Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2021 Nov 16;60(23):18490–18502. doi: 10.1021/acs.inorgchem.1c03056

Ammonia–Borane Dehydrogenation Catalyzed by Dual-Mode Proton-Responsive Ir-CNNH Complexes

Isabel Ortega-Lepe , Andrea Rossin ‡,*, Práxedes Sánchez , Laura L Santos , Nuria Rendón , Eleuterio Álvarez , Joaquín López-Serrano , Andrés Suárez †,*
PMCID: PMC8653221  PMID: 34784204

Abstract

graphic file with name ic1c03056_0013.jpg

Metal complexes incorporating proton-responsive ligands have been proved to be superior catalysts in reactions involving the H2 molecule. In this contribution, a series of IrIII complexes based on lutidine-derived CNNH pincers containing N-heterocyclic carbene and secondary amino NHR [R = Ph (4a), tBu (4b), benzyl (4c)] donors as flanking groups have been synthesized and tested in the dehydrogenation of ammonia–borane (NH3BH3, AB) in the presence of substoichiometric amounts (2.5 equiv) of tBuOK. These preactivated derivatives are efficient catalysts in AB dehydrogenation in THF at room temperature, albeit significantly different reaction rates were observed. Thus, by using 0.4 mol % of 4a, 1.0 equiv of H2 per mole of AB was released in 8.5 min (turnover frequency (TOF50%) = 1875 h–1), while complexes 4b and 4c (0.8 mol %) exhibited lower catalytic activities (TOF50% = 55–60 h–1). 4a is currently the best performing IrIII homogeneous catalyst for AB dehydrogenation. Kinetic rate measurements show a zero-order dependence with respect to AB, and first order with the catalyst in the dehydrogenation with 4a (−d[AB]/dt = k[4a]). Conversely, the reaction with 4b is second order in AB and first order in the catalyst (−d[AB]/dt = k[4b][AB]2). Moreover, the reactions of the derivatives 4a and 4b with an excess of tBuOK (2.5 equiv) have been analyzed through NMR spectroscopy. For the former precursor, formation of the iridate 5 was observed as a result of a double deprotonation at the amine and the NHC pincer arm. In marked contrast, in the case of 4b, a monodeprotonated (at the pincer NHC-arm) species 6 is observed upon reaction with tBuOK. Complex 6 is capable of activating H2 reversibly to yield the trihydride derivative 7. Finally, DFT calculations of the first AB dehydrogenation step catalyzed by 5 has been performed at the DFT//MN15 level of theory in order to get information on the predominant metal–ligand cooperation mode.

Short abstract

Iridium complexes based on CNNH ligands containing two potential proton-responsive sites—a lutidine scaffold and a secondary amino group—have been tested in the dehydrogenation of ammonia-borane. Upon reaction with base, depending on the amino group acidity, mono- or doubly deprotonated species exhibiting significantly different catalytic activities were observed.

Introduction

The controlled release of H2 from high content hydrogen compounds used as H2-storage systems is paramount to the use of this gas as an energy vector in the so-called Hydrogen Economy.1,2 In this context, ammonia–borane (NH3BH3, AB) has received increasing attention as a hydrogen storage material because of its high available hydrogen content (19.6 wt % H), moisture kinetic stability and easy H2 thermal release.3 In addition, AB dehydrogenation renders B–N oligomers and polymers as byproducts that are suitable starting materials for the synthesis of boron nitride (BN) ceramics.4 Although AB dehydrogenation can be performed under simple thermal conditions,5 insufficient kinetic control of the process leads to the formation of ill-defined B–N containing solids. This is a serious drawback for the H2-depleted material recycling. As a consequence, catalytic approaches to the H2 release from AB have been envisaged. The use of a catalyst allows for a fine control of the rate and extent of H2 production, as well as the nature of the H2-depleted byproducts. In addition, the H2 release temperature is lowered, with respect to pure AB, and the process can be performed with a lower energetic impact.

Catalysts based on transition metals have provided fast kinetics and a large extent of H2 release under relatively mild reaction conditions.6 Particularly, metal complexes stabilized by pincer ligands have been widely investigated because of the enhanced catalyst stability provided by the κ3–tridentate ligand coordination.716 Moreover, since ammonia–borane is a polar molecule with hydridic (BH) and acidic (NH) hydrogen atoms, transition-metal complexes based on ligands containing Brønsted basic sites have been shown to be particularly adequate to promote H2 release from AB through the transfer of the hydridic BH to the acidic metal center and the NH proton to the basic ligand functionality (bifunctional catalysts).1319 Prevalent examples of such systems are metal complexes based on ligands bearing secondary amine donors,20 which are capable of getting involved in reversible metal-amine/metal-amido interconversion (Figure 1a).1315,17,18

Figure 1.

Figure 1

Proposed mechanisms for the first step of ammonia–borane dehydrogenation with transition-metal complexes based on ligands containing Brønsted basic sites.

Similarly, metal complexes incorporating picoline- or lutidine-derived pincer ligands are able to participate in ligand-assisted H–X (X = H, C, O, N, B) bond activation upon deprotonation of the (acidic) CH2 pincer side-arms and concomitant dearomatization of the N-containing ring (Figure 1b).21 Although these derivatives have been found to be efficient catalysts in a plethora of hydrogenation reactions of polar organic substrates and alcohol dehydrogenation reactions, their application in AB dehydrogenation has been limitedly explored.16 Moreover, development of homogeneous catalytic systems based on lutidine-derived ligands has been mainly focused on phosphine-containing PNX (X = phosphine or N-donor) pincers, albeit substitution of the flanking PR2 groups by another strong σ-donor such as N-heterocyclic carbenes (NHCs) has also been briefly addressed. Upon deprotonation of the methylene CH2–NHC arms, these organometallics are catalytically active in ligand-assisted processes.2225

A step forward in catalysts design that has provided superior catalytic systems in hydrogenation and dehydrogenation reactions relies on the use of ligands containing two Brønsted acidic/basic sites, such as a lutidine fragment and a secondary amino group.25,26 Since two Brønsted functionalities are present in these ligands, the related complexes might participate in ligand-assisted processes through pyridine aromatization/dearomatization or amine-metal/amido-metal interconversion. Herein, we report on the synthesis of a series of Ir complexes based on lutidine-derived CNNH ligands bearing secondary amino and NHC side-arms and their catalytic performance in AB dehydrogenation. Detailed kinetic and NMR spectroscopy studies show that different species are formed under catalytic conditions, depending on the substituent of the secondary amino group. Moreover, DFT calculations have determined the preferred metal–ligand cooperation mode when a doubly deprotonated catalytic species is formed.

Results and Discussion

Synthesis and Structure of Ir-CNNH Complexes

Following previously reported procedures for the synthesis of Ir complexes based on related NHC-containing lutidine-derived ligands,23,24 the diolefin derivatives 2a and 2b were prepared by the reactions of Ir(acac)(COD) and the N-heterocyclic carbene ligand precursors 1a and 1b,25 respectively (Scheme 1). These complexes were isolated as yellow solids in good yields (99% and 77%, respectively), and spectroscopical and analytically characterized. In the 1H NMR spectrum of 2a, the olefinic hydrogens of one of the C=C moieties of the COD ligand appear as multiplets at δ 4.51 and 4.41 ppm, whereas the resonances corresponding to the other olefin fragment are shown at δ 2.95 and 2.82 ppm. Moreover, the 1H–1H EXSY experiment of 2a recorded at 298 K exhibits intense cross-peaks corresponding to the pairwise exchange of the olefinic signals corresponding to different C=C moieties. The observed dynamic behavior can be assigned to alkene site exchange involving the decoordination of one of the olefin fragments to produce a distorted tetrahedral intermediate, and subsequent recoordination of the uncoordinated C=C group to the opposite side (see the Supporting Information).23 Meanwhile, the C-2 of the NHC ligand fragment produces a singlet at δC 181.0 ppm in the 13C{1H} NMR spectrum. Similar NMR data and fluxional behavior were shown by complex 2b.

Scheme 1. Synthesis of the Ir-CNNH Complexes 2a, 2b, 3, and 4a4c.

Scheme 1

Conversely, heating a CH2Cl2 solution of Ir(acac)(COD) and the imidazolium salt 1c to 50 °C produced the expected coordination of the lutidine-derived CNNH pincer ligand along with the activation of one ortho C–H bond of the benzyl substituent, yielding the κ4-(CNHC,NPy,Namine,Caryl) complex 3 (Scheme 1). The solid-state structure of this derivative was determined by single-crystal X-ray diffraction (Figure 2). The complex is comprised of a stereogenic Ir center in an octahedral coordination geometry, in which the carbene carbon and the nitrogen donors of the CNNH ligand are coordinated in the meridional positions, as defined by the CNHC–Ir-Namine angle of 171.0°, and the metalated aryl fragment is located trans to the Br atom (Caryl–Ir-Br = 169.1°).

Figure 2.

Figure 2

ORTEP drawing at 30% ellipsoid probability of complex 3. Hydrogen atoms (except for the hydride ligand and the NHC and NH hydrogens) have been omitted for clarity. Selected bond lengths: Ir(2)–C(27), 1.959(10) Å; Ir(2)–N(7), 2.124(7) Å; Ir(2)–N(8), 2.132(8) Å; Ir(2)–Br(2), 2.6266(11) Å; and Ir(2)–C(48), 2.010(10) Å. Selected bond angles: C(27)–Ir(2)–N(8), 171.0(3)°; C(27)–Ir(2)–C(48), 97.4(4)°; C(27)–Ir(2)–N(7), 91.4(3)°; N(8)–Ir(2)–N(7), 79.7(3)°; and C(48)–Ir(2)–Br(2), 169.1(3)°.

Complexes 2a and 2b reacted with H2 (5 bar) in CH2Cl2, resulting in the formation of the bromodihydride complexes 4a and 4b, respectively (Scheme 1). Similarly, derivative 4c was obtained by exposing a CH2Cl2 solution of 3 at 60 °C to 4 bar of H2. Complexes 4a4c, which were isolated as air-stable yellow solids, were characterized analytically and spectroscopically. For example, in the 1H NMR spectrum of complex 4a, two mutually coupled doublets are detected in the hydride region, appearing at −19.05 and −23.25 ppm (2JHH = 6.9 Hz), which are assigned to the IrH hydrogens located trans and cis to the pyridine fragment, respectively, as determined by 1H–1H NOESY spectroscopy. In the same experiment, the NH hydrogen produces a broad doublet resonance at 6.15 ppm (3JHH = 11.1 Hz). In the 13C{1H} NMR spectrum, diagnostic signals for complex 4a are shown at 153.5 ppm, caused by the carbene carbon, and at 62.2 and 55.3 ppm, corresponding to the methylene NH- and NHC-arms of the pincer, respectively. Analogous NMR data were observed in the case of complexes 4b and 4c, with the notable exception of the resonances due to the NH groups in the 1H NMR spectra, which are shifted upfield, with respect to that of 4a, appearing as a broad doublet of doublets at 4.03 ppm (3JHH = 12.5 Hz, 3JHH = 3.2 Hz) for 4b, and 4.39 ppm (3JHH = 9.2 Hz, 3JHH = 9.2 Hz) in the case of 4c.

The solid-state structure of 4b, as determined by single-crystal X-ray diffraction, exhibits an octahedral geometry with the CNNH pincer adopting a meridional coordination and the two hydride ligands in cis to each other (Figure 3). Moreover, the existence of a hydrogen bond between the bromine ligand and the amino hydrogen is evident from the H···Br distance of 2.76 Å, which is shorter than the sum of the van der Waals radii of H and Br (2.97 Å).27

Figure 3.

Figure 3

ORTEP drawing at 30% ellipsoid probability of complex 4b. Hydrogen atoms (except for the hydride ligands and the NHC and NH hydrogens) have been omitted for clarity. Selected bond lengths: Ir(1)–C(1), 1.939(3) Å; Ir(1)–N(3), 2.116(3) Å; Ir(1)–N(4), 2.184(3) Å; Ir(1)–Br(1), 2.6385(4) Å; Ir(1)-H(1Ir), 1.569(18) Å; and Ir(1)–H(2Ir), 1.591(18) Å. Selected bond angles: C(1)–Ir(1)–N(4), 169.62(12)°; C(1)–Ir(1)–N(3), 91.83(12)°; N(3)–Ir(1)–Br(1), 88.86(8)°; and N(4)–Ir(1)–N(3), 78.04(10).

AB Dehydrogenation

Complexes 4a4c were tested as catalysts in the dehydrogenation of ammonia borane after a preliminary activation with tBuOK (see the Supporting Information for the experimental setup). The initial treatment with a base is necessary to trigger the catalysis; bare complex 4a (without a base) was tested in AB dehydrogenation, and it was found to be a poor catalyst (initial turnover frequency (TOF) of <5 h–1). Upon addition of a THF solution of 4a (0.4 mol %) and tBuOK (base:4a = 2.5) to a stirred AB solution in the same solvent, instantaneous vigorous gas evolution was observed, followed by the immediate formation of an off-white precipitate. Follow-up of the reaction by measuring the pressure increase in the system showed that 1.0 equiv of H2 per AB molecule was released in ca. 8.5 min at room temperature (Figure 4a). To the best of our knowledge, the high room-temperature catalytic activity of 4a (TOF = 1764 h–1) is only surpassed by the cationic Pd complexes [Pd(allyl)][BF4], [Pd(allyl)(2,4-hexadiene)][BF4] and [Pd(MeCN)4][BF4]2 reported by the group of Michalak in 2010,28 and by the Ru-PNHP and RuCl2(PN)2 derivatives reported by the Schneider group in 200913 and the Fagnou group in 2008,17 respectively. Note that the catalytic activity of 4a is slightly superior to that of the most active Ir-based catalyst reported to date: IrH2(POCOP), published by Goldberg and colleagues in 2006.7 Thus, 4a is currently the best-performing iridium-containing homogeneous catalyst for AB dehydrogenation. On the other hand, reactions performed with complexes 4b and 4c, using 0.8 mol % of catalyst loading under the same reaction conditions, yielded ca. 0.75–0.8 equiv of H2 in 5 h (for 4b, TOF50% = 55 h–1; for 4c, TOF50% = 60 h–1); thus, both catalysts exhibit significantly lower catalytic activity than 4a (Figure 4b).

Figure 4.

Figure 4

H2 evolution in the catalytic dehydrogenation of AB with (a) 4a (0.4 mol %) and (b) 4b and 4c (0.8 mol %). Conditions: THF, room temperature, tBuOK:4 = 2.5, [AB] = 1.6 M.

The registered ATR-IR spectrum of the insoluble H2-depleted material produced in the AB dehydrogenation reactions with 4a (amounting to >95% of the total BN byproducts, as estimated by mass balance) contains peaks corresponding to the N–H (3299 and 3248 cm–1) and B–H (2384 and 2313 cm–1) stretching modes (see Figure S7 in the Supporting Information). These absorptions, as well as the spectrum fingerprint, closely resemble those described for the mainly linear poly(aminoboranes), [NH2BH2]n, isolated from the dehydrogenation of AB with Brookhart́s Ir-POCOP complex,29 and Ru- and Fe-PNHP complexes.13,14 The 11B NMR spectrum of the soluble part of the reaction with 4a exhibits broad signals between δB 17–31 ppm, in agreement with the formation of species resulting from the release of more than 1 equiv of H2 (see Figure S8 in the Supporting Information). However, the small fraction of these products is not enough to account for an overall H2 yield exceeding 1.0 equiv.

The observed differences in the catalytic activity provided by 4a and complexes 4b and 4c led us to determine the reaction kinetic laws. For 4a, plotting of [AB]o–[AB] versus time gave a straight line (kobs = 3.4 × 10–3 M s–1), indicative of a pseudo-zero order relationship in AB (see Figures S10 and S11 in the Supporting Information). Initial rate experiments performed at different catalyst concentrations showed that hydrogen release has a first-order rate dependence in 4a (Figure S12 in the Supporting Information). The observed lack of dependence on AB concentration is in agreement with its rapid reaction with the catalyst and, therefore, with a rate-determining step that does not involve AB activation.

Next, kinetic isotope effects (KIE) values were determined using the deuterated AB isotopologues (NH3BD3, ND3BH3, and ND3BD3). A pronounced decrease in the reaction rate was observed with the N-deuterated substrate [ND3BH3, kH/kD = 8.4], while deuteration at boron produces a somewhat lower rate [NH3BD3, kH/kD = 1.4]. In the case of full AB deuteration, within the experimental error, the KIE is the product of the individual isotope effects measured with ND3BH3 and NH3BD3 [ND3BD3, kH/kD = 13.4]. This suggests a concerted, asynchronous transition state in the cleavage of the N–H and B–H bonds. However, this conclusion is tentative, since the observed KIEs should reflect the contributions of all the multiple steps of the catalytic cycle. We must bear in mind that AB activation does not seem to be the rate-determining step.30

The kinetic law was also inferred for the reaction with the catalyst precursor 4b. From the initial rate experiments performed at different AB and catalyst concentrations, a kinetic lawd[AB]/dt = kobs[AB]2 = k[4b][AB]2 (kobs = 1.3 × 10–4 M–1 s–1) was revealed, evidencing different rate laws for the reactions with complexes 4a and 4b (see Figures S14 and S15 in the Supporting Information). For the latter catalyst, the KIE values using deuterated NH3BD3, ND3BH3, and ND3BD3 were found to be 2.5, 2.8, and 7.7, respectively. These data suggest a mechanism where two AB molecules are involved in the rate-limiting step,12 with the N–H and B–H bonds splitting occurring in a concerted manner.

Mechanistic Insights

Taking into account previous studies with catalysts based on proton-responsive ligands,13,14,16,17 a plausible mechanism for AB dehydrogenation catalyzed by complexes 4 in the presence of an auxiliary base should involve initial hydrogen transfer from AB to the deprotonated form of the Ir-CNNH catalyst precursors with the concomitant formation of the “inorganic ethylene analogue” BH2=NH2, followed by ligand-assisted H2 elimination. Therefore, to gain insight into the species formed upon treatment of complexes 4 with a base, derivative 4a was initially reacted with tBuOK (2.5 equiv) in THF-d8 (Scheme 2). Reaction of the Ir-CNNH derivatives with a strong base could induce deprotonation of the pincer ligand at its methylene carbon in the side arm2225 and/or at the secondary amino group.20,25,31 Analysis of the deuterated solution through NMR spectroscopy showed the formation of a major species (ca. 90%), which was characterized as the highly air-sensitive amido iridate(III) 5,32 confirming the deprotonation of both the NH and the NHC–methylene arm of 4a. The 1H NMR spectrum of 5 features two mutually coupled doublets at δ −16.35 ppm and δ −18.61 ppm (2JHH = 5.5 Hz), corresponding to the hydride ligands. In the same experiment, the methylene CH2NPh hydrogens appear as doublets at δ 4.24 and 3.80 ppm (2JHH = 16.5 Hz), while the methine CH-NHC hydrogen is a singlet at δ 6.25 ppm. In the 13C{1H} NMR spectrum, the carbene carbon produces a singlet at δC 156.8 ppm, whereas the CH2N and CH-NHC bridges of the pincer give rise to resonances at δC 64.2 and 49.0 ppm, respectively.

Scheme 2. Formation of Iridate Complex 5.

Scheme 2

Crystals suitable for X-ray diffraction (XRD) analysis of 5 were obtained by cooling the above THF-d8 solution to −30 °C (Figure 5). The X-ray data confirmed the initially proposed structure. The potassium iridate 5 is dimeric in the solid state, and the two Ir centers are bound to the opposite ligands through the deprotonated CH-NHC arms. The Ir atoms are in an octahedral coordination geometry, with the pincer and one hydride ligand occupying the meridional plane [Σ(Ir) = 361.2°] and the other IrH hydrogen and the Ir(CNN*)H2 unit coordinated at the apical positions. The K+ counterions are surrounded by two molecules of THF (d(K–O) = 2.61 Å, avg.), an η4-coordinated mesityl ring (K–C distances of 3.30–3.39 Å), one carbon of the N-phenyl ring (d(K–C) = 3.14 Å), and the meridional hydride ligand (d(K–H) = 2.85 Å). The observed dinuclear solid-state structure of 5 should be favored by the large planarity of the deprotonated Ir-CNN* fragments.

Figure 5.

Figure 5

ORTEP drawing at 30% ellipsoid probability of the dimeric form of complex 5. Hydrogen atoms (except for the hydride ligands) have been omitted for clarity. Selected bond lengths: Ir(1)–C(1), 1.999(19) Å; Ir(1)–N(3), 2.033(13) Å; Ir(1)–N(4), 2.115(15) Å; and Ir(1)–C(13), 2.215(17) Å. Selected bond angles: C(1)–Ir(1)–N(4), 165.9(7)°; C(1)–Ir(1)–N(3), 90.4(7)°; N(3)–Ir(1)–N(4), 78.5(6)°; C(1)–Ir(1)–C(13), 98.9(6)°; and N(4)–Ir(1)–C(13), 90.6(6)°.

Although in the solid state, complex 5 can be regarded as a dimer, the structure of 5 in solution is that of a monomer, as evidenced by diffusion NMR studies.33 The 1H DOSY experiment of 5 afforded a diffusion coefficient D = 7.2 × 10–10 m2/s (log D = −9.14) from which a value of its hydrodynamic radius (rH) of 6.4 Å was estimated using the Stokes–Einstein equation. The radius for the dimeric structure calculated from its X-ray volume34 was rX-ray = 13.9 Å, which is approximately twice the rH of 5 in solution. Moreover, the diffusion coefficient D of 5 is very similar to that obtained for 4a using 1H DOSY spectroscopy [D = 8.2 × 10–10 m2/s; log D = −9.09; rH = 5.5 Å]. This further supports the existence of a monomeric form of 5 in solution.

In marked contrast, under the conditions used for the formation of 5, reaction of 4b with tBuOK (2.5 equiv) in THF-d8 gave rise to complex 6 (75%–80% NMR yield), which is selectively deprotonated at the NHC-bridge, as determined through NMR spectroscopy (Scheme 3). The 1H NMR experiment of 6 shows, in the hydride region, two mutually coupled doublets appearing at −16.49 and −18.41 ppm (2JHH = 5.6 Hz). Moreover, in the same spectrum, a doublet of doublets at 3.88 ppm (2JHH = 13.6 Hz, 3JHH = 2.8 Hz) and a multiplet at 3.39 ppm, corresponding to the CH2N hydrogens, and a broad doublet at δ 2.97 ppm (2JHH = 10.9 Hz) due to the NH, are also observed; whereas, the proton of the deprotonated methine-NHC arm produces a singlet at 5.48 ppm. In the 13C{1H} NMR spectrum of 6, the carbene carbon appears as a singlet at 150.0 ppm, while the N- and NHC-linkers of the pincer produce singlet resonances at 56.1 and 54.9 ppm, respectively. Further attempts to perform the deprotonation of the amino group of 4b were unsuccessful, likely due to the reduced NH acidity, in comparison to 4a.

Scheme 3. Formation of Deprotonated Complex 6 and Trihydride Derivative 7.

Scheme 3

Because of the fast kinetics, attempts to detect metallic intermediates through NMR spectroscopy during the dehydrogenation reaction catalyzed by 4a at room temperature were unsuccessful. However, at the end of the catalytic process, the main metal-containing species present in solution (>90%, from 1H NMR analysis) is the amido iridate(III) 5. Moreover, pressurization of a THF-d8 solution of 5 with H2 (4 bar) did not lead to the formation of new observable species (i.e., a trihydride Ir complex).

For 4b, the solution composition at the end of the dehydrogenation process, performed at the NMR scale in THF-d8, is more complex than that found for 4a: the 1H NMR spectrum of the reaction mixture reveals the presence of NHC-deprotonated iridium dihydride 6 as the major component (ca. 30%), together with at least four other unidentified Ir hydrides. In addition, reaction in THF-d8 of the in situ formed complex 6 with H2 (4 bar, 48 h) produced the selective formation of trihydride derivative 7, which is only stable under a hydrogen atmosphere (Scheme 3). Complex 7 gives rise in the hydride region of the 1H NMR spectrum to a multiplet signal integrating for 2H at −8.92 ppm, corresponding to the apical IrH hydrogens, and a doublet of doublets at −18.24 ppm (2JHH = 5.2, 5.2 Hz) that integrates for 1H, ascribed to the meridional hydride ligand. Moreover, the reprotonation of the pincer ligand side arm was confirmed from the presence of two mutually coupled doublets appearing at 5.02 and 5.10 ppm (2JHH = 13.5 Hz), corresponding to the methylene–NHC pincer arm.

DFT Calculations of the First AB Dehydrogenation Step

To get further insight into the reaction catalyzed by 4a, DFT calculations on the first step of the AB dehydrogenation mechanism were performed at the DFT//MN15 level of theory on the real structure of the supposed active catalytic species 5, dimeric in the solid state (as found in the crystal structure determined from X-ray data collection) but monomeric in THF solution (as confirmed by DOSY NMR experiments). In this structure, possibly in the form of a THF adduct with general formula [(CNN*)IrH2(thf)], the solvent O atom occupies the empty coordination site trans to one of the hydride ligands, with the IrIII ion in an octahedral coordination geometry. Successive THF/AB exchange in the metal coordination sphere is thermodynamically favored, leading to the complex [(CNN*)IrH2(AB)] (I, Figure 6). For this complex, two different isomeric forms are conceivable, depending on the interaction site of AB NH3 end. While the BH3 end is always directly bound to the metal center in a η1-fashion, the NH3 end may be interacting either with the carbanionic C atom on the pincer skeleton (“C-path”) or with the negative N atom of the NPh amido side arm (“N-path”). In principle, both sites are nucleophilic and both interactions are possible. No interaction with the central pyridine N atom or with the carbene arm was found. This is the typical example of “bifunctional catalyst” bearing an acidic (IrIII) and a basic (CH/NPh) reactive site at the same time, where the metal center and the pincer ligand act cooperatively to extract hydrogen from AB. This class of compound is widespread in the literature, and it has been frequently exploited in the dehydrogenation of BN lightweight inorganic hydrides.13,14,1619 The analysis of the relative stability of the two isomers Ia (HC···H–N interaction, Figure 6a) and Ib (PhN···H–N interaction, Figure 6b) revealed that the latter is slightly more stable [ΔGTHF(Ia/Ib) = −2.9 kcal/mol]. Given the small Gibbs energy difference between the two isomers, both forms were taken into account for the calculation of the reaction profiles. Consistent with the kinetic data from the KIE experiments, a simultaneous BH/NH bond activation occurs and the B–H bond activation is less difficult than that of the N–H bond, since AB coordination to the metal ion through its BH3 end is responsible for a B–H bond “preliminary” weakening. Indeed, in I, the B–H bond is already partially broken, as witnessed by the longer B–H distance on the coordinated bond if compared with the other two.

Figure 6.

Figure 6

Optimized structure of (CNN*)IrH21-BH-AB) with (a) NH3 interaction with the carbanionic C atom (Ia) and (b) with NH3 interaction with the amido N atom (Ib). Selected optimized bond lengths are reported in Ångstroms. H atoms on the pincer ligand not relevant for the discussion are omitted for clarity. Atom color code: white, H; gray, C; blue, N; pink, B; orange, Ir.

From this initial geometry, a transition state TS1C/TS1N of the CH/NPh protonation by NH3 could be found at ΔGTHF# = 3.9/5.9 kcal/mol along the d(N–H···CH)/d(N–H···NPh) reaction coordinate (Figure 7), with related “inorganic ethylene analogue” (BH2=NH2) evolution and formation of the trihydride complex [(CNN)IrH3] (II). Thermodynamics for this step is favorable in both cases, with ΔGTHF = −9.5 and −2.4 kcal/mol for CH and NPh protonation, respectively.

Figure 7.

Figure 7

Optimized structures of (a) TS1C (from Ia) and (b) TS1N (from Ib). Selected optimized bond lengths are reported in Ångstroms. H atoms on the pincer ligand not relevant for the discussion are omitted for clarity. Bonds involved in the TS transformation depicted as red dotted lines. See Figure 6 for the atom color code.

From II (again, in two isomeric forms IIa and IIb, depending on the protonated group of choice), different reaction paths were then considered. No H2 formation (nonclassical hydride) from the (classical) hydride ligands with concomitant IrIII → IrI reduction could be observed, at the computational level used in this study. The same can be said for a direct interaction of the hydride ligands of II with BH2=NH2; all the plausible geometries built in silico led to repulsive interactions and no maxima could be located on the corresponding reaction energy scan. The only mechanism featured by the presence of a maximum along the examined reaction coordinate is the direct H2 elimination from one hydride ligand on iridium and from the methylenic −CH2– side arm (IIa) or from the NHPh group (IIb), to regenerate the starting active species 5 and close the catalytic cycle. The corresponding transition states TS2C and TS2N are located at 17.6 and 8.5 kcal/mol above IIa and IIb, respectively (Figure 8). The overall G versus reaction coordinate profiles at comparison are reported in Figure 9. From the inspection of the two profiles, we can conclude that for this complex the “N path” (red trace) is to be preferred to the “C path” (blue trace), from both a kinetic (ΔG#) and a thermodynamic (ΔG) viewpoint. In addition, the rate-determining step in both cases is H2 elimination from the trihydride intermediate II rather than the easy AB activation by the bifunctional complex I. This result is consistent with the zero-order dependence of the reaction rate from AB concentration found experimentally, since AB is not involved in the rate-determining step.

Figure 8.

Figure 8

Optimized structures of (a) TS2C (from IIa) and (b) TS2N (from IIb). Selected optimized bond lengths are reported in Ångstroms. H atoms on the pincer ligand not relevant for the discussion are omitted for clarity. Bonds involved in the TS transformation depicted in red dotted lines. See Figure 6 for the atom color code.

Figure 9.

Figure 9

Gibbs energy (THF) vs. reaction coordinate profiles for the first AB dehydrogenation step mediated by Ia (blue line) and Ib (red line).

As for 4b, the adduct of the activated neutral species 6 with two AB molecules is thermodynamically favored: ΔGTHF for the mono- and bis(AB) adducts equal −1.4 and −0.7 kcal/mol, respectively. The result is in agreement with the experimental evidence of a second-order kinetic dependence from AB concentration. This interaction leads to the geometry III (Figure S49), where the first AB molecule sits in the empty coordination site of the five-coordinated 6 interacting through its BH and NH ends with iridium and the deprotonated side arm, respectively [optimized d(Ir···H–B) = 1.88 Å; optimized d(H–C···H–N) = 2.21 Å]. The second AB molecule weakly interacts with the two hydrides on the metal center to the opposite side of the Ir(CNN) plane [optimized d(Ir–Haxial···H–B) = 2.86 Å; optimized d(Ir–Hequatorial···H–N) = 1.74 Å]. No interaction between AB and the NHtBu group is observed. Therefore, this part of the complex is not involved in the catalytic process and no “N path” can be conceived, at odds with its phenyl-substituted analogue. Unfortunately, no plausible reaction mechanism involving two AB molecules in the rate-determining step could be found, at the computational level used (see the Supporting Information).

Conclusions

The results reported herein show that highly active IrIII catalysts have been developed by using the same lutidine-derived CNNH ligand scaffold with NH amino groups of different basicity. The phenyl-substituted amino complex 4a in particular is the most active iridium-based catalyst in AB dehydrogenation to date, to the best of our knowledge. The excellent catalytic activity of 4a is attributed to the formation of a doubly deprotonated species (the amido iridate 5) upon reaction with a 2-fold excess of tBuOK under catalytic conditions. Meanwhile, double deprotonation of Ir-CNNH complexes based on N-alkyl substituted ligands like 4b cannot be achieved, likely due to the reduced acidity of their amino group. Finally, density functional theory (DFT) calculations of the mechanisms of the first H2 release from AB catalyzed by 5 reveal that, despite the presence of two Brønsted basic sites (the amido fragment and the methine-NHC arm), the most favorable metal–ligand cooperation mode is based on reversible metal-amine/metal-amido interconversion. Given the widespread use of metal complexes containing proton-responsive ligands in many catalytic processes, this approach, based on the use of ligands with two proton-responsive sites, could lead to improved catalytic systems. Further applications in catalysis of metal complexes incorporating more than one Brønsted acidic/basic sites are currently being explored in our laboratories.

Experimental Section

General Procedures

All reactions and manipulations were performed under nitrogen or argon, either in a Braun Labmaster 100 glovebox or using standard Schlenk-type techniques. All solvents were distilled under nitrogen with the following desiccants: sodium-benzophenone-ketyl for diethyl ether (Et2O) and tetrahydrofuran (THF); sodium for pentane and toluene; CaH2 for dichloromethane and acetonitrile. Imidazolium salts 1a1c25 and Ir(acac)(COD)35 were prepared as previously described. Deuterated ammonia borane adducts (NH3BD3, ND3BH3, ND3BD3) were synthesized by known methods.36 All other reagents were purchased from commercial suppliers and used as received. NMR spectra were obtained on Bruker DRX-400 and AVANCEIII/ASCEND 400R spectrometers. 11B{1H} NMR shifts were referenced to external BF3·Et2O, while 13C{1H} and 1H shifts were referenced to the residual signals of deuterated solvents. All data are reported in ppm downfield from Me4Si. All NMR measurements were performed at 25 °C, unless otherwise stated. NMR signal assignations were confirmed by 2D NMR spectroscopy (1H–1H COSY, 1H–1H NOESY, 1H–13C HSQC and 1H–13C HMBC) for all the complexes. Elemental analyses were run by the Analytical Service of the Instituto de Investigaciones Químicas in a Leco TrueSpec CHN elemental analyzer. IR spectra were acquired on a Thermo Scientific Nicolet iS5 iD7 ATR instrument.

Computational Details

Calculations were performed on the real structures of the IrIII anionic complex 5 and the IrIII neutral complex 6 with the Gaussian1637 package at the DFT//MN1538 level. Effective core potentials (ECP) and associated SDD basis set39 supplemented with f-polarization functions (SDD(f))40 were used to describe the inner electronic shells and the d valence electrons of the Ir atom. All the other atoms were described with a 6-31++G(d,p) basis set.41 The structures of the reactants and complexes were fully optimized with this basis set without any symmetry restrictions in THF (ε = 7.42), which was introduced within the SMD solvation model.42 In these optimizations, individual solvation spheres were placed on the H atoms of AB and on the hydride ligands on Ir. The full geometry optimization was followed by the thermochemistry calculations. The nature of all the stationary points on the potential energy surface was confirmed by vibrational analysis. No scaling factors were applied to the calculated frequencies. The transition-state structures showed only one negative eigenvalue in their diagonalized force constant matrices, and their associated eigenvectors were confirmed to correspond to the motion along the reaction coordinate under consideration using the Intrinsic Reaction Coordinate (IRC) method.43

Synthesis of Ir-CNNH Complexes

Complex 2a

A solution of 1a (0.105 g, 0.23 mmol) and Ir(acac)(COD) (0.090 g, 0.23 mmol) in CH2Cl2 (10 mL) was stirred for 2 days at room temperature. Volatiles were removed under reduced pressure, and the residue was successively washed with pentane (3 × 8 mL) and Et2O (8 mL), and dried under vacuum. Yellow solid (0.179 g, 99%). Anal. Calcd (%) for C33H38BrIrN4: C 51.96, H 5.02, N 7.34; found: C 51.96, H 4.93, N 7.24. 1H NMR (400 MHz, CD2Cl2): δ 7.76 (t, 3JHH = 7.6 Hz, 1H, H arom Py), 7.40 (d, 3JHH = 7.6 Hz, 1H, H arom Py), 7.34 (d, 3JHH = 7.6 Hz, 1H, H arom Py), 7.26 (d, 3JHH = 1.7 Hz, 1H, H arom NHC), 7.19 (dd, 3JHH = 7.3 Hz, 3JHH = 7.3 Hz, 2H, 2 H arom Ph), 7.08 (s, 1H, H arom Mes), 6.98 (s, 1H, H arom Mes), 6.90 (d, 3JHH = 1.7 Hz, 1H, H arom NHC), 6.71 (m, 3H, 3H arom Ph), 6.45 (d, 2JHH = 15.6 Hz, 1H, CHH-NHC), 5.58 (d, 2JHH = 15.6 Hz, 1H, CHH–NHC), 4.96 (s, 1H, NH), 4.51 (m, 1H, =CH COD), 4.47 (m, 2H, 2 CHH–NH), 4.41 (m, 1H, =CH COD), 2.95 (m, 1H, =CH COD), 2.82 (m, 1H, =CH COD), 2.39 (s, 6H, 2 CH3), 1.95 (s, 3H, CH3), 1.76 (m, 2H, 2 CHH COD), 1.56 (m, 1H, CHH COD), 1.38 (m, 4H, 4 CHH COD), 1.12 (m, 1H, CHH COD). 13C{1H} NMR (101 MHz, CD2Cl2): δ 181.0 (C2–NHC), 158.7 (Cq arom), 156.5 (Cq arom), 148.4 (Cq arom), 139.1 (Cq arom), 137.7 (CH arom), 136.9 (Cq arom), 136.3 (Cq arom), 135.1 (Cq arom), 129.7 (CH arom), 129.5 (2 CH arom), 128.5 (CH arom), 123.6 (CH arom), 122.3 (CH arom), 121.2 (CH arom), 120.9 (CH arom), 117.7 (CH arom), 113.3 (2 CH arom), 83.0 (=CH COD), 82.6 (=CH COD), 56.6 (CH2–NHC), 53.4 (=CH COD), 52.5 (=CH COD), 49.3 (CH2NH), 34.3 (CH2 COD), 32.6 (CH2 COD), 30.0 (CH2 COD), 29.3 (CH2 COD), 21.2 (CH3), 20.3 (CH3), 18.1 (CH3).

Complex 2b

A solution of 1b (0.176 g, 0.40 mmol) and Ir(acac)(COD) (0.159 g, 0.40 mmol) in CH2Cl2 (15 mL) was stirred for 24 h at room temperature. Volatiles were removed under reduced pressure, and the residue was washed with pentane (3 × 10 mL) and dried under vacuum. Yellow solid (0.229 g, 77%). Anal. Calcd (%) for C31H42BrIrN4: C 50.12, H 5.70, N 7.54; found: C 49.92, H 5.78, N 7.44. 1H NMR (400 MHz, CD2Cl2): δ 7.72 (dd, 3JHH = 7.7 Hz, 3JHH = 7.7 Hz, 1H, H arom Py), 7.34 (m, 2H, 2 H arom Py), 7.23 (d, 3JHH = 1.5 Hz, 1H, H arom NHC), 7.07 (s, 1H, H arom Mes), 6.98 (s, 1H, H arom Mes), 6.86 (d, 3JHH = 1.5 Hz, 1H, H arom NHC), 6.40 (d, 2JHH = 15.7 Hz, 1H, CHH–NHC), 5.51 (d, 2JHH = 15.7 Hz, 1H, CHH–NHC), 4.49 (m, 1H, =CH COD), 4.41 (m, 1H, =CH COD), 3.89 (m, 2H, 2 CHHNH), 3.01 (m, 1H, =CH COD), 2.81 (m, 1H, =CH COD), 2.39 (s, 6H, 2 CH3), 1.96 (br m, 4H, CH3 + NH), 1.79 (m, 3H, 3 CHH COD), 1.56 (m, 1H, CHH COD), 1.43 (m, 1H, CHH COD), 1.32 (m, 2H, 2 CHH COD), 1.18 (s, 9H, C(CH3)3), 1.12 (m, 1H, CHH COD). 13C{1H} NMR (101 MHz, CD2Cl2): δ 180.9 (C2–NHC), 161.5 (Cq arom), 156.1 (Cq arom), 139.1 (Cq arom), 137.5 (CH arom), 136.9 (Cq arom), 136.3 (Cq arom), 135.1 (Cq arom), 129.7 (CH arom), 128.5 (CH arom), 123.6 (CH arom), 122.2 (CH arom), 121.5 (CH arom), 120.7 (CH arom), 82.9 (=CH COD), 82.5 (=CH COD), 56.7 (CH2–NHC), 52.4 (2 =CH COD), 50.7 (C(CH3)3), 48.8 (CH2NH), 34.3 (CH2 COD), 32.6 (CH2 COD), 30.1 (CH2 COD), 29.3 (C(CH3)3 + CH2 COD), 21.2 (CH3), 20.3 (CH3), 18.1 (CH3).

Complex 3

A solution of Ir(acac)(COD) (0.059 g, 0.15 mmol) in CH2Cl2 (10 mL) was added to a solution of 1c (0.070 g, 0.15 mmol) in CH2Cl2 (10 mL). The mixture was stirred at room temperature for 24 h, and at 50 °C for 36 h. Solvent was removed under reduced pressure, and the residue was washed with MeCN (2 × 7 mL) and dried under vacuum. Yellow solid (0.024 g, 24%). Anal. Calcd (%) for C26H28BrIrN4: C 46.70, H 4.22, N 8.38; found: C 46.80, H 4.53, N 8.81. 1H NMR (400 MHz, CD2Cl2): δ 7.68 (dd, 3JHH = 7.9 Hz, 3JHH = 7.9 Hz, 1H, H arom Py), 7.30 (d, 3JHH = 7.9 Hz, 1H, H arom Py), 7.28 (d, 3JHH = 2.2 Hz, 1H, H arom NHC), 7.15 (d, 3JHH = 7.9 Hz, 1H, H arom Py), 6.99 (s, 1H, H arom Mes), 6.98 (d, 3JHH = 2.2 Hz, 1H, H arom NHC), 6.90 (s, 1H, H arom Mes), 6.79 (d, 3JHH = 7.7 Hz, 1H, H arom Ph), 6.72 (dd, 3JHH = 7.7 Hz, 4JHH = 1.2 Hz, 1H, H arom Ph), 6.61 (ddd, 3JHH = 7.3 Hz, 3JHH = 7.3 Hz, 4JHH = 1.2 Hz, 1H, H arom Ph), 6.53 (dd, 3JHH = 7.3 Hz, 3JHH = 7.3 Hz, 1H, H arom Ph), 5.78 (d, 2JHH = 16.0 Hz, 1H, CHH–NHC), 5.41 (d, 2JHH = 16.1 Hz, 1H, CHH–NHC), 5.26 (m, 2H, CHHNH + NH), 4.50 (dd, 2JHH = 15.3 Hz, 3JHH = 7.1 Hz, 1H, CHHN), 4.15 (d, 2JHH = 16.1 Hz, 1H, CHHNH), 3.92 (d, 2JHH = 15.3 Hz, 1H, CHHNH), 2.35 (s, 3H, CH3), 2.30 (s, 3H, CH3), 2.10 (s, 3H, CH3), −17.73 (s, 1H, IrH). 13C{1H} NMR (101 MHz, CD2Cl2): δ 162.8 (Cq arom), 161.3 (C2–NHC), 150.8 (Cqarom), 148.2 (Cqarom), 140.0 (Ir–Cqarom), 138.2 (Cqarom), 138.1 (Cqarom), 137.7 (CHarom), 137.1 (Cqarom), 136.5 (CHarom), 136.4 (Cqarom), 129.4 (CH arom), 128.6 (CH arom), 124.9 (CH arom), 121.8 (CH arom), 121.7 (CH arom), 120.9 (CH arom), 120.9 (CH arom), 120.7 (CH arom), 120.2 (CH arom), 66.5 (CH2NH), 64.2 (CH2NH), 55.3 (CH2–NHC), 21.3 (CH3), 20.1 (CH3), 19.7 (CH3).

Complex 4a

In a Fisher–Porter vessel, a solution of 2a (0.172 g, 0.22 mmol) in CH2Cl2 (8 mL) was pressurized with 5 bar of H2 and stirred for 24 h at room temperature. The system was depressurized, solvent was evaporated, and the residue was washed with pentane (2 × 8 mL) cooled to −30 °C and dried. Yellow solid (0.043 g, 29%). Anal. Calcd (%) for C25H28BrIrN4: C, 45.73; H, 4.30; N, 8.53; found: C, 45.41; H, 4.47; N, 8.21. 1H NMR (400 MHz, CD2Cl2): δ 7.86 (dd, 3JHH = 7.7 Hz, 3JHH = 7.7 Hz, 1H, H arom Py), 7.46 (dd, 3JHH = 7.2 Hz, 3JHH = 7.2 Hz, 2H, 2 H arom Ph), 7.33 (m, 4H, 4 H arom), 7.14 (m, 2H, 2 H arom), 6.96 (m, 2H, 2 H arom), 6.72 (d, 3JHH = 1.8 Hz, 1H, H arom NHC), 6.41 (d, 2JHH = 15.0 Hz, 1H, CHH–NHC), 6.15 (br d, 3JHH = 11.1 Hz, 1H, NH), 5.05 (d, 2JHH = 15.0 Hz, 1H, CHH–NHC), 4.76 (dd, 2JHH = 14.5 Hz, 3JHH = 2.8 Hz, 1H, CHH–NH), 4.62 (d, 2JHH = 14.5 Hz, 3JHH = 11.1 Hz, 1H, CHH–NH), 2.34 (s, 3H, CH3), 2.14 (s, 3H, CH3), 1.92 (s, 3H, CH3), −19.05 (d, 2JHH = 6.9 Hz, 1H, IrH trans to Py), −23.25 (d, 2JHH = 6.9 Hz, 1H, IrH cis to Py). 13C{1H} NMR (101 MHz, CD2Cl2): δ 159.8 (Cq arom), 153.5 (C2–NHC), 149.5 (Cq arom), 138.0 (Cq arom), 137.5 (Cq arom), 136.5 (Cq arom), 135.6 (CH arom), 135.0 (Cq arom), 128.9 (2 CH arom + Cq arom), 128.7 (CH arom), 128.3 (CH arom), 125.0 (CH arom), 122.8 (CH arom), 120.9 (CH arom), 120.5 (CH arom), 120.1 (2 CH arom), 119.3 (CH arom), 62.2 (CH2NH), 55.3 (CH2–NHC), 20.9 (CH3), 18.8 (CH3), 17.9 (CH3).

Complex 4b

This complex was prepared as described above for 4a. Yellow solid (0.084 g, 40%). Anal. Calcd (%) for C23H32BrIrN4: C 43.39, H 5.07, N 8.80; found: C 43.50, H 5.14, N 8.42. 1H NMR (400 MHz, CD2Cl2): δ 7.77 (dd, 3JHH = 7.7 Hz, 3JHH = 7.7 Hz, 1H, H arom Py), 7.37 (d, 3JHH = 7.7 Hz, 1H, H arom Py), 7.36 (d, 3JHH = 7.7 Hz, 1H, H arom Py), 7.12 (d, 3JHH = 2.1 Hz, 1H, H arom NHC), 7.03 (s, 1H, H arom Mes), 6.96 (s, 1H, H arom Mes), 6.71 (d, 3JHH = 2.1 Hz, 1H, H arom NHC), 6.43 (d, 2JHH = 14.9 Hz, 1H, CHH–NHC), 4.98 (d, 2JHH = 14.9 Hz, 1H, CHH–NHC), 4.53 (dd, 2JHH = 13.5 Hz, 3JHH = 3.2 Hz, 1H, CHHNH), 4.03 (br dd, 3JHH = 12.5 Hz, 3JHH = 3.2 Hz, 1H, NH), 3.91 (dd, 2JHH = 13.5 Hz, 3JHH = 12.5 Hz, 1H, CHHNH), 2.39 (s, 3H, CH3), 2.18 (s, 3H, CH3), 1.94 (s, 3H, CH3), 1.33 (s, 9H, C(CH3)3), −19.02 (d, 1H, 2JHH = 7.1 Hz, IrH trans to Py), −23.92 (d, 1H, 2JHH = 7.1 Hz, IrH cis to Py). 13C{1H} NMR (101 MHz, CD2Cl2): δ 162.4 (Cq arom), 156.9 (C2–NHC), 153.7 (Cq arom), 138.3 (Cq arom), 138.0 (Cq arom), 137.0 (Cq arom), 135.5 (Cq arom), 135.4 (CH arom), 128.9 (CH arom), 128.5 (CH arom), 122.7 (CH arom), 120.9 (CH arom), 120.6 (CH arom), 119.5 (CH arom), 56.3 (CH2NH + C(CH3)3), 55.7 (CH2–NHC), 28.8 (C(CH3)3), 21.3 (CH3), 19.2 (CH3), 18.4 (CH3).

Complex 4c

In a Fisher–Porter vessel, a suspension of complex 3 (0.040 g, 0.06 mmol) in CH2Cl2 (5 mL) was pressurized with 4 bar of H2, and stirred at 60 °C for 24 h. The system was depressurized, solvent was evaporated and the residue was washed with and pentane (2 × 8 mL), and dried. Yellow solid (0.039 g, 97%). Anal. Calcd (%) for C26H30BrIrN4: C, 46.56; H, 4.51; N, 8.35; found: C, 46.46; H, 4.84; N, 8.17. 1H NMR (400 MHz, CD2Cl2): δ 7.72 (dd, 3JHH = 7.4 Hz, 3JHH = 7.4 Hz, 1H, H arom Py), 7.38 (m, 7H, 7 H arom), 7.20 (d, 3JHH = 7.4 Hz, 1H, H arom Py), 7.16 (s, 1H, H arom), 7.05 (s, 1H, H arom), 7.00 (s, 1H, H arom), 6.35 (d, 2JHH = 15.0 Hz, 1H, CHH–NHC), 5.05 (d, 2JHH = 15.0 Hz, 1H, CHH–NHC), 4.78 (d, 2JHH = 13.9 Hz, 1H, CHHNH), 4.39 (br dd, 3JHH = 9.2 Hz, 3JHH = 9.2 Hz, 1H, NH), 4.28 (dd, 2JHH = 15.0 Hz, 3JHH = 3.1 Hz, 1H, CHHNH), 3.92 (dd, 2JHH = 13.3 Hz, 3JHH = 12.9 Hz, 1H, CHHNH), 3.73 (dd, 2JHH = 13.6 Hz, 3JHH = 12.9 Hz, 1H, CHHNH), 2.40 (s, 3H, CH3), 2.22 (s, 3H, CH3), 1.96 (s, 3H, CH3), −18.83 (d, 1H, 2JHH = 6.9 Hz, IrH trans to Py), −23.28 (d, 1H, 2JHH = 6.9 Hz, IrH cis to Py). 13C{1H} NMR (101 MHz, CD2Cl2): δ 161.0 (Cq arom), 156.8 (C2–NHC), 153.3 (Cq arom), 138.1 (Cq arom), 137.9 (Cq arom), 136.9 (Cq arom), 135.2 (CH arom), 135.0 (Cq arom), 129.0 (2 CH arom), 128.8 (3 CH arom), 128.2 (CH arom), 128.0 (CH arom), 122.4 (CH arom), 120.7 (CH arom), 120.2 (CH arom), 119.1 (CH arom), 63.9 (CH2NH), 60.3 (CH2NH), 55.4 (CH2–NHC), 20.9 (CH3), 18.9 (CH3), 18.0 (CH3).

Complex 5

In a J. Young-valved NMR tube, a suspension of complex 4a (0.014 g, 0.02 mmol) in THF-d8 (0.5 mL) was treated with tBuOK (0.005 g, 0.04 mmol). The sample was analyzed by NMR spectroscopy, showing the formation of complex 5 in ca. 90% NMR yield. Crystals of 5 suitable for XRD analysis were obtained by cooling the above solution at −30 °C. 1H NMR (400 MHz, THF-d8): δ 7.15 (d, 3JHH = 8.6 Hz, 1H, H arom), 6.97 (d, 3JHH = 5.1 Hz, 2H, 2 H arom), 7.73 (m, 4H, 4 H arom), 6.41 (s, 1H, H arom), 6.25 (s, 1H, CH–NHC), 6.06 (d, 3JHH = 7.6 Hz, 2H, 2 H arom), 5.75 (m, 2H, 2 H arom), 4.24 (d, 2JHH = 16.5 Hz, 1H, CHH–NPh), 3.80 (d, 2JHH = 16.5 Hz, 1H, CHH–NPh), 2.29 (s, 3H, CH3), 2.21 (s, 3H, CH3), 1.77 (overlapped with solvent signal, CH3), −16.35 (d, 2JHH = 5.5 Hz, 1H, IrH), −18.61 (d, 2JHH = 5.5 Hz, 1H, IrH). 13C{1H} NMR (101 MHz, THF-d8): δ 164.4 (Cq arom), 164.0 (Cq arom), 159.0 (Cq arom), 156.8 (C2–NHC), 142.3 (Cq arom), 138.7 (Cq arom), 138.4 (Cq arom), 136.3 (Cq arom), 129.8 (CH arom), 129.0 (CH arom), 128.8 (CH arom), 128.3 (CH arom), 127.6 (CH arom), 120.0 (CH arom), 119.8 (CH arom), 116.7 (CH arom), 114.5 (CH arom), 108.3 (CH arom), 107.1 (CH arom), 105.4 (CH arom), 64.2 (CH2N), 49.0 (CH–NHC), 21.0 (CH3), 20.9 (CH3), 20.1 (CH3).

Complex 6

In a J. Young-valved NMR tube, a suspension of complex 4b (0.015 g, 0.02 mmol) in THF-d8 (0.5 mL) was treated with tBuOK (0.007 g, 0.06 mmol). The sample was analyzed by NMR spectroscopy, showing the formation of complex 6 in 75%–80% NMR yield. 1H NMR (400 MHz, THF-d8): δ 7.29 (dd, 3JHH = 7.5 Hz, 3JHH = 7.5 Hz, 1H, H arom Py), 6.81 (s, 1H, H arom Mes), 6.80 (d, 3JHH = 1.6 Hz, 1H, H arom NHC), 6.72 (s, 1H, H arom Mes), 6.31 (d, 3JHH = 1.6 Hz, 1H, H arom NHC), 6.16 (d, 3JHH = 7.3 Hz, 1H, H arom Py), 5.95 (d, 3JHH = 7.6 Hz, 1H, H arom Py), 5.48 (s,1H, NHC–CH), 3.88 (dd, 2JHH = 13.6 Hz, 3JHH = 2.8 Hz, 1H, CHHN), 3.39 (m, 1H, CHHN), 2.97 (br d, 2JHH = 10.9 Hz, 1H, NH), 2.25 (s, 3H, CH3), 2.04 (s, 3H, CH3), 1.77 (s, 3H, CH3), 1.31 (s, 9H, C(CH3)3), – 16.49 (d, 2JHH = 5.6 Hz, 1H, IrH), −18.41 (d, 2JHH = 5.6 Hz, 1H, IrH). 13C{1H} NMR (101 MHz, THF-d8): δ 164.5 (Cqarom), 161.5 (Cqarom), 150.0 (C2–NHC), 140.6 (Cqarom), 137.5 (Cqarom), 136.3 (Cqarom), 136.1 (Cqarom), 131.1 (CH arom), 128.4 (CH arom), 127.9 (CH arom), 120.0 (CH arom), 116.4 (CH arom), 112.5 (CH arom), 110.2 (CH arom), 56.8 (C(CH3)3), 56.1 (CH2NH), 54.9 (CH–NHC), 29.2 (C(CH3)3), 21.1 (CH3), 20.8 (CH3), 20.0 (CH3).

Complex 7

In a J. Young-valved NMR tube, a suspension of complex 4b (0.015 g, 0.02 mmol) in THF-d8 (0.5 mL) was treated with tBuOK (0.007 g, 0.06 mmol). The sample was pressurized with 4 bar H2, and analyzed by NMR spectroscopy after 48 h. 1H NMR (400 MHz, THF-d8): δ 7.52 (dd, 3JHH = 7.6 Hz, 3JHH = 7.6 Hz, 1H, H arom Py), 7.15 (m, 2H, 2 H arom Py), 7.00 (d, 3JHH = 1.9 Hz, 1H, H arom NHC), 6.82 (s, 1H, H arom Mes), 6.80 (s, 1H, H arom Mes), 6.56 (d, 3JHH = 1.9 Hz, 1H, H arom NHC), 5.10 (d, 2JHH = 13.5 Hz,1H, CHH–NHC), 5.02 (br, 1H, NH), 5.02 (d, 2JHH = 13.5 Hz,1H, CHH-NHC), 4.25 (dd, 2JHH = 14.4 Hz, 3JHH = 3.8 Hz, 1H, CHHN), 4.02 (dd, 2JHH = 14.4 Hz, 3JHH = 5.8 Hz, 1H, CHHN), 2.27 (s, 3H, CH3), 2.03 (s, 3H, CH3), 1.95 (s, 3H, CH3), 1.11 (s, 9H, C(CH3)3), −8.92 (m, 2H, 2 IrH), −18.24 (dd, 2JHH = 5.2 Hz, 2JHH = 5.2 Hz, 1H, IrH). 13C{1H} NMR (101 MHz, THF-d8): δ 164.5 (Cqarom), 164.0 (Cqarom), 153.8 (C2–NHC), 139.5 (Cqarom), 137.3 (Cqarom), 136.7 (Cqarom), 131.5 (CH arom), 128.6 (CH arom), 128.5 (CH arom), 120.9 (CH arom), 119.1 (CH arom), 118.6 (CH arom), 117.9 (CH arom), 59.0 (CH2–NHC), 57.7 (CH2NH), 50.7 (C(CH3)3), 27.5 (C(CH3)3), 21.1 (CH3), 19.6 (CH3), 19.4 (CH3); signals of two quaternary aromatic carbons could not be unambiguously assigned.

Ammonia-Borane Dehydrogenation

Experimental Setup for H2 Production

H2 generation was followed up using a Fisher-Porter vessel (25 mL) that was connected to a vacuum line and coupled to an ESI pressure gauge model GS4200-USB (0–6 bar) that was connected to a computer (see the Supporting Information for details).

Representative Procedure for AB Dehydrogenation

A solution of complex 4a (4.3 mg, 6.5 μmol) and tBuOK (1.8 mg, 0.016 mmol) in THF (0.5 mL) was added to a freshly prepared, stirred (750 rpm) solution of AB (50.0 mg, 1.62 mmol) in THF (0.5 mL) at room temperature. H2 generation was monitored by registering the increase of pressure in the system. At the end of the reaction, the supernatant was transferred to an NMR tube and analyzed via 11B NMR spectroscopy, to identify the soluble reaction byproducts. The insoluble residue was washed subsequently with THF and Et2O, and then dried under vacuum before recording the IR spectrum.

Acknowledgments

Financial support (FEDER contribution) from the Spanish Agencia Estatal de Investigación (PID2019-104159GB-I00/AEI/10.13039/501100011033) and Junta de Andalucía (P18-FR-3208) is gratefully acknowledged. A.R. would like to thank the CNR-RFBR (Russia) Bilateral Project (2018-2020) for financial support to this research activity. Mr. Arne Estelmann is thanked for preliminary experimental work.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.1c03056.

  • Experimental setup details, kinetic studies, selected NMR spectra, and X-ray diffraction data (PDF)

  • Cartesian coordinates, absolute Gibbs energies (THF) and imaginary frequencies (TSs) of the DFT-optimized structures (XYZ)

Accession Codes

CCDC 2113023–2113025 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, 909 Cambridge CB2 1EZ, UK; fax: + 44 1223 336033.

The authors declare no competing financial interest.

Supplementary Material

ic1c03056_si_001.pdf (4.5MB, pdf)
ic1c03056_si_002.xyz (32.4KB, xyz)

References

  1. a Yadav M.; Xu Q. Liquid-Phase Chemical Hydrogen Storage Materials. Energy Environ. Sci. 2012, 5, 9698–9725. 10.1039/c2ee22937d. [DOI] [Google Scholar]; b Dalebrook A. F.; Gan W.; Grasemann M.; Moret S.; Laurenczy G. Hydrogen Storage: Beyond Conventional Methods. Chem. Commun. 2013, 49, 8735–8751. 10.1039/c3cc43836h. [DOI] [PubMed] [Google Scholar]; c Demirci U. B.; Miele P. Chemical Hydrogen Storage: ‘Material’ Gravimetric Capacity versus ‘System’ Gravimetric Capacity. Energy Environ. Sci. 2011, 4, 3334–3341. 10.1039/c1ee01612a. [DOI] [Google Scholar]
  2. a Durbin D. J.; Malardier-Jugroot C. Review of Hydrogen Storage Techniques for On Board Vehicle Applications. Int. J. Hydrogen Energy 2013, 38, 14595–14617. 10.1016/j.ijhydene.2013.07.058. [DOI] [Google Scholar]; b Zhang F.; Zhao P.; Niu M.; Maddy J. The Survey of Key Technologies in Hydrogen Energy Storage. Int. J. Hydrogen Energy 2016, 41, 14535–14552. 10.1016/j.ijhydene.2016.05.293. [DOI] [Google Scholar]; c Sakintuna B.; Lamari-Darkrim F.; Hirscher M. Metal Hydride Materials for Solid Hydrogen Storage: A Review. Int. J. Hydrogen Energy 2007, 32, 1121–1140. 10.1016/j.ijhydene.2006.11.022. [DOI] [Google Scholar]
  3. a Stephens F. H.; Pons V.; Tom Baker R. Ammonia–Borane: The Hydrogen Source Par Excellence?. Dalton Trans. 2007, 2613–2626. 10.1039/B703053C. [DOI] [PubMed] [Google Scholar]; b Hamilton C. W.; Baker R. T.; Staubitz A.; Manners I. B–N Compounds for Chemical Hydrogen Storage. Chem. Soc. Rev. 2009, 38, 279–293. 10.1039/B800312M. [DOI] [PubMed] [Google Scholar]; c Staubitz A.; Robertson A. P. M.; Manners I. Ammonia-Borane and Related Compounds as Dihydrogen Sources. Chem. Rev. 2010, 110, 4079–4124. 10.1021/cr100088b. [DOI] [PubMed] [Google Scholar]; d Li H.; Yang Q.; Chen X.; Shore S. G. Ammonia Borane, Past as Prolog. J. Organomet. Chem. 2014, 751, 60–66. 10.1016/j.jorganchem.2013.08.044. [DOI] [Google Scholar]; e Marder T. B. Will We Soon Be Fueling our Automobiles with Ammonia-Borane?. Angew. Chem., Int. Ed. 2007, 46, 8116–8118. 10.1002/anie.200703150. [DOI] [PubMed] [Google Scholar]
  4. a Staubitz A.; Presa Soto A.; Manners I. Iridium-Catalyzed Dehydrocoupling of Primary Amine–Borane Adducts: A Route to High Molecular Weight Polyaminoboranes, Boron–Nitrogen Analogues of Polyolefins. Angew. Chem., Int. Ed. 2008, 47, 6212–6215. 10.1002/anie.200801197. [DOI] [PubMed] [Google Scholar]; b Clark T. J.; Lee K.; Manners I. Transition-Metal-Catalyzed Dehydrocoupling: A Convenient Route to Bonds between Main-Group Elements. Chem. - Eur. J. 2006, 12, 8634–8648. 10.1002/chem.200600981. [DOI] [PubMed] [Google Scholar]; c Staubitz A.; Robertson A. P. M.; Sloan M. E.; Manners I. Amine- and Phosphine-Borane Adducts: New Interest in Old Molecules. Chem. Rev. 2010, 110, 4023–4078. 10.1021/cr100105a. [DOI] [PubMed] [Google Scholar]; d Whittell G. R.; Manners I. Advances with Ammonia-Borane: Improved Recycling and Use as a Precursor to Atomically Thin BN Films. Angew. Chem., Int. Ed. 2011, 50, 10288–10289. 10.1002/anie.201103830. [DOI] [PubMed] [Google Scholar]; e Colebatch A. L.; Weller A. S. Amine–Borane Dehydropolymerization:Challenges andOpportunities. Chem. - Eur. J. 2019, 25, 1379–1390. 10.1002/chem.201804592. [DOI] [PMC free article] [PubMed] [Google Scholar]; f Han D.; Anke F.; Trose M.; Beweries T. Recent Advances in Transition Metal Catalysed Dehydropolymerisation of Amine Boranes and Phosphine Boranes. Coord. Chem. Rev. 2019, 380, 260–286. 10.1016/j.ccr.2018.09.016. [DOI] [Google Scholar]
  5. a Smythe N. C.; Gordon J. C. Ammonia Borane as a Hydrogen Carrier: Dehydrogenation and Regeneration. Eur. J. Inorg. Chem. 2010, 2010, 509–521. 10.1002/ejic.200900932. [DOI] [Google Scholar]; b Bluhm M. E.; Bradley M. G.; Butterick R. III; Kusari U.; Sneddon L. G. Amineborane-Based Chemical Hydrogen Storage: Enhanced Ammonia Borane Dehydrogenation in Ionic Liquids. J. Am. Chem. Soc. 2006, 128, 7748–7749. 10.1021/ja062085v. [DOI] [PubMed] [Google Scholar]
  6. a Rossin A.; Peruzzini M. Ammonia-Borane and Amine-Borane Dehydrogenation Mediated by Complex Metal Hydrides. Chem. Rev. 2016, 116, 8848–8872. 10.1021/acs.chemrev.6b00043. [DOI] [PubMed] [Google Scholar]; b Alcaraz G.; Sabo-Etienne S. Coordination and Dehydrogenation of Amine–Boranes at Metal Centers. Angew. Chem., Int. Ed. 2010, 49, 7170–7179. 10.1002/anie.201000898. [DOI] [PubMed] [Google Scholar]; c Bhunya S.; Malakar T.; Ganguly G.; Paul A. Combining Protons and Hydrides by Homogeneous Catalysis for Controlling the Release of Hydrogen from Ammonia-Borane: Present Status and Challenges. ACS Catal. 2016, 6, 7907–7934. 10.1021/acscatal.6b01704. [DOI] [Google Scholar]
  7. Denney M. C.; Pons V.; Hebden T. J.; Heinekey D. M.; Goldberg K. I. Efficient Catalysis of Ammonia Borane Dehydrogenation. J. Am. Chem. Soc. 2006, 128, 12048–12049. 10.1021/ja062419g. [DOI] [PubMed] [Google Scholar]
  8. Bhattacharya P.; Krause J. A.; Guan H. Mechanistic Studies of Ammonia Borane Dehydrogenation Catalyzed by Iron Pincer Complexes. J. Am. Chem. Soc. 2014, 136, 11153–11161. 10.1021/ja5058423. [DOI] [PubMed] [Google Scholar]
  9. Nugent J. W.; García-Melchor M.; Fout A. R. Cobalt-Catalyzed Ammonia Borane Dehydrogenation: Mechanistic Insight and Isolation of a Cobalt Hydride-Amidoborane Complex. Organometallics 2020, 39, 2917–2927. 10.1021/acs.organomet.0c00459. [DOI] [Google Scholar]
  10. Esteruelas M. A.; Nolis P.; Oliván M.; Oñate E.; Vallribera A.; Vélez A. Ammonia Borane Dehydrogenation Promoted by a Pincer-Square-Planar Rhodium(I) Monohydride: A Stepwise Hydrogen Transfer from the Substrate to the Catalyst. Inorg. Chem. 2016, 55, 7176–7181. 10.1021/acs.inorgchem.6b01216. [DOI] [PubMed] [Google Scholar]
  11. Luconi L.; Osipova E. S.; Giambastiani G.; Peruzzini M.; Rossin A.; Belkova N. V.; Filippov O. A.; Titova E. M.; Pavlov A. A.; Shubina E. S. Amine Boranes Dehydrogenation Mediated by an Unsymmetrical Iridium Pincer Hydride: (PCN) vs (PCP) Improved Catalytic Performance. Organometallics 2018, 37, 3142–3153. 10.1021/acs.organomet.8b00488. [DOI] [Google Scholar]
  12. Rossin A.; Bottari G.; Lozano-Vila A. M.; Paneque M.; Peruzzini M.; Rossi A.; Zanobini F. Catalytic Amine-Borane Dehydrogenation by a PCP-Pincer Palladium Complex: A Combined Experimental and DFT Analysis of the Reaction Mechanism. Dalton Trans. 2013, 42, 3533–3541. 10.1039/c2dt32273k. [DOI] [PubMed] [Google Scholar]
  13. a Käβ M.; Friedrich A.; Drees M.; Schneider S. Ruthenium Complexes with Cooperative PNP Ligands: Bifunctional Catalysts for the Dehydrogenation of Ammonia-Borane. Angew. Chem., Int. Ed. 2009, 48, 905–907. 10.1002/anie.200805108. [DOI] [PubMed] [Google Scholar]; b Marziale A. N.; Friedrich A.; Klopsch I.; Drees M.; Celinski V. R.; Schmedt auf der Günne J.; Schneider S. The Mechanism of Borane-Amine Dehydrocoupling with Bifunctional Ruthenium Catalysts. J. Am. Chem. Soc. 2013, 135, 13342–13355. 10.1021/ja311092c. [DOI] [PubMed] [Google Scholar]
  14. Glüer A.; Förster M.; Celinski V. R.; Schmedt auf der Günne J.; Holthausen M. C.; Schneider S. Highly Active Iron Catalyst for Ammonia Borane Dehydrocoupling at Room Temperature. ACS Catal. 2015, 5, 7214–7217. 10.1021/acscatal.5b02406. [DOI] [Google Scholar]
  15. Elsby M. R.; Ghostine K.; Das U. K.; Gabidullin B. M.; Baker R. T. Iron-SNS and -CNS Complexes: Selective Caryl-S Bond Cleavage and Amine-Borane Dehydrogenation Catalysis. Organometallics 2019, 38, 3844–3851. 10.1021/acs.organomet.9b00524. [DOI] [Google Scholar]
  16. Luconi L.; Demirci U. B.; Peruzzini M.; Giambastiani G.; Rossin A. Ammonia Borane and Hydrazine Bis(Borane) Dehydrogenation Mediated by an Unsymmetrical (PNN) Ruthenium Pincer Hydride: Metal–Ligand Cooperation for Hydrogen Production. Sustainable Energy Fuels 2019, 3, 2583–2596. 10.1039/C9SE00241C. [DOI] [Google Scholar]
  17. Blaquiere N.; Diallo-Garcia S.; Gorelsky S. I.; Black D. A.; Fagnou K. Ruthenium-Catalyzed Dehydrogenation of Ammonia Boranes. J. Am. Chem. Soc. 2008, 130, 14034–14035. 10.1021/ja804235t. [DOI] [PubMed] [Google Scholar]
  18. Takahashi H.; Watanabe T.; Tobita H. Bifunctional Iron-Amino Complexes: Highly Efficient Catalysts for Dehydrogenation of Ammonia-Borane. Chem. Lett. 2018, 47, 296–299. 10.1246/cl.171052. [DOI] [Google Scholar]
  19. a Conley B. L.; Williams T. J. Dehydrogenation of Ammonia-Borane by Shvo’s Catalyst. Chem. Commun. 2010, 46, 4815–4817. 10.1039/c003157g. [DOI] [PubMed] [Google Scholar]; b Conley B. L.; Guess D.; Williams T. J. A Robust, Air-Stable, Reusable Ruthenium Catalyst for Dehydrogenation of Ammonia Borane. J. Am. Chem. Soc. 2011, 133, 14212–14215. 10.1021/ja2058154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. a Zhao B.; Han Z.; Ding K. The N-H Functional Group in Organometallic Catalysis. Angew. Chem., Int. Ed. 2013, 52, 4744–4788. 10.1002/anie.201204921. [DOI] [PubMed] [Google Scholar]; b Dub P. A.; Gordon J. C. The Role of the Metal-Bound N–H Functionality in Noyori-Type Molecular Catalysts. Nat. Rev. Chem. 2018, 2, 396–408. 10.1038/s41570-018-0049-z. [DOI] [Google Scholar]
  21. a Milstein D. Metal-Ligand Cooperation by Aromatization-Dearomatization as a Tool in Single Bond Activation. Philos. Trans. R. Soc., A 2015, 373, 20140189. 10.1098/rsta.2014.0189. [DOI] [PubMed] [Google Scholar]; b Shimbayashi T.; Fujita K. Recent Advances in Homogeneous Catalysis via Metal-Ligand Cooperation Involving Aromatization and Dearomatization. Catalysts 2020, 10, 635. 10.3390/catal10060635. [DOI] [Google Scholar]
  22. a Filonenko G. A.; Cosimi E.; Lefort L.; Conley M. P.; Copéret C.; Lutz M.; Hensen E. J. M.; Pidko E. A. Lutidine-Derived Ru-CNC Hydrogenation Pincer Catalysts with Versatile Coordination Properties. ACS Catal. 2014, 4, 2667–2671. 10.1021/cs500720y. [DOI] [Google Scholar]; b Filonenko G. A.; Smykowski D.; Szyja B. M.; Li G.; Szczygiel J.; Hensen E. J. M.; Pidko E. A. Catalytic Hydrogenation of CO2 to Formates by a Lutidine-Derived Ru-CNC Pincer Complex: Theoretical Insight into the Unrealized Potential. ACS Catal. 2015, 5, 1145–1154. 10.1021/cs501990c. [DOI] [Google Scholar]; c Sun Y.; Koehler C.; Tan R.; Annibale V. T.; Song D. Ester Hydrogenation Catalyzed by Ru-CNN Pincer Complexes. Chem. Commun. 2011, 47, 8349–8351. 10.1039/c1cc12601f. [DOI] [PubMed] [Google Scholar]; d Del Pozo C.; Iglesias M.; Sánchez F. Pincer-Type Pyridine-Based N-Heterocyclic Carbene Amine Ru(II) Complexes as Efficient Catalysts for Hydrogen Transfer Reactions. Organometallics 2011, 30, 2180–2188. 10.1021/om101179a. [DOI] [Google Scholar]; e Fogler E.; Balaraman E.; Ben-David Y.; Leitus G.; Shimon L. J. W.; Milstein D. New CNN-Type Ruthenium Pincer NHC Complexes. Mild, Efficient Catalytic Hydrogenation of Esters. Organometallics 2011, 30, 3826–3833. 10.1021/om200367j. [DOI] [Google Scholar]; f Hernández-Juárez M.; Vaquero M.; Álvarez E.; Salazar V.; Suárez A. Hydrogenation of Imines Catalysed by Ruthenium(II) Complexes Based on Lutidine-Derived CNC Pincer Ligands. Dalton Trans. 2013, 42, 351–354. 10.1039/C2DT31907A. [DOI] [PubMed] [Google Scholar]; g Hernández-Juárez M.; López-Serrano J.; Lara P.; Morales-Cerón J. P.; Vaquero M.; Álvarez E.; Salazar V.; Suárez A. Ruthenium(II) Complexes Containing Lutidine-Derived Pincer CNC Ligands: Synthesis, Structure, and Catalytic Hydrogenation of C-N Bonds. Chem. - Eur. J. 2015, 21, 7540–7555. 10.1002/chem.201406040. [DOI] [PubMed] [Google Scholar]; h Hernández-Juárez M.; López-Serrano J.; González-Herrero P.; Rendón N.; Álvarez E.; Paneque M.; Suárez A. Hydrogenation of an Iridium-Coordinated Imidazol-2-ylidene Ligand Fragment. Chem. Commun. 2018, 54, 3843–3846. 10.1039/C8CC00420J. [DOI] [PubMed] [Google Scholar]; i Hernández-Juárez M.; Sánchez P.; López-Serrano J.; Lara P.; González-Herrero P.; Rendón N.; Álvarez E.; Paneque M.; Suárez A. Metalated Ir-CNP Complexes Containing Imidazolin-2-ylidene and Imidazolidin-2-ylidene Donors - Synthesis, Structure, Luminescence, and Metal-Ligand Cooperative Reactivity. Eur. J. Inorg. Chem. 2020, 2020, 3944–3953. 10.1002/ejic.202000681. [DOI] [Google Scholar]; j Sánchez P.; Hernández-Juárez M.; Rendón N.; López-Serrano J.; Álvarez E.; Paneque M.; Suárez A. Hydroboration of Carbon Dioxide with Catechol and Pinacol Borane Using an Ir-CNP* Pincer Complex. Water Influence on the Catalytic Activity. Dalton Trans. 2018, 47, 16766–16776. 10.1039/C8DT03951H. [DOI] [PubMed] [Google Scholar]
  23. Sánchez P.; Hernández-Juárez M.; Álvarez E.; Paneque M.; Rendón N.; Suárez A. Synthesis, Structure and Reactivity of Pd and Ir Complexes Based on New Lutidine-Derived NHC/Phosphine Mixed Pincer Ligands. Dalton Trans. 2016, 45, 16997–17009. 10.1039/C6DT03652J. [DOI] [PubMed] [Google Scholar]
  24. Sánchez P.; Hernández-Juárez M.; Rendón N.; López-Serrano J.; Álvarez E.; Paneque M.; Suárez A. Selective, Base-Free Hydrogenation of Aldehydes Catalyzed by Ir Complexes Based on Proton-Responsive Lutidine-Derived CNP Ligands. Organometallics 2021, 40, 1314–1327. 10.1021/acs.organomet.1c00109. [DOI] [Google Scholar]
  25. Sánchez P.; Hernández-Juárez M.; Rendón N.; López-Serrano J.; Santos L. L.; Álvarez E.; Paneque M.; Suárez A. Hydrogenation/Dehydrogenation of N-Heterocycles Catalyzed by Ruthenium Complexes Based on Multimodal Proton-Responsive CNN(H) Pincer Ligands. Dalton Trans. 2020, 49, 9583–9587. 10.1039/D0DT02326D. [DOI] [PubMed] [Google Scholar]
  26. For relevant examples, see:; a Fogler E.; Garg J. A.; Hu P.; Leitus G.; Shimon L. J. W.; Milstein D. System with Potential Dual Modes of Metal-Ligand Cooperation: Highly Catalytically Active Pyridine-Based PNNH-Ru Pincer Complexes. Chem. - Eur. J. 2014, 20, 15727–15731. 10.1002/chem.201405295. [DOI] [PubMed] [Google Scholar]; b Kar S.; Rauch M.; Kumar A.; Leitus G.; Ben-David Y.; Milstein D. Selective Room-Temperature Hydrogenation of Amides to Amines and Alcohols Catalyzed by a Ruthenium Pincer Complex and Mechanistic Insight. ACS Catal. 2020, 10, 5511–5515. 10.1021/acscatal.0c01406. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Srimani D.; Mukherjee A.; Goldberg A. F. G.; Leitus G.; Diskin-Posner Y.; Shimon L. J. W.; Ben-David Y.; Milstein D. Cobalt-Catalyzed Hydrogenation of Esters to Alcohols: Unexpected Reactivity Trend Indicates Ester Enolate Intermediacy. Angew. Chem., Int. Ed. 2015, 54, 12357–12360. 10.1002/anie.201502418. [DOI] [PubMed] [Google Scholar]; d Daw P.; Chakraborty S.; Garg J. A.; Ben-David Y.; Milstein D. Direct Synthesis of Pyrroles by Dehydrogenative Coupling of Diols and Amines Catalyzed by Cobalt Pincer Complexes. Angew. Chem., Int. Ed. 2016, 55, 14373–14377. 10.1002/anie.201607742. [DOI] [PubMed] [Google Scholar]; e Kumar A.; Janes T.; Espinosa-Jalapa N. A.; Milstein D. Manganese Catalyzed Hydrogenation of Organic Carbonates to Methanol and Alcohols. Angew. Chem., Int. Ed. 2018, 57, 12076–1043. 10.1002/anie.201806289. [DOI] [PubMed] [Google Scholar]; f Kumar A.; Espinosa-Jalapa N. A.; Leitus G.; Diskin-Posner Y.; Avram L.; Milstein D. Direct Synthesis of Amides by Dehydrogenative Coupling of Amines with either Alcohols or Esters: Manganese Pincer Complex as Catalyst. Angew. Chem., Int. Ed. 2017, 56, 14992–14996. 10.1002/anie.201709180. [DOI] [PubMed] [Google Scholar]; g Espinosa-Jalapa N. A.; Kumar A.; Leitus G.; Diskin-Posner Y.; Milstein D. Synthesis of Cyclic Imides by Acceptorless Dehydrogenative Coupling of Diols and Amines Catalyzed by a Manganese Pincer Complex. J. Am. Chem. Soc. 2017, 139, 11722–11725. 10.1021/jacs.7b08341. [DOI] [PubMed] [Google Scholar]; h Tan X.; Wang Y.; Liu Y.; Wang F.; Shi L.; Lee K.-H.; Lin Z.; Lv H.; Zhang X. Highly Efficient Tetradentate Ruthenium Catalyst for Ester Reduction: Especially for Hydrogenation of Fatty Acid Esters. Org. Lett. 2015, 17, 454–457. 10.1021/ol503456j. [DOI] [PubMed] [Google Scholar]; i Shi L.; Tan X.; Long J.; Xiong X.; Yang S.; Xue P.; Lv H.; Zhang X. Direct Catalytic Hydrogenation of Simple Amides: A Highly Efficient Approach from Amides to Amines and Alcohols. Chem. - Eur. J. 2017, 23, 546–548. 10.1002/chem.201604904. [DOI] [PubMed] [Google Scholar]
  27. a Rowland R. S.; Taylor R. Intermolecular Nonbonded Contact Distances in Organic Crystal Structures: Comparison with Distances Expected from van der Waals Radii. J. Phys. Chem. 1996, 100, 7384–7391. 10.1021/jp953141+. [DOI] [Google Scholar]; b Reedijk J. Coordination Chemistry Beyond Werner: Interplay Between Hydrogen Bonding and Coordination. Chem. Soc. Rev. 2013, 42, 1776–1783. 10.1039/C2CS35239G. [DOI] [PubMed] [Google Scholar]
  28. Kim S.-K.; Han W.-S.; Kim T.-J.; Kim T.-Y.; Nam S. W.; Mitoraj M.; Piekoś Ł.; Michalak A.; Hwang S.-J.; Kang S. O. Palladium Catalysts for Dehydrogenation of Ammonia Borane with Preferential B-H Activation. J. Am. Chem. Soc. 2010, 132, 9954–9955. 10.1021/ja101685u. [DOI] [PubMed] [Google Scholar]
  29. Staubitz A.; Sloan M. E.; Robertson A. P. M.; Friedrich A.; Schneider S.; Gates P. J.; Auf der Günne J. S.; Manners I. Catalytic Dehydrocoupling/Dehydrogenation of N-Methylamine-Borane and Ammonia-Borane: Synthesis and Characterization of High Molecular Weight Polyaminoboranes. J. Am. Chem. Soc. 2010, 132, 13332–13345. 10.1021/ja104607y. [DOI] [PubMed] [Google Scholar]
  30. a Gómez-Gallego M.; Sierra M. A. Kinetic Isotope Effects inthe Study of Organometallic Reaction Mechanisms. Chem. Rev. 2011, 111, 4857–4963. 10.1021/cr100436k. [DOI] [PubMed] [Google Scholar]; b Simmons E. M.; Hartwig J. F. On the Interpretation of Deuterium Kinetic Isotope Effects in C–H Bond Functionalizations by Transition-Metal Complexes. Angew. Chem., Int. Ed. 2012, 51, 3066–3072. 10.1002/anie.201107334. [DOI] [PubMed] [Google Scholar]
  31. Espinosa-Jalapa N. A.; Nerush A.; Shimon L. J. W.; Leitus G.; Avram L.; Ben-David Y.; Milstein D. Manganese-Catalyzed Hydrogenation of Esters to Alcohols. Chem. - Eur. J. 2017, 23, 5934–5938. 10.1002/chem.201604991. [DOI] [PubMed] [Google Scholar]
  32. Tejel C.; del Río M. P.; Ciriano M. A.; Reijerse E. J.; Hartl F.; Zális S.; Hetterscheid D. G. H.; Tsichlis i Spithas N.; de Bruin B. Ligand-Centred Reactivity of Bis(picolyl)amine Iridium: Sequential Deprotonation, Oxidation and Oxygenation of a “Non-Innocent” Ligand. Chem. - Eur. J. 2009, 15, 11878–11889. 10.1002/chem.200901017. [DOI] [PubMed] [Google Scholar]
  33. Morris K. F.; Johnson C. S. Jr. Diffusion-Ordered Two-Dimensional Nuclear Magnetic Resonance Spectroscopy. J. Am. Chem. Soc. 1992, 114, 3139–3141. 10.1021/ja00034a071. [DOI] [Google Scholar]
  34. The X-ray volume (VX-ray) was calculated by dividing the crystallographic unit-cell volume by the number of molecular entities contained in the unit cell, assuming that these have a spherical shape, and subtracting the van der Waals volumes of the THF-solvated K+ ions. See:; Zhao Y. H.; Abraham M. H.; Zissimos A. M. Fast Calculation of van der Waals Volume as a Sum of Atomic and Bond Contributions and Its Application to Drug Compounds. J. Org. Chem. 2003, 68, 7368–7373. 10.1021/jo034808o. [DOI] [PubMed] [Google Scholar]
  35. Rueping M.; Koenigs R. M.; Borrmann R.; Zoller J.; Weirich T. E.; Mayer J. Size-Selective, Stabilizer-Free, Hydrogenolytic Synthesis of Iridium Nanoparticles Supported on Carbon Nanotubes. Chem. Mater. 2011, 23, 2008–2010. 10.1021/cm1032578. [DOI] [Google Scholar]
  36. Ramachandran P. V.; Gagare P. D. Preparation of Ammonia Borane in High Yield and Purity, Methanolysis, and Regeneration. Inorg. Chem. 2007, 46, 7810–7817. 10.1021/ic700772a. [DOI] [PubMed] [Google Scholar]
  37. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, Rev. B.01; Gaussian: Wallingford, CT, 2016. [Google Scholar]
  38. Yu H. S.; He X.; Li S. L.; Truhlar D. G. MN15: A Kohn–Sham Global-Hybrid Exchange–Correlation Density Functional with Broad Accuracy for Multi-reference and Single-reference Systems and Noncovalent Interactions. Chem. Sci. 2016, 7, 5032–5051. 10.1039/C6SC00705H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. a Leininger T.; Nicklass A.; Stoll H.; Dolg M.; Schwerdtfeger P. The Accuracy of the Pseudopotential Approximation. II. A Comparison of Various Core Sizes for Indium Pseudopotentials in Calculations for Spectroscopic Constants of InH, InF, and InCl. J. Chem. Phys. 1996, 105, 1052–1059. 10.1063/1.471950. [DOI] [Google Scholar]; b Kuchle W.; Dolg M.; Stoll H.; Preuss H. Energy-Adjusted Pseudopotentials for the Actinides. Parameter Sets and Test Calculations for Thorium and Thorium Monoxide. J. Chem. Phys. 1994, 100, 7535–7542. 10.1063/1.466847. [DOI] [Google Scholar]; c Häussermann U.; Dolg M.; Stoll H.; Preuss H.; Schwerdtfeger P.; Pitzer R. M. Accuracy of Energy-Adjusted Quasirelativistic ab initio Pseudopotentials. Mol. Phys. 1993, 78, 1211–1224. 10.1080/00268979300100801. [DOI] [Google Scholar]; d Andrae D.; Häußermann U.; Dolg M.; Stoll H.; Preuß H. Energy-Adjusted ab initio Pseudopotentials for the Second and Third Row Transition Elements. Theoret. Chim. Acta 1990, 77, 123–141. 10.1007/BF01114537. [DOI] [Google Scholar]
  40. Ehlers A. W.; Böhme M.; Dapprich S.; Gobbi A.; Höllwarth A.; Jonas V.; Köhler K. F.; Stegmann R.; Veldkamp A.; Frenking G. A Set of f-Polarization Functions for Pseudopotential Basis Sets of the Transition Metals Sc-Cu, Y-Ag and La-Au. Chem. Phys. Lett. 1993, 208, 111–114. 10.1016/0009-2614(93)80086-5. [DOI] [Google Scholar]
  41. a Clark T.; Chandrasekhar J.; Spitznagel G. W.; Schleyer P. V. R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-row Elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. 10.1002/jcc.540040303. [DOI] [Google Scholar]; b Francl M. M.; Pietro W. J.; Hehre W. J.; Binkley J. S.; DeFrees D. J.; Pople J. A.; Gordon M. S. Self-consistent Molecular Orbital Methods. XXIII. A Polarization-type Basis Set for Second-row Elements. J. Chem. Phys. 1982, 77, 3654–3665. 10.1063/1.444267. [DOI] [Google Scholar]; c Hariharan P. C.; Pople J. A. The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theoret. Chim. Acta 1973, 28, 213–222. 10.1007/BF00533485. [DOI] [Google Scholar]; d Hehre W. J.; Ditchfield R.; Pople J. A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. 10.1063/1.1677527. [DOI] [Google Scholar]
  42. 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. 10.1021/jp810292n. [DOI] [PubMed] [Google Scholar]
  43. a Fukui K. The Path of Chemical Reactions - The IRC Approach. Acc. Chem. Res. 1981, 14, 363–368. 10.1021/ar00072a001. [DOI] [Google Scholar]; b Hratchian H. P., Schlegel H. B. In Theory and Applications of Computational Chemistry; Frenking G., Kim K. S., Scuseria G. E., Eds.; Elsevier: Amsterdam, 2005. [Google Scholar]

Associated Data

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

Supplementary Materials

ic1c03056_si_001.pdf (4.5MB, pdf)
ic1c03056_si_002.xyz (32.4KB, xyz)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES