53. Kumar, A.; Bhatti, T.M.; Goldman, A.S. Dehydrogenation of Alkanes and Aliphatic Groups by Pincer-Ligated Metal Complexes. Chem. Rev. 2017, 117, 12357–12384. [Google Scholar] [CrossRef]
  54. Kumar, A.; Zhou, T.; Emge, T.J.; Mironov, O.; Saxton, R.J.; Krogh-Jespersen, K.; Goldman, A.S. Dehydrogenation of n-Alkanes by Solid-Phase Molecular Pincer-Iridium Catalysts. High Yields of α-Olefin Product. J. Am. Chem. Soc. 2015, 137, 9894–9911. [Google Scholar] [CrossRef] [PubMed]
  55. Shada, A.D.R.; Miller, A.J.M.; Emge, T.J.; Goldman, A.S. Catalytic Dehydrogenation of Alkanes by PCP-Pincer Iridium Complexes Using Proton and Electron Acceptors. ACS Catal. 2021, 11, 3009–3016. [Google Scholar] [CrossRef]
  56. Kovalenko, O.O.; Wendt, O.F. An Electron Poor Iridium Pincer Complex for Catalytic Alkane Dehydrogenation. Dalton Trans. 2016, 45, 15963–15969. [Google Scholar] [CrossRef] [PubMed]
  57. Zhou, X.; Malakar, S.; Zhou, T.; Murugesan, S.; Huang, C.; Emge, T.J.; Krogh-Jespersen, K.; Goldman, A.S. Catalytic Alkane Transfer Dehydrogenation by PSP-Pincer-Ligated Ruthenium. Deactivation of an Extremely Reactive Fragment by Formation of Allyl Hydride Complexes. ACS Catal. 2019, 9, 4072–4083. [Google Scholar] [CrossRef]
  58. Cherney, A.H.; Kadunce, N.T.; Reisman, S.E. Enantioselective and Enantiospecific Transition-Metal-Catalyzed Cross-Coupling Reactions of Organometallic Reagents to Construct C–C Bonds. Chem. Rev. 2015, 115, 9587–9652. [Google Scholar] [CrossRef]
  59. Rossi, R.; Bellina, F.; Lessi, M. Alkenylation Reactions of Heteroarenes by Transition-Metal Catalysts. Synthesis 2010, 24, 4131–4153. [Google Scholar] [CrossRef]
  60. Messaoudi, S.; Brion, J.; Alami, M. Transition-Metal-Catalyzed Direct C–H Alkenylation, Alkynylation, Benzylation, and Alkylation of (Hetero)Arenes. Eur. J. Org. Chem. 2010, 2010, 6495–6516. [Google Scholar] [CrossRef]
  61. Cai, H.; Tu, Y.-Q.; Lu, K.; Chen, Q.-L.; Zhang, F.-M.; Zhang, X.-M.; Pan, Y.-J.; Yan, Z.-B. Aromatic and Olefinic C-H Alkenylation by Catalysis with Spirocyclic NHC Ru(IV) Pincer Complex. Sci. China Chem. 2023, 66, 2791–2796. [Google Scholar] [CrossRef]
  62. De Sarkar, S.; Liu, W.; Kozhushkov, S.I.; Ackermann, L. Weakly Coordinating Directing Groups for Ruthenium(II)-Catalyzed C-H Activation. Adv. Synth. Catal. 2014, 356, 1461–1479. [Google Scholar] [CrossRef]
  63. Wang, Z.; Fang, H.; Liu, G.; Huang, Z. Ruthenium-Catalyzed Dual Dehydrogenative Silylation of C(sp3)–H Bonds: Access to Diverse Silicon-Centered Spirocycles. Org. Lett. 2021, 23, 7603–7607. [Google Scholar] [CrossRef]
  64. Xu, Z.; Huang, W.-S.; Zhang, J.; Xu, L.-W. Recent Advances in Transition-Metal-Catalyzed Silylations of Arenes with Hydrosilanes: C–X Bond Cleavage or C–H Bond Activation Synchronized with Si–H Bond Activation. Synthesis 2015, 47, 3645–3668. [Google Scholar] [CrossRef]
  65. Gruver, B.C.; Adams, J.J.; Arulsamy, N.; Roddick, D.M. Acceptor Pincer Chemistry of Osmium: Catalytic Alkane Dehydrogenation by (CF3PCP)Os(Cod)(H). Organometallics 2013, 32, 6468–6475. [Google Scholar] [CrossRef]
  66. Obligacion, J.V.; Semproni, S.P.; Chirik, P.J. Cobalt-Catalyzed C–H Borylation. J. Am. Chem. Soc. 2014, 136, 4133–4136. [Google Scholar] [CrossRef]
  67. Obligacion, J.V.; Semproni, S.P.; Pappas, I.; Chirik, P.J. Cobalt-Catalyzed C(sp2)-H Borylation: Mechanistic Insights Inspire Catalyst Design. J. Am. Chem. Soc. 2016, 138, 10645–10653. [Google Scholar] [CrossRef] [PubMed]
  68. Obligacion, J.V.; Chirik, P.J. Mechanistic Studies of Cobalt-Catalyzed C(sp2)–H Borylation of Five-Membered Heteroarenes with Pinacolborane. ACS Catal. 2017, 7, 4366–4371. [Google Scholar] [CrossRef]
  69. Pabst, T.P.; Obligacion, J.V.; Rochette, É.; Pappas, I.; Chirik, P.J. Cobalt-Catalyzed Borylation of Fluorinated Arenes: Thermodynamic Control of C(sp2)-H Oxidative Addition Results in Ortho-to-Fluorine Selectivity. J. Am. Chem. Soc. 2019, 141, 15378–15389. [Google Scholar] [CrossRef] [PubMed]
  70. Li, H.; Obligacion, J.V.; Chirik, P.J.; Hall, M.B. Cobalt Pincer Complexes in Catalytic C–H Borylation: The Pincer Ligand Flips Rather Than Dearomatizes. ACS Catal. 2018, 8, 10606–10618. [Google Scholar] [CrossRef]
  71. Pabst, T.P.; Chirik, P.J. A Tutorial on Selectivity Determination in C(sp2)–H Oxidative Addition of Arenes by Transition Metal Complexes. Organometallics 2021, 40, 813–831. [Google Scholar] [CrossRef]
  72. Obligacion, J.V.; Bezdek, M.J.; Chirik, P.J. C(sp2)–H Borylation of Fluorinated Arenes Using an Air-Stable Cobalt Precatalyst: Electronically Enhanced Site Selectivity Enables Synthetic Opportunities. J. Am. Chem. Soc. 2017, 139, 2825–2832. [Google Scholar] [CrossRef] [PubMed]
  73. Lee, B.; Pabst, T.P.; Hierlmeier, G.; Chirik, P.J. Exploring the Effect of Pincer Rigidity on Oxidative Addition Reactions with Cobalt(I) Complexes. Organometallics 2023, 42, 708–718. [Google Scholar] [CrossRef]
  74. Lee, B.; Pabst, T.P.; Chirik, P.J. Effect of Pincer Methylation on the Selectivity and Activity in (PNP)Cobalt-Catalyzed C(sp2)–H Borylation. Organometallics 2021, 40, 3766–3774. [Google Scholar] [CrossRef] [PubMed]
  75. Obligacion, J.V.; Zhong, H.; Chirik, P.J. Insights into Activation of Cobalt Pre-Catalysts for C(sp2)−H Functionalization. Isr. J. Chem. 2017, 57, 1032–1036. [Google Scholar] [CrossRef] [PubMed]
  76. Léonard, N.G.; Bezdek, M.J.; Chirik, P.J. Cobalt-Catalyzed C(sp2)–H Borylation with an Air-Stable, Readily Prepared Terpyridine Cobalt(II) Bis(Acetate) Precatalyst. Organometallics 2017, 36, 142–150. [Google Scholar] [CrossRef]
  77. Pabst, T.P.; Chirik, P.J. Development of Cobalt Catalysts for the Meta-Selective C(sp2)–H Borylation of Fluorinated Arenes. J. Am. Chem. Soc. 2022, 144, 6465–6474. [Google Scholar] [CrossRef] [PubMed]
  78. Roque, J.B.; Shimozono, A.M.; Pabst, T.P.; Hierlmeier, G.; Peterson, P.O.; Chirik, P.J. Kinetic and Thermodynamic Control of C(sp2)–H Activation Enables Site-Selective Borylation. Science 2023, 382, 1165–1170. [Google Scholar] [CrossRef] [PubMed]
  79. Ren, H.; Zhou, Y.; Bai, Y.; Cui, C.; Driess, M. Cobalt-Catalyzed Regioselective Borylation of Arenes: N-Heterocyclic Silylene as an Electron Donor in the Metal-Mediated Activation of C−H Bonds. Chem. Eur. J. 2017, 23, 5663–5667. [Google Scholar] [CrossRef] [PubMed]
  80. Arevalo, R.; Pabst, T.P.; Chirik, P.J. C(sp2)–H Borylation of Heterocycles by Well-Defined Bis(Silylene)Pyridine Cobalt(III) Precatalysts: Pincer Modification, C(sp2)–H Activation, and Catalytically Relevant Intermediates. Organometallics 2020, 39, 2763–2773. [Google Scholar] [CrossRef] [PubMed]
  81. Roque, J.B.; Pabst, T.P.; Chirik, P.J. C(sp2)–H Activation with Bis(Silylene)Pyridine Cobalt(III) Complexes: Catalytic Hydrogen Isotope Exchange of Sterically Hindered C–H Bonds. ACS Catal. 2022, 12, 8877–8885. [Google Scholar] [CrossRef]
  82. Semproni, S.P.; Hojilla Atienza, C.C.; Chirik, P.J. Oxidative Addition and C–H Activation Chemistry with a PNP Pincer-Ligated Cobalt Complex. Chem. Sci. 2014, 5, 1956–1960. [Google Scholar] [CrossRef]
  83. Esteruelas, M.A.; Oliván, M.; Vélez, A. POP–Rhodium-Promoted C–H and B–H Bond Activation and C–B Bond Formation. Organometallics 2015, 34, 1911–1924. [Google Scholar] [CrossRef]
  84. Tanabe, T.; Brennessel, W.W.; Clot, E.; Eisenstein, O.; Jones, W.D. Synthesis, Structure, and Reductive Elimination in the Series Tp′Rh(PR3)(ArF)H.; Determination of Rhodium–Carbon Bond Energies of Fluoroaryl Substituents. Dalton Trans. 2010, 39, 10495–10509. [Google Scholar] [CrossRef] [PubMed]
  85. Budiman, Y.P.; Perutz, R.N.; Steel, P.G.; Radius, U.; Marder, T.B. Applications of Transition Metal-Catalyzed Ortho-Fluorine-Directed C–H Functionalization of (Poly)Fluoroarenes in Organic Synthesis. Chem. Rev. 2024, 124, 4822–4862. [Google Scholar] [CrossRef] [PubMed]
  86. Kalläne, S.I.; Teltewskoi, M.; Braun, T.; Braun, B. C–H and C–F Bond Activations at a Rhodium(I) Boryl Complex: Reaction Steps for the Catalytic Borylation of Fluorinated Aromatics. Organometallics 2015, 34, 1156–1169. [Google Scholar] [CrossRef]
  87. Gair, J.J.; Qiu, Y.; Chan, N.H.; Filatov, A.S.; Lewis, J.C. Rhodium Complexes of 2,6-Bis(Dialkylphosphinomethyl)Pyridines: Improved C–H Activation, Expanded Reaction Scope, and Catalytic Direct Arylation. Organometallics 2017, 36, 4699–4706. [Google Scholar] [CrossRef]
  88. Nguyen, D.H.; Pérez-Torrente, J.J.; Jiménez, M.V.; Modrego, F.J.; Gómez-Bautista, D.; Lahoz, F.J.; Oro, L.A. Unsaturated Iridium(III) Complexes Supported by a Quinolato–Carboxylato ONO Pincer-Type Ligand: Synthesis, Reactivity, and Catalytic C–H Functionalization. Organometallics 2013, 32, 6918–6930. [Google Scholar] [CrossRef]
  89. Press, L.P.; Kosanovich, A.J.; McCulloch, B.J.; Ozerov, O.V. High-Turnover Aromatic C-H Borylation Catalyzed by POCOP-Type Pincer Complexes of Iridium. J. Am. Chem. Soc. 2016, 138, 9487–9497. [Google Scholar] [CrossRef] [PubMed]
  90. Hung, M.U.; Press, L.P.; Bhuvanesh, N.; Ozerov, O.V. Examination of a Series of Ir and Rh PXL Pincer Complexes as (Pre)Catalysts for Aromatic C-H Borylation. Organometallics 2021, 40, 1004–1013. [Google Scholar] [CrossRef]
  91. Lee, C.-I.; Zhou, J.; Reinbenspies, J.H.; Ozerov, O.V. Synthesis of Triborylalkenes from Terminal Alkynes by Iridium-Catalyzed Tandem C-H Borylation and Diboration. Angew. Chem. Int. Ed. 2015, 54, 14003–14007. [Google Scholar] [CrossRef]
  92. Choi, J.; MacArthur, A.H.R.; Brookhart, M.; Goldman, A.S. Dehydrogenation and Related Reactions Catalyzed by Iridium Pincer Complexes. Chem. Rev. 2011, 111, 1761–1779. [Google Scholar] [CrossRef]
  93. Shi, Y.; Suguri, T.; Dohi, C.; Yamada, H.; Kojima, S.; Yamamoto, Y. Highly Active Catalysts for the Transfer Dehydrogenation of Alkanes: Synthesis and Application of Novel 7-6-7 Ring-Based Pincer Iridium Complexes. Chem. Eur. J. 2013, 19, 10672–10689. [Google Scholar] [CrossRef]
  94. Zuo, W.; Braunstein, P. N-Heterocyclic Dicarbene Iridium(III) Pincer Complexes Featuring Mixed NHC/Abnormal NHC Ligands and Their Applications in the Transfer Dehydrogenation of Cyclooctane. Organometallics 2012, 31, 2606–2615. [Google Scholar] [CrossRef]
  95. Zhang, X.; Wang, D.Y.; Emge, T.J.; Goldman, A.S. Dehydrogenation of Ketones by Pincer-Ligated Iridium: Formation and Reactivity of Novel Enone Complexes. Inorganica Chim. Acta 2011, 369, 253–259. [Google Scholar] [CrossRef]
  96. Wang, Y.; Huang, Z.; Liu, G.; Huang, Z. A New Paradigm in Pincer Iridium Chemistry: PCN Complexes for (De)Hydrogenation Catalysis and Beyond. Acc. Chem. Res. 2022, 55, 2148–2161. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, X.; Wu, S.-B.; Leng, X.; Chung, L.W.; Liu, G.; Huang, Z. N-Bridged Pincer Iridium Complexes for Highly Efficient Alkane Dehydrogenation and the Relevant Linker Effects. ACS Catal. 2020, 10, 6475–6487. [Google Scholar] [CrossRef]
  98. Yao, W.; Zhang, Y.; Jia, X.; Huang, Z. Selective Catalytic Transfer Dehydrogenation of Alkanes and Heterocycles by an Iridium Pincer Complex. Angew. Chem. Int. Ed. 2014, 53, 1390–1394. [Google Scholar] [CrossRef] [PubMed]
  99. Yao, W.; Jia, X.; Leng, X.; Goldman, A.S.; Brookhart, M.; Huang, Z. Catalytic Alkane Transfer-Dehydrogenation by PSCOP Iridium Pincer Complexes. Polyhedron 2016, 116, 12–19. [Google Scholar] [CrossRef]
  100. Göttker-Schnetmann, I.; Brookhart, M. Mechanistic Studies of the Transfer Dehydrogenation of Cyclooctane Catalyzed by Iridium Bis(Phosphinite) p-XPCP Pincer Complexes. J. Am. Chem. Soc. 2004, 126, 9330–9338. [Google Scholar] [CrossRef] [PubMed]
  101. Göttker-Schnetmann, I.; White, P.S.; Brookhart, M. Synthesis and Properties of Iridium Bis(Phosphinite) Pincer Complexes (p-XPCP)IrH2, (p-XPCP)Ir(CO), (p-XPCP)Ir(H)(Aryl), and {(p-XPCP)Ir}2{μ-N2} and Their Relevance in Alkane Transfer Dehydrogenation. Organometallics 2004, 23, 1766–1776. [Google Scholar] [CrossRef]
  102. Biswas, S.; Blessent, M.J.; Gordon, B.M.; Zhou, T.; Malakar, S.; Wang, D.Y.; Krogh-Jespersen, K.; Goldman, A.S. Origin of Regioselectivity in the Dehydrogenation of Alkanes by Pincer-Iridium Complexes: A Combined Experimental and Computational Study. ACS Catal. 2021, 11, 12038–12051. [Google Scholar] [CrossRef]
  103. Lu, Y.J.; Zhang, X.; Malakar, S.; Krogh-Jespersen, K.; Hasanayn, F.; Goldman, A.S. Formation of Enamines via Catalytic Dehydrogenation by Pincer-Iridium Complexes. J. Org. Chem. 2020, 85, 3020–3028. [Google Scholar] [CrossRef]
  104. Ahuja, R.; Punji, B.; Findlater, M.; Supplee, C.; Schinski, W.; Brookhart, M.; Goldman, A.S. Catalytic Dehydroaromatization of N-Alkanes by Pincer-Ligated Iridium Complexes. Nat. Chem. 2011, 3, 167–171. [Google Scholar] [CrossRef] [PubMed]
  105. Thawani, A.; Rajeev, R.; Sunoj, R.B. On the Mechanism of the Dehydroaromatization of Hexane to Benzene by an Iridium Pincer Catalyst. Chem. Eur. J. 2013, 19, 4069–4077. [Google Scholar] [CrossRef] [PubMed]
  106. Laviska, D.A.; Guan, C.; Emge, T.J.; Wilklow-Marnell, M.; Brennessel, W.W.; Jones, W.D.; Krogh-Jespersen, K.; Goldman, A.S. Addition of C-C and C-H Bonds by Pincer-Iridium Complexes: A Combined Experimental and Computational Study. Dalton Trans. 2014, 43, 16354–16365. [Google Scholar] [CrossRef] [PubMed]
  107. Rubashkin, S.B.; Chu, W.Y.; Goldberg, K.I. Lowering the Barrier to C−H Activation at Ir(III) through Pincer Ligand Design. Organometallics 2021, 40, 1296–1302. [Google Scholar] [CrossRef]
  108. Wilklow-Marnell, M.; Li, B.; Zhou, T.; Krogh-Jespersen, K.; Brennessel, W.W.; Emge, T.J.; Goldman, A.S.; Jones, W.D. Catalytic Dehydrogenative C-C Coupling by a Pincer-Ligated Iridium Complex. J. Am. Chem. Soc. 2017, 139, 8977–8989. [Google Scholar] [CrossRef] [PubMed]
  109. Ahlstrand, D.A.; Polukeev, A.V.; Marcos, R.; Ahlquist, M.S.G.; Wendt, O.F. Csp3−H Activation without Chelation Assistance in an Iridium Pincer Complex Forming Cyclometallated Products. Chem. Eur. J. 2017, 23, 1748–1751. [Google Scholar] [CrossRef]
  110. Li, N.; Yang, X.; Zhu, Y.; Wang, F.; Gong, J.; Song, M. Pincer Iridium(III)-Catalyzed Enantioselective C(sp3)-H Functionalization via Carbenoid C-H Insertion of 3-Diazooxindoles with 1,4-Cyclohexadiene. Chin. Chem. Lett. 2022, 33, 2437–2441. [Google Scholar] [CrossRef]
  111. Li, N.; Zhu, W.J.; Huang, J.J.; Hao, X.Q.; Gong, J.F.; Song, M.P. Chiral NCN Pincer Iridium(III) Complexes with Bis(Imidazolinyl)Phenyl Ligands: Synthesis and Application in Enantioselective C-H Functionalization of Indoles with α-Aryl-α-Diazoacetates. Organometallics 2020, 39, 2222–2234. [Google Scholar] [CrossRef]
  112. Owens, C.P.; Varela-Álvarez, A.; Boyarskikh, V.; Musaev, D.G.; Davies, H.M.L.; Blakey, S.B. Iridium(III)-Bis(Oxazolinyl)Phenyl Catalysts for Enantioselective C-H Functionalization. Chem. Sci. 2013, 4, 2590–2596. [Google Scholar] [CrossRef]
  113. Yang, J.; Delolo, F.G.; Spannenberg, A.; Jackstell, R.; Beller, M. A Selective and General Cobalt-Catalyzed Hydroaminomethylation of Olefins to Amines. Angew. Chem. Int. Ed. 2022, 61, e202112597. [Google Scholar] [CrossRef] [PubMed]
  114. Migliorini, F.; Monciatti, E.; Romagnoli, G.; Parisi, M.L.; Taubert, J.; Vogt, M.; Langer, R.; Petricci, E. Switching Mechanistic Pathways by Micellar Catalysis: A Highly Selective Rhodium Catalyst for the Hydroaminomethylation of Olefins with Anilines in Water. ACS Catal. 2023, 13, 2702–2714. [Google Scholar] [CrossRef]
  115. Rätze, K.H.G.; Kortuz, W.; Kirschtowski, S.; Jokiel, M.; Hamel, C.; Sundmacher, K. Optimal Experimental Design for the Identification of a Reaction Kinetic Model for the Hydroaminomethylation of 1-Decene in a Thermomorphic Multiphase System. Chem. Eng. J. 2023, 469, 143713. [Google Scholar] [CrossRef]
  116. Delolo, F.G.; Vieira, G.M.; Avendaño-Villarreal, J.A.; de Oliveira Dias, A.; dos Santos, E.N.; Gusevskaya, E.V. Working Together to Avoid Unwanted Reactions: Hydroformylation/O-Acylation of Terpene-Based Hydroxyolefins. J. Catal. 2023, 421, 20–29. [Google Scholar] [CrossRef]
  117. Vemuri, G.N.; Hughes, J.R.; Iovine, P.M. Synthesis and Characterization of Terpene-Derived Cationic Bolaamphiphiles. Tetrahedron 2022, 124, 133008. [Google Scholar] [CrossRef]
  118. Knapp, S.M.M.; Shaner, S.E.; Kim, D.; Shopov, D.Y.; Tendler, J.A.; Pudalov, D.M.; Chianese, A.R. Mechanistic Studies of Alkene Isomerization Catalyzed by CCC-Pincer Complexes of Iridium. Organometallics 2014, 33, 473–484. [Google Scholar] [CrossRef]
  119. Arora, V.; Narjinari, H.; Nandi, P.G.; Kumar, A. Recent Advances in Pincer–Nickel Catalyzed Reactions. Dalton Trans. 2021, 50, 3394–3428. [Google Scholar] [CrossRef] [PubMed]
  120. Patel, U.N.; Pandey, D.K.; Gonnade, R.G.; Punji, B. Synthesis of Quinoline-Based NNN-Pincer Nickel(II) Complexes: A Robust and Improved Catalyst System for C–H Bond Alkylation of Azoles with Alkyl Halides. Organometallics 2016, 35, 1785–1793. [Google Scholar] [CrossRef]
  121. Patel, U.N.; Jain, S.; Pandey, D.K.; Gonnade, R.G.; Vanka, K.; Punji, B. Mechanistic Aspects of Pincer Nickel(II)-Catalyzed C–H Bond Alkylation of Azoles with Alkyl Halides. Organometallics 2018, 37, 1017–1025. [Google Scholar] [CrossRef]
  122. Soni, V.; Jagtap, R.A.; Gonnade, R.G.; Punji, B. Unified Strategy for Nickel-Catalyzed C-2 Alkylation of Indoles through Chelation Assistance. ACS Catal. 2016, 6, 5666–5672. [Google Scholar] [CrossRef]
  123. Jagtap, R.A.; Soni, V.; Punji, B. Expeditious and Solvent-Free Nickel-Catalyzed C−H Arylation of Arenes and Indoles. ChemSusChem 2017, 10, 2242–2248. [Google Scholar] [CrossRef]
  124. Gartia, Y.; Ramidi, P.; Jones, D.E.; Pulla, S.; Ghosh, A. Nickel Complex Catalyzed Efficient Activation of Sp3 and Sp2 C–H Bonds for Alkylation and Arylation of Oxygen Containing Heterocyclic Molecules. Catal. Lett. 2014, 144, 507–515. [Google Scholar] [CrossRef]
  125. Mandapati, P.; Braun, J.D.; Sidhu, B.K.; Wilson, G.; Herbert, D.E. Catalytic C–H Bond Alkylation of Azoles with Alkyl Halides Mediated by Nickel(II) Complexes of Phenanthridine-Based N^N^-N Pincer Ligands. Organometallics 2020, 39, 1989–1997. [Google Scholar] [CrossRef]
  126. Mao, L.; Bertermann, R.; Rachor, S.G.; Szabó, K.J.; Marder, T.B. Palladium-Catalyzed Oxidative Borylation of Allylic C–H Bonds in Alkenes. Org. Lett. 2017, 19, 6590–6593. [Google Scholar] [CrossRef] [PubMed]
  127. Arumugam, V.; Kaminsky, W.; Nallasamy, D. Pd(II) Pincer Type Complex Catalyzed Tandem C–H and N–H Activation of Acetanilide in Aqueous Media: A Concise Access to Functionalized Carbazoles in a Single Step. Green Chem. 2016, 18, 3295–3301. [Google Scholar] [CrossRef]
Figure 1. General structure of pincer compounds (a) and some examples (b).
Figure 1. General structure of pincer compounds (a) and some examples (b).
Inorganics 12 00221 g001
Scheme 1. Reactions to explore reactivity of PNP–Fe pincer in C–H borylation. Pin = pinacolate.
Scheme 1. Reactions to explore reactivity of PNP–Fe pincer in C–H borylation. Pin = pinacolate.
Inorganics 12 00221 sch001
Scheme 2. Proposed catalytic cycle for C–H borylation with NNP–Fe pincers.
Scheme 2. Proposed catalytic cycle for C–H borylation with NNP–Fe pincers.
Inorganics 12 00221 sch002
Figure 2. Deuterium and tritium labelling via HIE catalysis with a CNC–Fe pincer catalyst.
Figure 2. Deuterium and tritium labelling via HIE catalysis with a CNC–Fe pincer catalyst.
Inorganics 12 00221 g002
Scheme 3. Activation of CNC–Fe pincer with H2.
Scheme 3. Activation of CNC–Fe pincer with H2.
Inorganics 12 00221 sch003
Scheme 4. Proposed mechanism for HIE with an CNC–Fe pincer.
Scheme 4. Proposed mechanism for HIE with an CNC–Fe pincer.
Inorganics 12 00221 sch004
Scheme 5. Sigma-complex-assisted metathesis (σ-CAM) in HIE with a PCP–Fe(II) pincer.
Scheme 5. Sigma-complex-assisted metathesis (σ-CAM) in HIE with a PCP–Fe(II) pincer.
Inorganics 12 00221 sch005
Scheme 6. Nucleophilic C–H activation with NNN–Ru pincer.
Scheme 6. Nucleophilic C–H activation with NNN–Ru pincer.
Inorganics 12 00221 sch006
Scheme 7. Alkane dehydrogenation catalyzed with a PSP–Ru pincer.
Scheme 7. Alkane dehydrogenation catalyzed with a PSP–Ru pincer.
Inorganics 12 00221 sch007
Scheme 8. Proposed catalytic cycle of direct alkenylation with a CCC–Ru(IV) pincer.
Scheme 8. Proposed catalytic cycle of direct alkenylation with a CCC–Ru(IV) pincer.
Inorganics 12 00221 sch008
Scheme 9. Dehydrogenative silylation with a PCP–Ru(II) pincer.
Scheme 9. Dehydrogenative silylation with a PCP–Ru(II) pincer.
Inorganics 12 00221 sch009
Scheme 10. Deactivation reaction of PCP–Os(II) via H2.
Scheme 10. Deactivation reaction of PCP–Os(II) via H2.
Inorganics 12 00221 sch010
Scheme 11. C–H borylation with complex 13 as catalyst. Conditions A: 0.5 equiv B2Pin2, B: 1 equiv B2Pin2; in THF. Condition C: 0.05 equiv B2Pin2; solvent-free (neat). Numbers in parenthesis indicate the position of the borylation. Yields are the conversion percentage of B2Pin2.
Scheme 11. C–H borylation with complex 13 as catalyst. Conditions A: 0.5 equiv B2Pin2, B: 1 equiv B2Pin2; in THF. Condition C: 0.05 equiv B2Pin2; solvent-free (neat). Numbers in parenthesis indicate the position of the borylation. Yields are the conversion percentage of B2Pin2.
Inorganics 12 00221 sch011
Scheme 12. C–H borylation with PNP–Co pincers.
Scheme 12. C–H borylation with PNP–Co pincers.
Inorganics 12 00221 sch012
Scheme 13. Catalyst resting states observed via 1H and 31P RMN. (2,6-lutidine, 10 mol% 13, THF-d8, 80 °C).
Scheme 13. Catalyst resting states observed via 1H and 31P RMN. (2,6-lutidine, 10 mol% 13, THF-d8, 80 °C).
Inorganics 12 00221 sch013
Scheme 14. C–H borylation proposed with PNP–Co pincer complexes.
Scheme 14. C–H borylation proposed with PNP–Co pincer complexes.
Inorganics 12 00221 sch014
Scheme 15. DFT calculations of C–H activation and isomerization in C–H borylation.
Scheme 15. DFT calculations of C–H activation and isomerization in C–H borylation.
Inorganics 12 00221 sch015
Scheme 16. Competition deuterium kinetic isotope experiment (KIE) in C–H borylation with NNN-Co(I) pincer (18).
Scheme 16. Competition deuterium kinetic isotope experiment (KIE) in C–H borylation with NNN-Co(I) pincer (18).
Inorganics 12 00221 sch016
Scheme 17. C–H borylation with CNC–Co(I) pincer complex. (a) Reaction of 19 with 1,3-difluorobenzene and its isomerization. (b) Switchable C–H borylation.
Scheme 17. C–H borylation with CNC–Co(I) pincer complex. (a) Reaction of 19 with 1,3-difluorobenzene and its isomerization. (b) Switchable C–H borylation.
Inorganics 12 00221 sch017
Scheme 18. Catalyst resting states in C–H borylation with SiNSi–Co pincer complexes using 2-methylfurane (a) and 2,6-lutidine (b) as substrates.
Scheme 18. Catalyst resting states in C–H borylation with SiNSi–Co pincer complexes using 2-methylfurane (a) and 2,6-lutidine (b) as substrates.
Inorganics 12 00221 sch018
Scheme 19. Catalytic hydrogen isotope exchange (HIE) of arenes with 21 and benzene-d6.
Scheme 19. Catalytic hydrogen isotope exchange (HIE) of arenes with 21 and benzene-d6.
Inorganics 12 00221 sch019
Scheme 20. HIE catalysis with different cobalt(III) pincer precursors. [C6D6] = 0.125 M for 20 and 21 and 0.25 M for 24a and 24b.
Scheme 20. HIE catalysis with different cobalt(III) pincer precursors. [C6D6] = 0.125 M for 20 and 21 and 0.25 M for 24a and 24b.
Inorganics 12 00221 sch020
Scheme 21. C–H borylation of arenes with a POP–Rh(I) pincer.
Scheme 21. C–H borylation of arenes with a POP–Rh(I) pincer.
Inorganics 12 00221 sch021
Scheme 22. Direct arylation with a PNP–Rh(I) pincer.
Scheme 22. Direct arylation with a PNP–Rh(I) pincer.
Inorganics 12 00221 sch022
Scheme 23. C–H borylation or arenes with an ONO-Ir(III) pincer.
Scheme 23. C–H borylation or arenes with an ONO-Ir(III) pincer.
Inorganics 12 00221 sch023
Scheme 24. Proposed mechanism for C–H borylation catalyzed with a PCP–Ir(III) pincer.
Scheme 24. Proposed mechanism for C–H borylation catalyzed with a PCP–Ir(III) pincer.
Inorganics 12 00221 sch024
Scheme 25. Substrate scope of triborylation of alkynes with an NNSi–Ir pincer.
Scheme 25. Substrate scope of triborylation of alkynes with an NNSi–Ir pincer.
Inorganics 12 00221 sch025
Figure 3. PCN–Ir-type catalysts employed as homogeneous catalysts in alkane dehydrogenation.
Figure 3. PCN–Ir-type catalysts employed as homogeneous catalysts in alkane dehydrogenation.
Inorganics 12 00221 g003
Scheme 26. General reaction for transfer dehydrogenation of COA catalyzed with PCN–Ir.
Scheme 26. General reaction for transfer dehydrogenation of COA catalyzed with PCN–Ir.
Inorganics 12 00221 sch026
Scheme 27. Transfer dehydrogenation (TD) of heterocycles catalyzed with Ir pincers.
Scheme 27. Transfer dehydrogenation (TD) of heterocycles catalyzed with Ir pincers.
Inorganics 12 00221 sch027
Scheme 28. ortho-C(sp2)-H cyclometallation of complex 46.
Scheme 28. ortho-C(sp2)-H cyclometallation of complex 46.
Inorganics 12 00221 sch028
Scheme 29. Study of regioselectivity. (a) Iridium pincer catalysts used for the study of regioselectivity. (b) Dehydrogenation of cyclododecane/n-pentane using iridium pincer catalysts.
Scheme 29. Study of regioselectivity. (a) Iridium pincer catalysts used for the study of regioselectivity. (b) Dehydrogenation of cyclododecane/n-pentane using iridium pincer catalysts.
Inorganics 12 00221 sch029
Scheme 30. Catalytic dehydrogenation cycle proposed for PCP–Ir pincer catalysts.
Scheme 30. Catalytic dehydrogenation cycle proposed for PCP–Ir pincer catalysts.
Inorganics 12 00221 sch030
Scheme 31. Formation of enamines via catalytic dehydrogenation. (a) Competitive study of COA/DIPEA. (b) Dehydrogenation pathway of amines catalyzed with PCP–Ir pincers.
Scheme 31. Formation of enamines via catalytic dehydrogenation. (a) Competitive study of COA/DIPEA. (b) Dehydrogenation pathway of amines catalyzed with PCP–Ir pincers.
Inorganics 12 00221 sch031
Scheme 32. Proposed mechanism for benzene formation from hexane using a PCP–Ir pincer.
Scheme 32. Proposed mechanism for benzene formation from hexane using a PCP–Ir pincer.
Inorganics 12 00221 sch032
Scheme 33. Enantioselective C–H activation. (a) Chiral NCN–Ir(III) pincer catalysts. Enantioselective direct alkylation with (b) α-aryl-α-diazoacetates with protected indoles and (c) 3-diazo-oxindoles substituted with 1,4-cyclohexadiene, employing NCN–Ir(III) pincer catalysts.
Scheme 33. Enantioselective C–H activation. (a) Chiral NCN–Ir(III) pincer catalysts. Enantioselective direct alkylation with (b) α-aryl-α-diazoacetates with protected indoles and (c) 3-diazo-oxindoles substituted with 1,4-cyclohexadiene, employing NCN–Ir(III) pincer catalysts.
Inorganics 12 00221 sch033
Scheme 34. Proposed reaction mechanism for enantioselective alkyl catalysis. Pincer ligand is omitted in Complex 59 for clarity.
Scheme 34. Proposed reaction mechanism for enantioselective alkyl catalysis. Pincer ligand is omitted in Complex 59 for clarity.
Inorganics 12 00221 sch034
Scheme 35. Enantioselective functionalization reactions of diazoesters catalyzed with pincer complex 60.
Scheme 35. Enantioselective functionalization reactions of diazoesters catalyzed with pincer complex 60.
Inorganics 12 00221 sch035
Scheme 36. Proposed catalytic cycle for isomerization using CCC–Ir(III) pincer complex (61) as catalyst.
Scheme 36. Proposed catalytic cycle for isomerization using CCC–Ir(III) pincer complex (61) as catalyst.
Inorganics 12 00221 sch036
Figure 4. NNN–Ni Pincer compounds.
Figure 4. NNN–Ni Pincer compounds.
Inorganics 12 00221 g004
Scheme 37. Direct alkylation with an NNN–Ni pincer.
Scheme 37. Direct alkylation with an NNN–Ni pincer.
Inorganics 12 00221 sch037
Scheme 38. Proposed catalytic cycle for C–H borylation with NCN–Pd.
Scheme 38. Proposed catalytic cycle for C–H borylation with NCN–Pd.
Inorganics 12 00221 sch038
Scheme 39. Carbazoles formation mechanism using NCN–Pd pincers.
Scheme 39. Carbazoles formation mechanism using NCN–Pd pincers.
Inorganics 12 00221 sch039
Table 1. Kinetic Isotope Effect (KIE) and rate law determination for C(sp2)–H borylation with Co pincer compounds.
Table 1. Kinetic Isotope Effect (KIE) and rate law determination for C(sp2)–H borylation with Co pincer compounds.
Cat. SubstrateExperimental
Conditions
KIERate LawRef.
13Inorganics 12 00221 i0013 mol%, B2Pin2, 80 °C2.9(1)rate=kobs[13][substrate]1[B2Pin2]0[67]
14-pyrrInorganics 12 00221 i0023 mol%, B2Pin2, 80 °C1.6(1)rate=kobs[14-pyrr][substrate]1[B2Pin2]0[67]
13Inorganics 12 00221 i0031 mol%, HBPin, 80 °C1.9(1)rate=kobs[13][substrate]0[HBPin]0[68]
14-MeInorganics 12 00221 i00410 mol%, B2Pin2, 50 °C1.1(1), 0.9(1) *rate=kobs[14-Me][substrate]0[B2Pin2]0 [HBPin]0.5[69]
* ortho and meta borylation, respectively.
Table 2. Transfer dehydrogenation of COA/TBE catalyzed with PCN–Ir.
Table 2. Transfer dehydrogenation of COA/TBE catalyzed with PCN–Ir.
EntryCat.Cat. Loading (mol%)T (°C)TOF (h−1) 1
1350.215026
2360.21503
3370.21502
4380.2150333
5390.21506
6400.21506
7410.025200191
847 20.033200448
948 20.017200328
1 t = 1 h (35–40), 18 h (41), 4 h (47), 15 h (48). 2 Catalysts used as control.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Serrano-García, J.S.; Amaya-Flórez, A.; R.-Galindo, J.; González-Sebastián, L.; Delgado-Rangel, L.H.; Morales-Morales, D. C–H Activation via Group 8–10 Pincer Complexes: A Mechanistic Approach. Inorganics 2024, 12, 221. https://doi.org/10.3390/inorganics12080221

AMA Style

Serrano-García JS, Amaya-Flórez A, R.-Galindo J, González-Sebastián L, Delgado-Rangel LH, Morales-Morales D. C–H Activation via Group 8–10 Pincer Complexes: A Mechanistic Approach. Inorganics. 2024; 12(8):221. https://doi.org/10.3390/inorganics12080221

Chicago/Turabian Style

Serrano-García, Juan S., Andrés Amaya-Flórez, Jordi R.-Galindo, Lucero González-Sebastián, Luis Humberto Delgado-Rangel, and David Morales-Morales. 2024. "C–H Activation via Group 8–10 Pincer Complexes: A Mechanistic Approach" Inorganics 12, no. 8: 221. https://doi.org/10.3390/inorganics12080221

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop