Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Stereoselective formation and catalytic activity of hydrido(acylphosphane)(chlorido)(pyrazole)rhodium(III) complexes. Experimental and DFT studies

Virginia San Nacianceno a, Susan Azpeitia a, Lourdes Ibarlucea a, Claudio Mendicute-Fierro a, Antonio Rodríguez-Diéguez b, José M. Seco a, Eider San Sebastian a and María A. Garralda *a
aFacultad de Química de San Sebastián, Universidad del País Vasco (UPV/EHU), Apdo. 1072, 20080 San Sebastián, Spain. E-mail: mariaangeles.garralda@ehu.es
bFacultad de Ciencias. Universidad de Granada, Avda. Fuenteventura s/n, 18071 Granada, Spain

Received 6th May 2015 , Accepted 2nd June 2015

First published on 3rd June 2015


Abstract

The reaction of [{RhCl(COD)}2] (COD = 1,5-cyclooctadiene) with L = pyrazole (Hpz), 3(5)-methylpyrazole (Hmpz) or 3,5-dimethylpyrazole (Hdmpz) and PPh2(o-C6H4CHO) (Rh[thin space (1/6-em)]:[thin space (1/6-em)]L[thin space (1/6-em)]:[thin space (1/6-em)]P = 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1) gives hydridoacyl complexes [RhHCl{PPh2(o-C6H4CO)}(L)2] (1). Stereoselective formation of 1-Hpz and 1-Hmpz with pyrazoles trans to hydrido and phosphorus and hydrogen bond formation with O-acyl and chlorido occur. 1-Hmpz is a mixture of two linkage isomers in a 9[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio, with two 5-methylpyrazole ligands or with one 3- and one 5-methylpyrazole ligand, respectively. Fluxional 1-Hdmpz undergoes metallotropic tautomerization and is a mixture of equal amounts of 1a-Hdmpz and 1b-Hdmpz, with hydrido trans to pyrazole or chlorido, respectively. Complexes 1 readily exchange hydrido by chlorido to afford [RhCl2{PPh2(o-C6H4CO)}(L)2] (2-Hpz, 2-Hmpz and 2-Hdmpz) as single isomers with cis chloridos and two N–H⋯Cl hydrogen bonds. The reaction of 1 with PPh3 or PPh2OH affords static [RhHCl{PPh2(o-C6H4CO)}(PPh3)L] (3) or [RhHCl{PPh2(o-C6H4CO)}(PPh2OH)L] (4) respectively with trans P-atoms and pyrazoles forming N–H⋯Cl hydrogen bonds. 3-Hpz and 3-Hmpz contain single species with hydrido cis to chlorido, while 3-Hdmpz is a mixture of equal amounts of 3a-Hdmpz and 3b-Hdmpz. Complexes 4, with an additional O–H⋯O hydrogen bond, selectively contain only the cis-H,Cl species with all the three ligands. The reaction of [{RhCl(COD)}2] with L and PPh2(o-C6H4CHO) (Rh[thin space (1/6-em)]:[thin space (1/6-em)]L[thin space (1/6-em)]:[thin space (1/6-em)]P = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2) led to complexes with trans P-atoms, [RhHCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHO)-κP}L] (5-Hpz, 5a-Hdmpz and 5b-Hdmpz), at room temperature, and to [RhCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHOH)}(Hmpz)] (6-Hmpz) or [RhCl{PPh2(o-C6H4CO)}2L] (7) with hydrogen evolution in refluxing benzene. DFT calculations were used to predict the correct isomers, their ratios and the particular intramolecular hydrogen bonds in these complexes. Single crystal X-ray diffraction analysis was performed on 2-Hpz, 3a-Hdmpz and 7-Hpz. Complexes 1 are efficient homogeneous catalysts (0.5 mol% loading) in the hydrolysis of amine– or ammonia–borane (AB) to generate up to 3 equivalents of hydrogen in the presence of air.


Introduction

Acylhydrido transition metal species are well known to be involved in many stoichiometric and catalytic reactions such as hydroacylation1 or aldehyde decarbonylation reactions,2 among others. Recently we have reported that hydridoirida-β-diketones, which contain an acyl and a hydroxycarbene group stabilized by a strong O⋯H⋯O intramolecular hydrogen bond or {(acylphosphane)(diphenylphosphinous acid)}rhodium(III) complexes, which contain a hydrogen bond between the acidic hydrogen atom of the coordinated phosphane and the oxygen atom of a coordinated acyl group, are efficient and robust homogeneous catalysts for the hydrolysis of ammonia– or amine–boranes under air to release hydrogen gas.3 Pyrazoles, five-membered heterocycles containing two adjacent nitrogen atoms, possess a wide range of useful properties4 and as ligands to transition metals they have attracted considerable interest.5 Pyrazoles possess a planar aromatic ring with a pyridine-type nitrogen suitable for coordination and an adjacent N–H group that can behave as a hydrogen-bond donor. In metal complexes of simple pyrazoles both inter- and intra-molecular hydrogen bonding can be found. Intermolecular forces have been successfully employed for the construction of supramolecular assemblies with various and interesting structures.6 Intramolecular hydrogen bonds are important in establishing the conformation of the complexes.7 Halide, hydroxide or peroxo species are known as hydrogen-bond acceptors towards metal-coordinated pyrazoles8 and the presence of strong hydrogen bonds between pyrazole and pyrazolate moieties in late transition metal complexes has also been reported.9 Therefore, we thought it interesting to prepare acylrhodium derivatives containing pyrazoles, which could form hydrogen bonds involving the oxygen atom of the acyl group and behave as catalysts for the hydrolysis of amine–boranes to release hydrogen. Hydrogen is a feasible alternative to fossil fuels but its safe storage and delivery still remains a challenge.10 Ammonia–borane (AB), a chemical hydrogen storage material with high hydrogen contents (H3N–BH3, 19.6 wt%), is considered a potential hydrogen source. The homogeneous transition-metal catalysed dehydrogenation of ammonia– or amine–boranes under mild conditions, which usually affords up to one equivalent of hydrogen gas and requires an inert atmosphere, has been the subject of recent intensive research.11 Hydrolysis reactions can afford up to three equivalents of hydrogen per equivalent of an amine–borane adduct. Transition metal heterogeneous systems,12 which often require an inert atmosphere and include highly efficient rhodium nanoparticles,13 allow fast H2 release from these amine–borane adducts at room temperature. The efficient homogeneously catalyzed version of this reaction has been reported more recently using the afore-mentioned hydridoirida-β-diketones or hydrido{(acylphosphane)(diphenylphosphinous acid)}rhodium(III) complexes,3 and also iridium-PNP complexes,14 coordinatively unsaturated iridium-carbene derivatives,15 dicarbonylruthenacyclic compounds16 or ruthenium-bipyridine-p-cymene complexes.17

We now report on the preparation of new hydrido(acylphosphane)rhodium(III) complexes containing pyrazole (Hpz), 3(5)-methylpyrazole (Hmpz) or 3,5-dimethylpyrazole (Hdmpz). DFT calculations were carried out aimed at identifying the thermodynamically most stable isomer of those complexes that could not be unambiguously characterized otherwise. The catalytic activity of the complexes to promote the hydrolysis of AB or amine–boranes in air to release hydrogen is also tested.

Results and discussion

The reaction of [{RhCl(COD)}2] (COD = 1,5-cyclooctadiene) with PPh2(o-C6H4CHO) (Rh/P = 1[thin space (1/6-em)]:[thin space (1/6-em)]1) in the presence of pyrazole (Rh/Hpz = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) leads to the displacement of COD and the formation of the hydridoacyl complex [RhHCl{PPh2(o-C6H4CO)}(Hpz)2] (1-Hpz), shown in Scheme 1i, in line with the well-known ability of the aldehyde-phosphane ligand to promote the chelate-assisted oxidative addition of the aldehyde to late transition metal complexes.2e The NMR spectra indicate the presence of only one species in solution. The 31P{1H} NMR spectrum contains only one doublet due to coupling with rhodium (J(Rh,P) = 166 Hz) at 79.0 ppm, the characteristic low field due to the five-member ring effect.18 The 1H NMR spectrum shows two sets of pyrazole signals at 8.47, 7.64 and 6.35 ppm or 7.00, 6.95 and 5.85 ppm, respectively, due to two non-equivalent ligands. These resonances are sharp and different for protons 3 and 5 of each pyrazole ring, indicating the absence of metallotropic tautomerization and both pyrazoles being firmly bonded to rhodium.8,19 In the low field region, the N–H resonances, also sharp, appear at 14.02 and 11.75 ppm and suggest the presence of hydrogen bonding to oxygen and chlorido. Adscription of the former to N–H⋯O bonding between the acyl group and the pyrazole trans to phosphorus appears likely.3d A doublet of doublet in the high field region, due to a hydrido bonded to rhodium (J(Rh,H) = 21.0 Hz) and cis to phosphorus (J(P,H) = 15.4 Hz) is observed. The chemical shift, at −15.79 ppm, is consistent with a hydrido trans to an electronegative atom. Due to the equivalent trans influence exerted by chlorido and N-donor ligands,20 it is difficult to determine the relative cis/trans disposition of the chlorido and the hydrido ligands. Also, the identification of the partners involved in the generation of intramolecular hydrogen bonds to the donor group of the pyrazole may be difficult due to the dependency of the hydrogen bond strength on its geometry, as well as on the nature of the substituent in the trans position to the acceptor ligand.21 DFT calculations (see below) indicate the structure depicted in Scheme 1i, with hydrido cis to chlorido and trans to pyrazole, to be the most stable for 1-Hpz.
image file: c5dt01705j-s1.tif
Scheme 1 In C6H6, 298 K. (i) R = R′ = H, Hpz; (ii) R = Me, R′ = H, Hmpz; (iii) R = R′ = Me, Hdmpz.

The reaction with 3(5)-methylpyrazole affords a mixture of two isomers shown in Scheme 1ii, 1a-Hmpz[thin space (1/6-em)]:[thin space (1/6-em)]1a′-Hmpz = 9[thin space (1/6-em)]:[thin space (1/6-em)]1. The spectroscopic features, similar to those observed for 1-Hpz, indicate the absence of metallotropic tautomerization. In this case we believe that isomers 1a-Hmpz and 1a′-Hmpz are linkage isomers. 1H-Pyrazoles exist in solution as mixtures of two tautomeric forms whereas in the solid state, with very few exceptions, only one tautomer is observed.4 When coordinated to transition metals, 3(5)-(R)pyrazoles (R = alkyl or aryl) usually appear in the 5-(R)pyrazole coordination mode, with the R group in the most distant position from the coordinating N-atom, in order to avoid steric congestion.22 The tautomerism of 3(5)-methylpyrazole, involving the proton transfer between nitrogen sites has been thoroughly studied.23 This ligand is unique and a few examples of complexes that contain both coordination modes, 5-methylpyrazole and 3-methylpyrazole, on the same metal atom have been reported.6a,24 Due to steric constraints, in our case the most abundant isomer, 1a-Hmpz, is expected to contain two less demanding 5-methylpyrazole ligands and the minor isomer, 1a′-Hmpz, could contain at least one 3-methylpyrazole ligand. DFT calculations (vide infra) confirm this assumption and show that isomer 1a′-Hmpz, containing both coordination modes 5- and 3-methylpyrazole, must be the minor linkage isomer.

When the reaction is performed using 3,5-dimethylpyrazole, a different behaviour is observed. The 1H and 31P{1H} NMR spectra at 298 K indicate the presence of almost equal amounts of the two species 1a-Hdmpz and 1b-Hdmpz in solution. The 31P{1H} NMR spectrum contains two doublets in the 77–78 ppm range and the 1H NMR spectrum shows two doublets of doublets in a very narrow range, between −15.5 and −16.5 ppm, due to the corresponding hydridos. We believe that both isomers differ in the group trans to hydrido as shown in Scheme 1iii, pyrazole in 1a-Hdmpz or chlorido in 1b-Hdmpz. DFT calculations (see below) confirm the similar stability of these complexes, 1a-Hdmpz and 1b-Hdmpz. Furthermore, in the 1H NMR spectrum, the resonances due to the pyrazole groups are extremely broad, suggesting that both isomers undergo metallotropic tautomerization and proton transfer in addition to metal exchange between nitrogen sites.6b,22a,b,25 The metallotropic tautomerization is much less frequent than that observed for the free ligands. On lowering the temperature, the signals due to the pyrazole become sharper, so that by 213 K the expected resonances due to four non-equivalent groups, two per isomer, with eight non-equivalent methyl groups, can be observed. These results show that at low temperatures the metallotropic tautomerization, which can be attributed to steric congestion, becomes inhibited. The resonances due to the phosphanes and the hydridos remain unaltered in the 298–213 K range.

Complexes 1 are stable at room temperature in the solid state but show low stability in solution, therefore single crystals for a crystallographic study could not be obtained. As mentioned earlier, theoretical calculations were performed to determine the particular disposition of the ligands around the central metal and to describe the intramolecular hydrogen bonds established in the obtained compounds. On the basis of its reported accuracy,26 the M062X level DFT calculations were carried out in order to identify the most stable isomers and rotamers, involving different hydrogen bonding options, of complexes 1. Comparison of PCM corrected relative ΔG values for all possible isomers and rotamers of compounds 1 at the M062X level was shown to be an appropriate tool to identify the thermodynamically most stable isomer/rotamer in each case. The absolute Gibbs free energies of the most stable structures (Tables 1 and SI-1, ESI) were used in each case to estimate the relative stability (ΔG) of the rotamers/isomers.

Table 1 Ranked PCM corrected relative free energy values (ΔG, kcal mol−1) calculated at the M062X/6-311++G(2d,2p)//B3LYP/6-31G(d,p)&LANL2DZ level of theory, for the most stable isomers/rotamers of compounds 1-Hpz, 1-Hmpz and 1-Hdmpz, respectively
Compound type   ΔGsolv cis/trans H,Cl-dispositiona R/R′ H-bond
trans cis trans cis
a cis or trans disposition of chlorido with respect to hydrido. b Nature of R and R′ in the pyrazole trans to phosphorus. c Nature of R and R′ in the pyrazole cis to phosphorus. d Acceptor group of the hydrogen bond established by the pyrazole trans to the phosphorus atom. e Acceptor group of the hydrogen bond established by the pyrazole cis to the phosphorus atom.
1-Hpz 1-Hpz 0.00 cis-H,Cl H/H H/H O-Acyl Chlorido
1-Hmpz 1a-Hmpz 0.00 cis-H,Cl H/Me H/Me O-Acyl Chlorido
1a′-Hmpz 1.35 cis-H,Cl H/Me Me/H O-Acyl Chlorido
1-Hdmpz 1b-Hdmpz 0.00 trans-H,Cl Me/Me Me/Me O-Acyl Chlorido
1b′-Hdmpz 0.02 trans-H,Cl Me/Me Me/Me Chlorido Chlorido
1a-Hdmpz 0.35 cis-H,Cl Me/Me Me/Me O-Acyl Chlorido


Up to 12 alternative isomers/rotamers with chlorido cis to phosphorus were considered in the case of 1-Hpz-type structures (Table SI-1). The isomer with chlorido trans to phosphorus and both Hpz ligands cis to each other was found to be 12.51 kcal mol−1 less stable than compound 1-Hpz (data not shown). In view of this significant thermodynamic destabilization, the theoretical study of additional trans (chlorido/phosphorus) isomers was found unnecessary. As observed in Table 1, the cis-H,Cl isomer 1-Hpz was identified as the most stable and unique structure, since the energetically closest structure, a trans-H,Cl isomer, turned out to be 1.86 kcal mol−1 less stable (Table SI-1), a significant energy difference, which rules out the possibility of the experimental observation of both species in the NMR spectra. This is in agreement with the experimental data. Compound 1-Hpz corresponds therefore to the cis-H,Cl isomer shown in Fig. SI-1a, where the pyrazoles trans and cis to phosphorus would establish hydrogen bonds with the acceptors O-acyl and chlorido, respectively. Any rotamer/isomer where the hydrido acts as a hydrogen bond acceptor was shown to be consistently unstable with respect to 1-Hpz (see Fig. SI-2). With respect to 1-Hmpz-type structures, the inclusion of all plausible tautomeric species led to the analyses of 48 structures, which, again, cover all plausible combinations of intramolecular hydrogen bonds (Table SI-1). As shown in Table 1, complexes 1a-Hmpz and 1a′-Hmpz were identified as the major and minor species (9[thin space (1/6-em)]:[thin space (1/6-em)]1), respectively, as concluded from the significant energy difference among the two (ΔG = 1.35 kcal mol−1). As observed in Fig. SI-1 (b and c), compounds 1a-Hmpz and 1a′-Hmpz are tautomeric species, where the pyrazole cis to phosphorus is methylated either on the 5- or 3-position, respectively. Both of them are cis-H,Cl isomers.

Experimental NMR data for 1-Hdmpz-type compounds revealed the existence of almost equimolar amounts of two species. In agreement with this, DFT results pointed out that two or even three species may exist simultaneously in solution. A slightly more abundant species (60%) was identified to consist of two rotamers of the trans-H,Cl isomer, 1b-Hdmpz and 1b′-Hdmpz (Fig. SI-1d and e), where the pyrazole trans to phosphorus would establish alternating hydrogen bonds with the O-acyl (1b-Hdmpz) and the chlorido (1b′-Hdmpz), respectively (ΔG = 0.02 kcal mol−1). The cis-H,Cl isomer 1a-Hdmpz (Fig. SI-1f) would be a slightly less abundant species (40%), with a moderate energy difference with the former (ΔG = 0.35 kcal mol−1). We thus observe that the difference in energy between trans-H,Cl and cis-H,Cl isomers is much smaller for the Hdmpz ligand than for the Hpz or Hmpz ligands, and allows the formation of both isomers in the Hdmpz case. In an attempt to identify the driving force that provokes the lower stability difference between isomers, steric clashes between the methyl groups present as substituents on the Hdmpz ligand (cis to phosphorus and also cis to acyl) with the neighbouring O-acyl were encountered. This steric effect would destabilize the cis-H,Cl geometry reducing the energy gap with the trans-H,Cl geometry, containing a Hdmpz ligand trans to acyl.

In a chloroform solution complexes 1 readily exchange the hydrido with chlorido to afford complexes [RhCl2{PPh2(o-C6H4CO)}(L)2] (L = Hpz, 2-Hpz; Hmpz, 2-Hmpz; Hdmpz, 2-Hdmpz). Such a reaction has several precedents.27 As shown in Scheme 2, in all the three cases single species are obtained, which are static in solution on the NMR time scale at room temperature. An X-ray diffraction study on complex 2-Hpz (see below) indicates cis chloridos and the existence of N–H⋯Cl hydrogen bonds.


image file: c5dt01705j-s2.tif
Scheme 2 In CHCl3, 298 K. 2-Hpz, R = R′ = H; 2-Hmpz, R = H, R′ = Me; 2-Hdmpz, R = R′ = Me.

It is remarkable that upon exchange of hydrido by chlorido, the 31P NMR signal moves significantly towards a higher field and appears in the 57–59 ppm range. A reduction in the coupling constant, J(Rh,P) = 136 Hz, is also observed. Their 13C{1H} NMR spectra show the resonance due to the acyl group in the low field, 222–228 ppm range, as a doublet of doublet due to coupling with rhodium (J(Rh,C) = 28 Hz) and with a cis phosphorus atom (J(P,C) = 3 Hz).

Complexes 1 react with triphenylphosphane that displaces the pyrazole ligand trans to phosphorus to give complexes [RhHCl{PPh2(o-C6H4CO)}(PPh3)L] (L = Hpz, 3-Hpz; Hmpz, 3-Hmpz; Hdmpz, 3a-Hdmpz and 3b-Hdmpz) as shown in Scheme 3i. The reaction is stereoselective for Hpz and Hmpz affording only one isomer, which, in accord with the DFT studies performed for complexes 1, we assume to be the cis-H,Cl species, while Hdmpz affords an almost equimolar mixture of two isomers that, as shown by the NMR spectra, can only differ in the group trans to the hydrido, pyrazole in 3a-Hdmpz or chlorido in 3b-Hdmpz.


image file: c5dt01705j-s3.tif
Scheme 3 In CH2Cl2, 298 K. (i) PPh3: 3-Hpz, R = R′ = H; 3-Hmpz, R = Me, R′ = H; 3a-Hdmpz and 3b-Hdmpz, R = R′ = Me. (ii) PPh2(O)H: 4-Hpz, R = R′ = H; 4a-Hmpz, R = Me, R′ = H; 4a′-Hmpz, R = H, R′ = Me; 4-Hdmpz, R = R ′ = Me.

The 31P{1H} NMR spectra show two doublets of doublets in the 65–62 and the 44–38 ppm range, due to the chelating acyl-phosphane and PPh3 respectively. The J(P,P) coupling constants, in the 376–379 Hz range, indicate two mutually trans phosphorus atoms. In the 13C{1H} NMR spectra, the resonance due to the acyl group appears as a doublet (J(Rh,C) in the range 28–35 Hz) in the expected low field (232–235 ppm range),28 and the resonances due to the 3 and 5 carbon atoms of the pyrazole rings are different, indicating the absence of tautomerization in all cases. In the 1H NMR spectra, the hydrido appears as a doublet of doublet of doublet due to coupling to rhodium and to two cis phosphorus atoms, in the −14.0 to −14.5 ppm range, in a slightly lower field than in complexes 1. Sharp singlets in the 11.0–11.8 ppm range are also observed, due to the pyrazole being involved in N–H⋯Cl hydrogen bonding, and the resonances for protons or methyl groups on the 3 and 5 carbon atoms of the pyrazole rings are different. An X-ray diffraction study of 3a-Hdmpz confirms the structure depicted in Scheme 3i.

The reaction of complexes 1 with diphenylphosphane oxide affords complexes [RhHCl{PPh2(o-C6H4CO)} {(PPh2O)H}L] (L = Hpz, 4-Hpz; Hmpz, 4a-Hmpz and 4a′-Hmpz; Hdmpz, 4-Hdmpz) shown in Scheme 3ii. As in the reaction with PPh3, 1-Hpz affords a single isomer 4-Hpz. 1-Hmpz gives two linkage isomers, 4a-Hmpz[thin space (1/6-em)]:[thin space (1/6-em)]4a′-Hmpz = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, as in the starting material. At variance with the reaction with PPh3, 1-Hdmpz, which consists of an almost equimolar mixture of two isomers, affords now a single isomer 4-Hdmpz. The spectroscopic features of complexes 4 are similar to those reported for complexes 3 (see Experimental). In the low field region of the 1H NMR spectra, a sharp singlet is observed, in the 10.7–11.2 ppm range due to N–H⋯Cl. Another sharp singlet is observed in the 12.0–12.7 ppm range that can be attributed to the presence of a hydrogen bond between the oxygen atom of the diphenylphosphinous acid and the oxygen atom of the coordinated acyl group.3d,29 This O–H⋯O hydrogen bond appears to contribute to the stereoselectivity of this reaction to afford a single isomer also in the Hdmpz case. DFT calculations (see below) show this isomer to be the cis-H,Cl species.

In accordance with these observations, the preparation of complexes [RhHCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHO)-κP}L] (L = Hpz, 5-Hpz; Hdmpz, 5a-Hdmpz and 5b-Hdmpz), by reaction of [{RhCl(COD)}2] with PPh2(o-C6H4CHO) in the presence of pyrazoles (Rh/Hpz/P = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2), leads to a mixture of two isomers 5a-Hdmpz[thin space (1/6-em)]:[thin space (1/6-em)]5b-Hdmpz = 7[thin space (1/6-em)]:[thin space (1/6-em)]3 when using Hdmpz while Hpz leads to a single isomer, 5-Hpz (Scheme 4i). Complexes 5 contain a P-monodentate phosphane–aldehyde ligand with a dangling aldehyde group less suitable for hydrogen bond formation. The spectroscopic features of complexes 5 are as expected (see Experimental). When using Hmpz, the NMR spectra show the formation of a mixture of the corresponding complex with a dangling aldehyde, 5-Hmpz, which could not be obtained pure, and a hydroxyalkyl-phosphane derivative 6-Hmpz, which is the product of the insertion of an aldehyde into the Rh–H bond.27e,30


image file: c5dt01705j-s4.tif
Scheme 4 (i) In C6H6, 298 K. 5-Hpz, R = R′ = H; 5a-Hdmpz and 5b-Hdmpz, R = R′ = CH3. (ii) In C6H6, 353 K, 5 min. (iii) In C6H6, 353 K, 90 min. 7-Hpz, R = R′ = H; 7-Hmpz, R = H, R′ = CH3; 7-Hdmpz, R = R′ = CH3.

Complex [RhCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHOH)}(Hmpz)], 6-Hmpz, could be obtained in refluxing benzene (5 min) as shown in Scheme 4ii. It is a mixture of two isomers 6a-Hmpz[thin space (1/6-em)]:[thin space (1/6-em)]6b-Hmpz in a 6[thin space (1/6-em)]:[thin space (1/6-em)]4 ratio, analogous to our previously reported pyridine containing species [RhCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHOH)}(py)].27e The 13C{1H} NMR spectrum shows a doublet of doublets in the 90–95 ppm range (J(Rh,C) = 22; J(P,C) = 100 Hz) showing the formation of hydroxyalkyl groups bonded to rhodium and trans to the phosphorus atom of the acyl-phosphane chelate. Accordingly, the 31P{1H} NMR spectrum shows two doublets of doublets for each isomer with phosphanes mutually cis (J(P,P) in the 16–19 Hz range). Isomers 6a-Hmpz and 6b-Hmpz differ in the group, chlorido or pyrazole, trans to acyl. Sharp singlets at 12.09 and 11.49 ppm in the 1H NMR spectrum, can be attributed to N–H⋯Cl hydrogen bonds. Longer reaction times led to the formation of the diacyl derivative [RhCl{PPh2(o-C6H4CO)}2(Hmpz)] (7-Hmpz) as a single isomer, with hydrogen evolution (see Scheme 4iii). Hpz and Hdmpz also afford complexes 7-Hpz and 7-Hdmpz respectively. The 31P{1H} NMR spectra of complexes 7 show ABX spin patterns with J(P,P) = 340 Hz, indicating the trans disposition of both phosphorus atoms while the 13C{1H} NMR spectra show in the low field region two doublets due to two inequivalent acyl groups trans to chlorido and pyrazole respectively. An X-ray diffraction study of 7-Hpz confirms the structure depicted in Scheme 4iii.

The same theoretical study as before was carried out to identify the particular isomers/rotamers obtained for 3-Hdmpz-, 4-Hdmpz- and 5-Hdmpz-type complexes. Phosphorus atoms were considered to be in the trans position to each other in all cases since NMR data for the mentioned complexes were unambiguous in this respect. Structures with intramolecular hydrogen bonds established between Hdmpz and any acceptor atom other than chlorido were not considered in these cases (Table SI-2), since our DFT studies on complexes 1 have shown that pyrazole ligands cis to phosphorus form hydrogen bonds exclusively with chlorido. DFT results (Table 2) for 3-Hdmpz-type complexes revealed that two species with a moderate energy difference (0.60 kcal mol−1) may exist in a 3[thin space (1/6-em)]:[thin space (1/6-em)]7 ratio, namely, the cis-H,Cl 3a-Hdmpz and the trans-H,Cl 3b-Hdmpz isomers (Fig. SI-3a and b, respectively), the trans-H,Cl isomer being the most abundant species. The latter result is similar to that previously described for 1-Hdmpz and agrees with the experimental NMR data, which show the presence of two different species in solution. In the case of 4-Hdmpz-type complexes, and in agreement with the NMR data, the cis-H,Cl isomer 4-Hdmpz was predicted to be the only species in solution (Fig. SI-3c). The energy difference (2.99 kcal mol−1) with the energetically closest structure ruled out any other possibility (Table SI-2). As depicted in Fig. SI-3c, the acidic hydrogen atom belonging to the diphenylphosphinous acid would establish a hydrogen bond with the O-acyl moiety. In this case, a combination of a lesser steric hindrance of PPh2OH versus PPh3 along with fixation of the O-acyl atom by O–H⋯O bond formation, relieves the steric clashes between pyrazole methyl substituents and the neighbouring O-acyl, thus stabilizing the cis-H,Cl conformation. Theoretical calculations on 5-Hdmpz-type complexes revealed that the 5a-Hdmpzcis-H,Cl isomer is 0.82 kcal mol−1 more stable that the trans-H,Cl isomer 5b-Hdmpz (Table 2) and that both isomers may exist in an 8[thin space (1/6-em)]:[thin space (1/6-em)]2 ratio. The latter results are in agreement with the corresponding NMR data, which contain signals due to two different species in solution. As observed in Fig. SI-3d and SI-3e a quantitatively significant N–H⋯Cl intramolecular bond is established, while a second hydrogen bond, which may be quantitatively much less meaningful, is also observed between the hydrogen atom of the aldehyde moiety and either the hydrido (5a-Hdmpz) or the O-acyl group (5b-Hdmpz). A visual inspection of 5a-Hdmpz and 5b-Hdmpz suggests a suboptimal spatial disposition of the acyl-phosphane ligand in the trans-H,Cl isomer, derived from its attempt to establish a H-bond with the H atom of the aldehyde moiety of the PPh2(o-C6H4CHO) ligand, which could rationalize the observed isomer ratios.

Table 2 Ranked PCM corrected relative free energy values (ΔG, kcal mol−1) calculated at the M062X/6-311++G(2d,2p)//B3LYP/6-31G(d,p)&LANL2DZ level of theory, for the most stable isomers/rotamers of compounds 3-Hdmpz, 4-Hdmpz and 5-Hdmpz, respectively
Compound type   ΔGsolv cis/transa H-bond
OH/CHOb Pyrazoled
a cis or trans disposition of chlorido with respect to hydrido. b Acceptor group of the hydrogen bond established by OH in 4-Hdmpz or H in CHO of 5-Hdmpz. c Considering the moderate polarity of the C–H bond in aldehydes, the quantitative significance of the indicated hydrogen bonds may be modest. Still they could be identified based on the directionality of the groups involved. The oxygen atom in CHO of 5-Hdmpz was found to not be able to establish hydrogen bonds. d Acceptor group of the hydrogen bond established by the pyrazole.
3-Hdmpz 3b-Hdmpz 0.00 trans-H,Cl Chlorido
3a-Hdmpz 0.60 cis-H,Cl Chlorido
4-Hdmpz 4-Hdmpz 0.00 cis-H,Cl O-Acyl Chlorido
5-Hdmpz 5a-Hdmpz 0.00 cis-H,Cl Hydridoc Chlorido
5b-Hdmpz 0.82 trans-H,Cl O-Acylc Chlorido


The X-ray structures of 2-Hpz (dichloromethane adduct), 3a-Hdmpz (chloroform adduct) and 7-Hpz have been determined. Selected bond distances and angles are listed in Table 3. Fig. 1–3 show molecular drawings for these complexes together with the atomic labelling scheme used. All the three complexes show a slightly distorted octahedral arrangement, with two positions occupied by the phosphorus and carbon atoms of an acyl-phosphane ligand and a chlorido ligand trans to the acyl carbon atom. 2-Hpz also contains two mutually cis pyrazole ligands that are trans to a second chlorido and to phosphorus. Compounds 3a-Hdmpz or 7-Hpz contain a single pyrazole ligand and hydrido and triphenylphosphane or a second acyl-phosphane ligand, respectively. In the latter compounds the pyrazole ligand lies trans to hydrido or to an acyl carbon atom, respectively, thus allowing the phosphorus atoms to adopt a trans arrangement. The Rh1–N1 distances: 2.046(2), 2.205(2) and 2.221(6) in 2-Hpz, 3a-Hdmpz and 7-Hpz respectively and the Rh1–N3 distance of 2.136(2) Å in 2-Hpz reflect the trans influence order of the ligands: chlorido < phosphane < hydrido ≈ acyl.31 In all the three structures the pyrazole moieties and the chlorido ligands are mutually cis and the formation of N–H⋯Cl hydrogen bonds is observed. The pyrazole protons involved in these hydrogen bonds could be clearly located as Fourier peaks in a difference map. The N⋯Cl distances, between 3.017 Å and 3.067 Å, for all the three complexes are as expected for moderately strong hydrogen bonds, however the N–H⋯Cl angles, between 128(1) and 140(1)°, are in the lower range of the corresponding values.7,32 Nevertheless, the importance of the hydrogen bonding can also be assessed by the dihedral angle between two planes, that of the pyrazole ring (including the N–H bond) and the plane defined by the coordinated N atom of pyrazole, chlorido, and their trans atoms to the metal and the metal, which should be close to 0°.8 In 2-Hpz these angles are very small, 4.8 and 6.6° for the planes involving the N4 and N2 atoms, respectively, indicating that the N–H vector is directed in order to maximize the intramolecular hydrogen bonding. For 3a-Hdmpz and 7-Hpz the values of these dihedral angles are 9.8 and 13.7°, respectively. In 2-Hpz the planes of the heterocycles are almost perpendicular (θ = 86.8°), and this feature may be related to the presence of the hydrogen bonds. Also in 2-Hpz the chlorido trans to acyl forms another hydrogen bond, N2–H2N⋯Cl1′′, weaker, with the pyrazole trans to chlorido in a neighbouring molecule (N2–Cl1′′ = 3.241(2) Å). This intermolecular interaction leads to the formation of dimers.


image file: c5dt01705j-f1.tif
Fig. 1 ORTEP view of 2-Hpz showing the atomic numbering and the intramolecular hydrogen bonds (30% probability ellipsoids). The hydrogen atoms except for two have been omitted for clarity.

image file: c5dt01705j-f2.tif
Fig. 2 ORTEP view of 3a-Hdmpz showing the atomic numbering and the intramolecular hydrogen bond (30% probability ellipsoids). The hydrogen atoms except for two have been omitted for clarity.

image file: c5dt01705j-f3.tif
Fig. 3 ORTEP view of 7-Hpz showing the atomic numbering and the intramolecular hydrogen bond (30% probability ellipsoids). The hydrogen atoms except for one have been omitted for clarity.
Table 3 Selected bond lengths (Å) and angles (°) for 2-Hpz, 3a-Hdmpz and 7-Hpz, including the intramolecular hydrogen bonding
2-Hpz 3a-Hdmpz 7-Hpz
Rh1–N1 2.046(2) Rh1–N1 2.205(2) Rh1–N1 2.221(6)
Rh1–N3 2.136(2) Rh1–P2 2.3443(9) Rh1–P2 2.351(3)
Rh1–P1 2.2415(6) Rh1–P1 2.2628(9) Rh1–P1 2.318(3)
Rh1–Cl1 2.5491(5) Rh1–Cl1 2.5073(6) Rh1–Cl1 2.523(2)
Rh1–Cl2 2.3614(5) Rh1–H1 1.401(4) Rh1–C20 2.004(8)
Rh1–C1 1.992(2) Rh1–C1 1.990(2) Rh1–C1 2.01(1)
C1–O1 1.208(3) C1–O1 1.216(4) C1–O1 1.21(1)
N3–N4 1.349(3) C20–O2 1.20(1)
N1–N2 1.347(2) N1–N2 1.361(2) N1–N2 1.35(1)
 
Cl1–Rh1–C1 179.02(7) Cl1–Rh1–C1 176.94(9) Cl1–Rh1–C1 173.1(3)
Cl2–Rh1–N1 175.71(5) P1–Rh1–P2 171.98(3) P1–Rh1–P2 168.35(9)
Cl1–Rh1–N1 89.14(5) Cl1–Rh1–N1 87.71(7) Cl1–Rh1–N1 85.8(2)
P1–Rh1–N3 177.95(5) N1–Rh1–H1 178(2) C20–Rh1–N1 173.1(3)
 
N2⋯Cl1 3.031(2) N2⋯Cl1 3.067(3) N2⋯Cl1 3.017(8)
H2N⋯Cl1 2.3073(5) H2N⋯Cl1 2.4582(8) H2N⋯Cl1 2.399(2)
N2–H2N⋯Cl1 140.3(1) N2–H2N⋯Cl1 128.4(5) N2–H2N⋯Cl1 129.3(5)
N4⋯Cl2 3.034(2)
H4N⋯Cl2 2.3376(6)
N4–H4N⋯Cl2 130.3(1)


Complexes 1 and 4, containing N–H⋯O or O–H⋯O hydrogen bonds respectively involving the acyl groups, were tried as catalysts for the hydrolysis of AB, TBAB (tert-butylamine-borane), and DMAB (dimethylamine-borane) in THF/H2O = 1/1 mixtures, at 298 K, in the presence of air. According to eqn (1), up to three equivalents of hydrogen per equivalent of an adduct may be released.

 
RR′HN–BH3 + 4H2O → RR′NH2+ + B(OH)3 + OH + 3H2(1)

With an initial AB, TBAB or DMAB concentration of 0.46 M and a 0.5 mol% catalyst loading of complexes 1, the hydrolysis and the release of 75–90% of the maximum hydrogen content that could be produced, was observed within 25 min (Table 4). The nature of the product gas was verified as H2 by NMR of the catalytic hydrolysis of AB in a THF-d8/D2O mixture, which shows the gradual disappearance of the quartet centred at δ = 1.37 ppm (1JB–H = 92 Hz) due to AB and the formation of HD at δ = 4.52 (t) ppm (1JD–H = 43 Hz). Blank tests under the same experimental conditions, but in the absence of a catalyst, show that the hydrolysis of AB, TBAB or DMAB affords a maximum amount of only 10% of the hydrogen content after 4 min Fig. SI-4).

Table 4 % Conversion and time required, for the hydrolysis of ammonia– and amine–boranes with complexes 1 or 4 (0.5 mol%) in THF/H2O = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio, at 298 K
Substrate Catalyst % Conv. Time/s Catalyst % Conv. Time/s
AB 1-Hpz 75 1500 4-Hpz 92 3000
1-Hmpz 90 1500 4-Hmpz 90 3300
1-Hdmpz 88 1500 4-Hdmpz 85 3600
TBAB 1-Hpz 81 1500 4-Hpz 78 3600
1-Hmpz 75 1800 4-Hmpz 82 3600
1-Hdmpz 79 1500 4-Hdmpz 82 3000
DMAB 1-Hpz 75 2400 4-Hpz 74 8400
1-Hmpz 81 2400 4-Hmpz 80 5400
1-Hdmpz 85 1800 4-Hdmpz 88 4200


The catalytic activity for complexes 1-Hpz, 1-Hmpz and 1-Hdmpz was similar. As shown in Fig. 4, the amount of hydrogen evolved from all the three substrates was analogous. When using complexes 4 a similar hydrogen release requires at least 50 min reaction times (Table 4). After completion of the catalytic reactions, the 11B NMR spectra of the remaining solutions contain only singlets in the 16–20 ppm range, which indicates the transformation of the borane adducts into borate species.13a,b In the presence of Hg similar results were obtained, confirming the homogeneous nature of the catalytic reaction.13cFig. 4 shows the kinetic profiles when using 1-Hmpz as a catalyst. The kinetic profiles obtained in the hydrolysis of AB and TBAB can be considered to follow a first order reaction rate model with respect to the [substrate]. When the experimental points are fitted to a first order law by non-linear least-squares regression, the values of the rate constants kobs = (1.89 ± 0.06) × 10−3 and (1.98 ± 0.10) × 10−3 s−1 can be obtained respectively for AB and TBAB. The kinetic profile of the hydrolysis of DMAB does not show a defined kinetic model. The activity shown by complexes 1 is comparable to that observed for acylrhodium complexes containing pyridine and diphenylphosphinous acid that hydrolyse AB and TBAB at 298 K with rate constants kobs = (1.54 ± 0.07) × 10−3 and (0.95 ± 0.05) × 10−3 s−1, respectively.3d


image file: c5dt01705j-f4.tif
Fig. 4 Hydrogen release from AB (○), TBAB (□), or DMAB (△) as substrates, with an initial substrate concentration of 0.46 M, in the presence of 0.5 mol% of 1-Hdmpz as the catalyst in THF/H2O mixtures. T, 298 K. Open marks represent the experimental points. Solid lines represent the regression line.

Conclusions

The stereoselective formation of hydrido(acylphosphane)(chlorido)(pyrazole)rhodium(III) complexes, with chlorido trans to acyl and hydrido trans to pyrazole and stabilized by the formation of hydrogen bonds, is accomplished. The formation of N–H⋯Cl and N–H⋯O hydrogen bonding occurs and the former is preferred to the latter. Steric hindrance lowers the stereoselectivity of the reaction by lowering the energy difference between isomers. In close agreement with the experimental NMR data, DFT calculations were able to predict the isomer/rotamer proportions and the particular intramolecular hydrogen bonds established in each case. The bis(pyrazole) complexes are efficient homogeneous catalysts for the hydrolysis of ammonia– or amine–borane adducts in air to produce hydrogen.

Experimental section

General procedures

The preparation of the metal complexes was carried out at room temperature, 298 K, under nitrogen by standard Schlenk techniques. [{RhCl(COD)}2]33 and PPh2(o-C6H4CHO)34 were prepared as previously reported. Microanalysis was carried out with a Leco CHNS–932 microanalyser. IR spectra were recorded with a Nicolet FTIR 510 spectrophotometer in the range 4000–400 cm−1 using KBr pellets. NMR spectra were recorded with Bruker Avance DPX 300 or Bruker Avance 500 spectrometers. 1H and 13C{1H} (TMS internal standard), 31P{1H} (H3PO4 external standard) NMR spectra were measured in CDCl3 solutions. The resonances of the pyrazoles were assigned by performing COSY and HSQC 2D experiments.
Preparation of [RhClH{PPh2(o-C6H4CO)}(L)2] (L = Hpz, 1-Hpz; Hmpz, 1-Hmpz; Hdmpz, 1-Hdmpz). To a benzene solution of [{RhCl(COD)}2] (0.06 mmol) was added a stoichiometric amount (0.24 mmol) of the corresponding ligand, whereupon a yellow solid was formed. Addition of PPh2(o-C6H4CHO) (0.12 mmol) and stirring at room temperature for 1 h afforded a yellow solution. Addition of hexane gave pale yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Data for1-Hpz. Yield: 60%. IR (KBr, cm−1): 3292 (s), ν(N–H); 2063 (s), ν(Rh–H); 1629 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C25H23ClN4OPRh; C 53.16, H 4.11, N 9.92; found C 52.89, H 4.17, N 9.92. 1H NMR: (CDCl3): δ −15.79 (dd, 1H, J(Rh,H) = 21.0 Hz, J(P,H) = 15.4 Hz, RhH); Hpz: 14.02 (s, 1H, NH⋯O); 11.75 (s, 1H, NH); 8.47 (s, 1H, CH); 7.64 (s, 1H, CH); 7.00 (s, 1H, CH); 6.95 (s, 1H, CH); 6.35 (s, 1H, CH); 5.85 (s, 1H, CH); phenyl: 8.23, 7.96, 7.51, 7.37, 7.20, 7.14 and 7.10 (m, 14H). 31P{1H} NMR (CDCl3): δ 79.0 (d, J(Rh,P) = 166 Hz). Data for1-Hmpz. Yield: 60%. IR (KBr, cm−1): 3212 (m), ν(N–H); 2044 (m), ν(Rh–H); 1634 (m), ν(C[double bond, length as m-dash]O). Anal. Calcd for C27H27ClN4OPRh; C 54.70, H 4.59, N 9.45; found C 54.80, H 4.51, N 9.17. Data for1a-Hmpz. 1H NMR: (CDCl3): δ −15.78 (dd, 1H, J(Rh,H) = 21.0 Hz, J(P,H) = 15.0 Hz, RhH); Hmpz: 13.50 (s, 1H, NH⋯O); 11.12 (s, 1H, NH); 8.31 (s, 1H, CH); 6.88 (s, 1H, CH); 6.08 (s, 1H, CH); 5.54 (s, 1H, CH); 2.40 (s, 3H, CH3); 1.86 (s, 3H, CH3); phenyl: 8.24, 7.94, 7.65, 7.49, 7.23 and 7.10 (m, 14H). 31P{1H} NMR (CDCl3): δ 79.2 (d, J(Rh,P) = 166 Hz). Data for1a′-Hmpz. 1H NMR: (CDCl3): δ −16.27 (dd, 1H, J(Rh,H) = 21.6 Hz, J(P,H) = 15.1 Hz, RhH); Hmpz: 13.88 (s, 1H, NH⋯O); 11.96 (s, 1H, NH); 2.42 (s, 3H, CH3); 1.60 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 79.0 (d, J(Rh,P) = 163 Hz). Data for1-Hdmpz. Yield: 40%. IR (KBr, cm−1): 3258 (m), ν(N–H); 2066 (m), ν(Rh–H); 1637 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C29H31ClN4OPRh·0.25(C6H6); C 57.20, H 5.12, N 8.75; found C 57.76, H 4.93, N 8.37. 1H NMR: (CDCl3, 213 K): δ −16.14 (dd, 1H, J(Rh,H) = 30.7 Hz, J(P,H) = 14.6 Hz, RhH); −15.63 (dd, 1H, J(Rh,H) = 20.9 Hz, J(P,H) = 15.3 Hz, RhH); Hdmpz: 12.17 (s, 1H, NH⋯O); 11.17 (s, 1H, NH); 10.63 (s, H, NH); 10.55 (s, H, NH); 5.76 (s, 2H, CH); 5.56 (s, 1H, CH), 5.23 (s, 1H, CH); 2.15 (s, 3H, CH3); 2.13 (s, 3H, CH3); 2.02 (s, 3H, CH3); 2.00 (s, 3H, CH3); 1.83 (s, 3H, CH3); 1.79 (s, 6H, CH3); 1.33 (s, 3H, CH3); phenyl: 8.07, 7.84, 7.77, 7.69, 7.47, 7.41 and 7.12 (m, 28H). 31P{1H} NMR (CDCl3, 298 K): δ 77.9 (d, J(Rh,P) = 166 Hz); 77.5 (d, J(Rh,P) = 162 Hz).
Preparation of [RhCl2{PPh2(o-C6H4CO)}(L)2] (L = Hpz, 2-Hpz; Hmpz, 2-Hmpz; Hdmpz, 2-Hdmpz). A chloroform solution of the corresponding 1 (0.05 mmol) was stirred at room temperature for 48 h. Addition of hexane gave yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Data for2-Hpz. Yield: 60%. IR (KBr, cm−1): 3295 (s), ν(N–H); 1664 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C25H22Cl2N4OPRh; C 50.11, H 3.70, N 9.35; found C 50.34, H 3.64, N 9.05. 1H NMR: (CDCl3): δ Hpz: 12.73 (s, 1H, NH); 12.69 (s, 1H, NH); 8.11 (s, 1H, CH); 7.63 (s, 1H, CH); 7.03 (s, 1H, CH); 6.86 (s, 1H, CH); 6.36 (s, 1H, CH); 5.78 (s, 1H, CH); phenyl: 8.46, 7.80, 7.54, 7.50 and 7.17 (m, 14H). 31P{1H} NMR (CDCl3): δ 59.2 (d, J(Rh,P) = 137 Hz). 13C{1H} NMR (CDCl3): δ 228.5 (dd, J(Rh,CO) = 28 Hz, J(P,CO) = 3 Hz, C[double bond, length as m-dash]O); Hpz: 140.4 (s, CH); 132.2 (s, CH); 129.4 (s, CH); 124.8 (s, CH); 107.4 (s, CH); 106.4 (s, CH); phenyl: 139.8 (d, J = 3 Hz), 136.2 (d, J = 10 Hz), 132.6 (d, J = 7 Hz), 131.8 (d, J = 3 Hz), 131.6 (d, J = 10 Hz), 131.3 (s), 130.2 (d, J = 3 Hz), 129.2 (d, J = 3 Hz), 128.3 (d, J = 11 Hz), 128.1 (d, J = 11 Hz) (18C). Data for2-Hmpz. Yield: 57%. IR (KBr, cm−1): 3239 (m), 3189 (s), ν(N–H); 1662 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C27H26Cl2N4OPRh; C 51.70, H 4.18, N 8.93; found C 51.31, H 4.19, N 8.53. 1H NMR: (CDCl3): δ Hmpz: 12.20 (s, 1H, NH); 12.11 (s, 1H, NH); 7.83 (s, 1H, CH); 6.77 (s, 1H, CH); 6.07 (s, 1H, CH); 5.49 (s, 1H, CH); 2.35 (s, 3H, CH3); 1.92 (s, 3H, CH3); phenyl: 8.47, 8.08, 7.79, 7.60, 7.49, 7.20 and 7.01 (m, 14H). 31P{1H} NMR (CDCl3): δ 58.7 (d, J(Rh,P) = 136 Hz). 13C{1H} NMR (CDCl3): δ 227.7 (dd, J(Rh,CO) = 28 Hz, J(P,CO) = 4 Hz, C[double bond, length as m-dash]O); Hmpz: 140.8 (s, CH); 140.7 (s, CH); 106.7 (s, CH); 105.9 (s, CH); 11.2 (s, CH3); 10.9 (s, CH3); phenyl: 140.3 (s), 136.2 (d, J = 10 Hz), 132.4 (d, J = 6 Hz), 132.4 (s), 132.0 (s), 131.6 (d, J = 9 Hz), 131.2 (s), 129.9 (d, J = 3 Hz), 128.2 (d, J = 11 Hz), 127.7 (d, J = 11 Hz) (18C). Data for2-Hdmpz. Yield: 59%. IR (KBr, cm−1): 3371 (s), ν(N–H); 1673 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C29H30Cl2N4OPRh·0.25(CHCl3); C 51.27, H 4.45, N 8.17; found C 51.31, H 4.36, N 8.31. 1H NMR: (CDCl3): δ Hdmpz: 12.63 (s, 1H, NH); 11.14 (s, 1H, NH); 5.80 (s, 1H, CH); 5.27 (s, 1H, CH); 2.25 (s, 3H, CH3); 2.09 (s, 3H, CH3); 1.66 (s, 3H, CH3); 1.31 (s, 3H, CH3); phenyl: 8.09, 7.92, 7.67, 7.48, 7.39 and 7.11(m, 14H). 31P{1H} NMR (CDCl3): δ 57.4 (d, J(Rh,P) = 137 Hz). 13C{1H} NMR (CDCl3): δ 221.8 (dd, J(Rh,CO) = 27 Hz, J(P,CO) = 3 Hz, C[double bond, length as m-dash]O); Hdmpz: 107.3 (s, CH); 106.4 (s, CH); 14.4 (s, CH3); 11.8 (s, CH3); 11.1 (s, CH3); 10.8 (s, CH3); phenyl: 140.5 (s), 139.6 (s), 135.3 (d, J = 9 Hz), 132.7 (s), 32.5 (s), 131.9 (d, J = 9 Hz), 131.7 (d, J = 6 Hz), 131.1 (d, J = 3 Hz), 130.3 (d, J = 3 Hz), 128.0 (d, J = 10 Hz), 127.9 (d, J = 11 Hz), 125.2 (d, J = 16 Hz) (18C).
Preparation of [RhHCl{PPh2(o-C6H4CO)}(PPh3)L] (L = Hpz, 3-Hpz; Hmpz, 3-Hmpz; Hdmpz, 3-Hdmpz). To a dichloromethane solution of the corresponding 1 (0.05 mmol) was added a slight excess (0.06 mmol) of triphenylphosphane. Stirring for 1 h at room temperature afforded a yellow solution. Addition of hexane gave yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Data for3-Hpz. Yield: 42%. IR (KBr, cm−1): 3231 (s), ν(N–H); 2046 (m), ν(Rh–H); 1621 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C40H34N2OP2Rh·0.25(CH2Cl2); C 61.96, H 4.46, N 3.59; found C 62.28, H 4.21, N 3.67. 1H NMR (CDCl3): δ −14.05 (ddd, 1H, J(Rh,H) = 17.1 Hz, J(P,H) = 11.9, 4.0 Hz, RhH); Hpz: 11.80 (s, 1H, NH); 6.69 (s, 1H, CH); 6.60 (s, 1H, CH); 5.54 (s, 1H, CH); phenyl: 8.31, 7.72, 7.58, 7.46, 7.42, 7.24, 7.10 and 7.00 (m, 29H). 31P{1H} NMR (CDCl3): δ 62.3 (dd, J(P,P) = 377 Hz, J(Rh,P) = 138 Hz, PPh2(o-C6H4CO)); 44.2 (dd, J(Rh,P) = 124 Hz, Ph3P). 13C{1H} NMR (CDCl3): δ 234.9 (d, J(Rh,C) = 32 Hz, C[double bond, length as m-dash]O); Hpz: 138.5 (s, CH); 129.1 (s, CH); 105.4 (s, CH); phenyl: 135.4 (d, J = 12 Hz), 134.2 (d, J = 11 Hz), 131.6 (d, J = 10 Hz), 131.3 (d, J = 5 Hz), 131.1 (d, J = 6 Hz), 130.8 (d, J = 4 Hz), 128.7 (d, J = 11 Hz), 126.7 (s) and 122.5 (d, J = 17 Hz) (36C). Data for3-Hmpz. Yield: 58%. IR (KBr, cm−1): 3242 (s), ν(N–H); 2036 (m), ν(Rh–H); 1623 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C41H36N2OP2Rh; C 63.70, H 4.69, N 3.62; found C 64.11, H 4.70, N 3.68. 1H NMR (CDCl3): δ −14.02 (ddd, 1H, J(Rh,H) = 19.0 Hz, J(P,H) = 12.3 Hz, 4.6 Hz, RhH); Hmpz: 11.25 (s, 1H, NH); 6.41 (s, 1H, CH); 5.25 (s, 1H, CH); 1.75 (s, 3H, CH3); phenyl: 8.32, 7.74, 7.58, 7.46, 7.41, 7.23, 7.13, 7.05 and 7.00 (m, 29H). 31P{1H} NMR (CDCl3): δ 62.6 (dd, J(P,P) = 379 Hz, J(Rh,P) = 137 Hz, PPh2(o-C6H4CO)); 44.4 (dd, J(Rh,P) = 124 Hz, Ph3P). 13C{1H} NMR (CDCl3, 303 K): δ 235.0 (d, J(Rh,C) = 35 Hz, C[double bond, length as m-dash]O); Hmpz: 139.1 (s, CH); 129.1 (s, C–CH3); 104.7 (s, CH); 10.7 (s, CH3); phenyl: 137.4 (s), 135.5 (d, J = 12 Hz), 134.3 (d, J = 11 Hz), 133.9 (s), 133.5 (s), 131.8 (d, J = 10 Hz), 131.3 (d, J = 5 Hz), 131.1 (d, J = 6 Hz), 130.8 (d, J = 4 Hz), 128.7 (d, J = 11 Hz) and 127.6 (d, J = 10 Hz) (36C). Data for3-Hdmpz. Yield: 58%. IR (KBr, cm−1): 3219 (s), ν(N–H); 2053 (m), ν(Rh–H); 1631 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C42H38ClN2OP2Rh; C 64.09, H 4.87, N 3.56; found C 63.92, H 4.32, N 3.80. 1H NMR (CDCl3): δ −14.73 (m, 1H, RhH); –14.51 (dt, 1H, J(Rh,H) = 27.3 Hz, J(P,H) = 7.7 Hz, RhH); Hdmpz: 11.50 (s, 1H, NH); 11.03 (s, 1H, NH); 5.35 (s, 1H, CH); 5.08 (s, 1H, CH); 1.79 (s, 3H, CH3); 1.72 (s, 3H, CH3); 1.48 (s, 3H, CH3); 0.98 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 65.1 (dd, J(P,P) = 379 Hz, J(Rh,P) = 133 Hz, PPh2(o-C6H4CO); 37.9 (dd, J(Rh,P) = 127 Hz, Ph3P) and δ 63.8 (dd, J(P,P) = 376 Hz, J(Rh,P) = 140 Hz, PPh2(o-C6H4CO)); 41.3 (dd, J(Rh,P) = 124 Hz, Ph3P). 13C{1H} NMR (CDCl3): δ 233.9 (d, J(Rh,C) = 28 Hz, C[double bond, length as m-dash]O); 232.0 (d, J(Rh,C) = 35 Hz, C[double bond, length as m-dash]O); Hdmpz: 129.8 (s, C–CH3); 129.1 (s, C–CH3); 105.8 (s, CH); 105.3 (s, CH); 13.7 (s, CH3); 13.3 (s, CH3); 10.7 (s, CH3); 10.6 (s, CH3).
Preparation of [RhHCl{PPh2(o-C6H4CO)}(PPh2OH)L] (L = Hpz, 4-Hpz; Hmpz, 4-Hmpz; Hdmpz, 4-Hdmpz). To a benzene solution of the corresponding 1 (0.05 mmol) was added diphenylphosphane oxide (0.055 mmol). After stirring for 1 h under reflux at 80 °C, the addition of hexane gave yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Data for4-Hpz. Yield: 54%. IR (KBr, cm−1): 3260(m), ν(N–H); 2072 (m), ν(Rh–H); 1625 (m), ν(C[double bond, length as m-dash]O). Anal. Calcd for C34H30ClN2O2P2Rh; C 58.43, H 4.33, N 4.01; found C 57.99, H 4.89, N 4.46. 1H NMR (CDCl3): δ −14.20 (dt, 1H, J(Rh,H) = 19.8 Hz, J(P,H) = 9.5 Hz, RhH); Hpz: 12.00 (s, 1H, PPh2P(O)H); 10.99 (s, 1H, NH); 6.82 (s, 1H, CH); 6.62 (s, 1H, CH); 5.77 (s, 1H, CH); phenyl: 8.21, 7.93, 7.68, 7.52, 7.44, 7.31, 7.12 and 7.02 (m, 24H). 31P{1H} NMR (CDCl3): δ 98.4 (dd, J(P,P) = 419 Hz, J(Rh,P) = 126 Hz, Ph2P(O)H); 60.6 (dd, J(Rh,P) = 120 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 246.8 (d, J(Rh,C) = 37 Hz, C[double bond, length as m-dash]O); Hpz: 138.7 (s, CH); 127.6 (s, CH); 106.0 (s, CH); phenyl: 135.2 (d, J = 13 Hz), 132.6 (d, J = 14 Hz), 131.6 (d, J = 10 Hz), 131.5 (d, J = 4 Hz), 131.4 (s), 131.2 (s), 130.5 (s), 130.4 (s), 129.3 (s), 128.9 (s), 128.8 (s), 128.0 (d, J = 11 Hz), 127.8 (d, J = 10 Hz), 127.0 (d, J = 10 Hz), 123.6 (d, J = 17 Hz) (30C). Data for4-Hmpz. Yield: 58%. IR (KBr, cm−1): 3260 (m), ν(N–H); 2018 (m), ν(Rh–H); 1626 (m), ν(C[double bond, length as m-dash]O). Anal. Calcd for C35H32ClN2O2P2Rh; C 58.96, H 4.52, N 3.93; found C 58.85, H 4.25, N 4.42. Data for4a-Hmpz. 1H NMR: (CDCl3): δ −14.20 (dt, 1H, J(Rh,H) = 19.7 Hz, J(P,H) = 9.5 Hz, RhH); Hmpz: 12.11 (s, 1H, PPh2P(O)H); 10.42 (s, 1H, NH); 6.62 (s, 1H, CH); 5.46 (s, 1H, CH); 1.62 (s, 3H, CH3); phenyl: 8.21, 7.92, 7.68, 7.52, 7.44, 7.36, 7.13 and 7.04 (m, 24H). 31P{1H} NMR (CDCl3): δ 98.4 (dd, J(P,P) = 423 Hz, J(Rh,P) = 126 Hz, Ph2P(O)H); 60.8 (dd, J(Rh,P) = 120 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 247.0 (d, J(Rh,C) = 37 Hz, C[double bond, length as m-dash]O); Hmpz: 139.2 (s, CH); 129.2 (s, C–CH3); 105.5 (s, CH); 10.6 (s, CH3); phenyl: 138.6 (s), 135.2 (d, J = 13 Hz), 132.7 (d, J = 14 Hz), 132.5 (d, J = 5 Hz), 131.8 (d, J = 10 Hz), 131.5 (d, J = 4 Hz), 131.4 (s), 131.2 (s), 130.7 (s), 130.6 (d, J = 5 Hz), 128.8 (s), 128.7 (s), 128.0 (d, J = 11 Hz), 127.6 (d, J = 9 Hz), 126.8 (d, J = 10 Hz), 123.5 (d, J = 16 Hz) (30C). Data for4a′-Hmpz. 1H NMR: (CDCl3): δ −14.57 (m, 1H, RhH); 12.22 (s, 1H, PPh2P(O)H); Hmpz: 11.16 (s, 1H, NH); 6.44 (s, 1H, CH); 5.41 (s, 1H, CH); 1.56 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 98.8 (dd, J(P,P) = 419 Hz, J(Rh,P) = 125 Hz, Ph2P(O)H); 60.6 (dd, J(Rh,P) = 121 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 138.9 (s, CH); Hmpz: 129.4 (s, C–CH3); 105.9 (s, CH); 10.9 (s, CH3). Data for4-Hdmpz. Yield: 53%. IR (KBr, cm−1): 3242 (m), ν(N–H); 2033 (m), ν(Rh–H); 1624 (m), ν(C[double bond, length as m-dash]O). Anal. Calcd for C36H34ClN2O2P2Rh; C 59.48, H 4.71, N 3.85; found C 59.39, H 4.67, N 4.02. 1H NMR (CDCl3): δ −14.53 (dt, 1H, J(Rh,H) = 19.3 Hz, J(P,H) = 8.6 Hz, RhH); Hdmpz: 12.68 (s, 1H, PPh2P(O)H); 10.67 (s, 1H, NH); 5.12 (s, 1H, CH); 1.55 (s, 3H, CH3); 1.46 (s, 3H, CH3); phenyl: 8.21, 8.15, 7.94, 7.72, 7.46, 7.17 and 7.08 (m, 24H). 31P{1H} NMR (CDCl3): δ 98.9 (dd, J(P,P) = 422 Hz, J(Rh,P) = 126 Hz, Ph2P(O)H); 60.7 (dd, J(Rh,P) = 120 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 246.0 (d, J(Rh,C) = 38 Hz, C[double bond, length as m-dash]O); Hdmpz: 105.5 (s, CH); 13.4 (s, CH3); 10.4 (s, CH3); phenyl: 138.2 (s), 135.0 (d, J = 13 Hz), 132.8 (d, J = 14 Hz), 132.2 (d, J = 5 Hz), 131.8 (s), 131.8 (d, J = 10 Hz), 131.6 (d, J = 4 Hz), 131.4 (s), 131.0 (s), 130.4 (s), 130.3 (s), 129.2 (s), 128.7 (s), 128.6 (s), 127.9 (d, J = 11 Hz), 127.5 (d, J = 10 Hz), 126.8 (d, J = 10 Hz), 123.5 (d, J = 16 Hz) (30C).
Preparation of [RhHCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHO)-κP}L] (L = Hpz, 5-Hpz; Hdmpz, 5-Hdmpz). To a benzene solution of [{RhCl(COD)}2] (0.06 mmol) was added a stoichiometric amount (0.12 mmol) of the corresponding ligand, whereupon a yellow solid was formed. Addition of PPh2(o-C6H4CHO) (0.30 mmol) and stirring at room temperature for 1 h followed by addition of hexane gave pale yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Data for5-Hpz. Yield: 45%. IR (KBr, cm−1): 3234 (s), ν(N–H); 2043 (m), ν(Rh–H); 1687 (s), ν(HC[double bond, length as m-dash]O); 1622 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C41H34N2O2P2Rh; C 62.57, H 4.35, N 3.56; found C 62.81, H 4.47, N 3.43. 1H NMR (CDCl3): δ −14.10 (ddd, 1H, J(Rh,H) = 17.2 Hz, J(P,H) = 12.1 Hz, 4.0 Hz, RhH); Hpz: 11.63 (s, 1H, NH); 10.27 (s, 1H, H–CO); 6.72 (s, 1H, CH); 6.57 (s, 1H, CH); 5.60 (s, 1H, CH); phenyl: 8.34, 7.70, 7.58, 7.47, 7.29, 7.22 and 7.13 (m, 28H). 31P{1H} NMR (CDCl3): δ 65.4 (dd, J(P,P) = 370 Hz, J(Rh,P) = 141 Hz, PPh2(o-C6H4CO)); 40.2 (dd, J(Rh,P) = 123 Hz, PPh2(o-C6H4CHO)). 13C{1H} NMR (CDCl3): δ 234.5 (d, J(Rh,C) = 31 Hz, C[double bond, length as m-dash]O); 191.7 (s, HC[double bond, length as m-dash]O); Hpz: 138.8 (s, CH); 127.8 (s, CH); 105.9 (s, CH); phenyl: 138.1 (d, J = 7 Hz), 135.5 (d, J = 13 Hz), 135.4 (d, J = 5 Hz), 134.9 (d, J = 13 Hz), 134.5 (d, J = 12 Hz), 132.3 (d, J = 7 Hz), 131.6 (s), 131.3 (s), 130.8 (d, J = 4 Hz), 129.9 (s), 129.5 (s), 129.3 (s), 128.8 (d, J = 11 Hz), 128.3 (d, J = 10 Hz), 128.1 (d, J = 10 Hz), 127.9 (d, J = 10 Hz), 122.7 (d, J = 17 Hz) (36C). Data for5-Hdmpz. Yield: 60%. IR (KBr, cm−1): 3202 (s), ν(N–H); 2057 (m), ν(Rh–H); 1687 (s), ν(HC[double bond, length as m-dash]O); 1626, ν(C[double bond, length as m-dash]O). Anal. Calcd for C43H38N2O2P2Rh; C 63.36, H 4.70, N 3.44; found C 63.22, H 4.97, N 3.27. Data for5a-Hdmpz. 1H NMR (CDCl3): δ −14.74 (m, 1H, RhH); 11.41 (s, 1H, NH); Hdmpz: 10.26 (s, 1H, H–CO); 5.15 (s, 1H, CH); 1.68 (s, 3H, CH3); 0.96 (s, 3H, CH3); phenyl: 8.30, 7.97, 7.82, 7.74, 7.64, 7.56, 7.44, 7.31, 7.12 and 7.06 (m, 28H). 31P{1H} NMR (CDCl3): δ 66.6 (dd, J(P,P) = 369 Hz, J(Rh,P) = 142 Hz, PPh2(o-C6H4CO)); 36.8 (dd, J(Rh,P) = 124 Hz, PPh2(o-C6H4CHO)). 13C{1H} NMR (CDCl3): δ 231.5 (d, J(Rh,C) = 32 Hz, C[double bond, length as m-dash]O); 190.4 (s, HC[double bond, length as m-dash]O); Hdmpz: 106.3 (s, CH); 12.7 (s, CH3); 10.6 (s, CH3). Data for5b-Hdmpz. 1H NMR (CDCl3): δ −14.37 (ddd, 1H, J(Rh,H) = 26.0 Hz, J(P,H) = 10.4 Hz, J(P,H) = 5.8 Hz, RhH); Hdmpz: 11.08 (s, 1H, NH); 9.91 (s, 1H, H–CO); 5.34 (s, 1H, CH); 1.83 (s, 3H, CH3); 1.16 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 65.8 (dd, J(P,P) = 375 Hz, J(Rh,P) = 137 Hz, PPh2(o-C6H4CO)); 35.9 (dd, J(Rh,P) = 127 Hz, PPh2(o-C6H4CHO)). 13C{1H} NMR (CDCl3): δ 229.2 (d, J(Rh,C) = 37 Hz, C[double bond, length as m-dash]O); 190.0 (s, HC[double bond, length as m-dash]O); Hdmpz: 105.3 (s, CH); 12.9 (s, CH3); 10.7 (s, CH3).
Preparation of [RhCl{PPh2(o-C6H4CO)}{PPh2(o-C6H4CHOH)}(Hmpz)] (6). To a benzene solution of [{RhCl(COD)}2] (0.06 mmol) was added a stoichiometric amount (0.12 mmol) of the corresponding ligand, whereupon a yellow solid was formed. Addition of PPh2(o-C6H4CHO) (0.30 mmol) and stirring for 5 min under reflux at 80 °C, afforded a yellow solution. Addition of hexane gave pale yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Yield: 40%. IR (KBr, cm−1): 3205 (m), ν(N–H); 1633 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C42H35N2O2P2Rh; C 62.97, H 4.53, N 3.50; found C 62.37, H 4.41, N 3.91. Data for6a. 1H NMR (CDCl3): δ Hmpz: 12.09 (s, 1H, NH); 6.51 (s, 1H, CH); 5.59 (s, 1H, CH); 1.94 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 55.2 (dd, J(P,P) = 19 Hz, J(Rh,P) = 152 Hz, PPh2(o-C6H4CHOH)); 29.3 (dd, J(Rh,P) = 83 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 239.8 (dd, J(Rh,C) = 35 Hz, J(P,C) = 9 Hz, C[double bond, length as m-dash]O); 90.6 (dd, J(Rh,C) = 22 Hz, J(P,C) = 103 Hz, CHOH); 10.9 (s, CH3). Data for6b. 1H NMR (CDCl3): δ Hmpz: 11.49 (s, 1H, NH); 6.09 (s, 1H, CH); 5.69 (s, 1H, CH); 1.87 (s, 3H, CH3). 31P{1H} NMR (CDCl3): δ 55.5 (dd, J(P,P) = 16 Hz, J(Rh,P) = 168 Hz, PPh2(o-C6H4CHOH)); 32.5 (dd, J(Rh,P) = 87 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 229.9 (d, J(Rh,C) = 33 Hz, C[double bond, length as m-dash]O); 94.2 (dd, J(Rh,C) = 22 Hz, J(P,C) = 99 Hz, CHOH); 11.0 (s, CH3).
Preparation of [RhCl{PPh2(o-C6H4CO)}2L] (L = Hpz, 7-Hpz; Hmpz, 7-Hmpz; Hdmpz, 7-Hdmpz). To a benzene solution of [{RhCl(COD)}2] (0.06 mmol) was added a stoichiometric amount (0.12 mmol) of the corresponding ligand, whereupon a yellow solid was formed. Addition of PPh2(o-C6H4CHO) (0.30 mmol) and stirring for 90 min under reflux at 80 °C, afforded a yellow solution. Addition of hexane gave pale yellow precipitates, which were filtered off, washed with hexane and dried under vacuum. Complex 7-Hmpz was recrystallized from dichloromethane/diethyl ether. Data for7-Hpz. Yield: 73%. IR (KBr, cm−1): 3225 (s), ν(N–H); 1641 (s), 1624 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C41H32N2O2P2Rh·0.25(C6H6); C 63.45, H 4.20, N 3.48; found C 63.40, H 4.34, N 3.95. 1H NMR (CDCl3): δ Hpz: 11.48 (s, 1H, NH); 7.08 (s, 1H, CH); 6.85 (s, 1H, CH); 5.86 (s, 1H, CH); phenyl: 8.17, 7.93, 7.71, 7.62, 7.49, 7.35, 7.25 and 7.05 (m, 28H). 31P{1H} NMR (CDCl3): δ 61.0 (dd, J(P,P) = 337 Hz, J(Rh,P) = 150 Hz, PPh2(o-C6H4CO)); 51.5 (dd, J(Rh,P) = 138 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 233.5 (d, br, J(Rh,C) = 33 Hz, C[double bond, length as m-dash]O); Hpz: 139.7 (s, CH); 127.4 (s, CH); 106.1 (s, CH); phenyl: 136.9 (d, J = 11 Hz), 133.7 (d, J = 10 Hz), 133.4 (d, J = 10 Hz), 132.0 (s), 131.5 (s), 131.2 (d, J = 7 Hz), 130.9 (s), 130.7 (s), 129.8 (d, J = 11 Hz), 129.0–127.7 (m), 127.5 (s), 123.9 (d, J = 7 Hz) and 123.8 (d, J = 6 Hz) (36C). Data for7-Hmpz. Yield: 62%. IR (KBr, cm−1): 3216 (s), ν(N–H); 1642 (s), 1621 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C42H34N2O2P2Rh·0.5(CH2Cl2); C, 60.66, H, 4.19, N, 3.33; found C 60.59, H 4.44, N 3.37. 1H NMR (CDCl3): δ Hmpz: 10.92 (s, 1H, NH); 6.88 (s, 1H, CH); 5.55 (s, 1H, CH); 1.81 (s, 3H, CH3); phenyl: 8.15, 7.92, 7.69, 7.58, 7.47, 7.27 and 7.01 (m, 28H). 31P{1H} NMR (CDCl3): δ 60.9 (dd, J(P,P) = 340 Hz, J(Rh,P) = 152 Hz, PPh2(o-C6H4CO)); 51.4 (dd, J(Rh,P) = 139 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 233.6 (d, J(Rh,C) = 33 Hz, C[double bond, length as m-dash]O); 232.8 (d, J(Rh,C) = 36 Hz, C[double bond, length as m-dash]O); Hmpz: 140.4 (s, CH); 105.5 (s, CH); 10.7 (s, CH3); phenyl: 136.9 (d, J = 10 Hz), 133.8 (d, J = 10 Hz), 133.5 (d, J = 10 Hz), 132.5 (s), 132.3–131.9 (m), 131.7 (d, J = 5. Hz), 131.5 (s), 131.4 (s), 131.3 (s), 130.9 (d, J = 12 Hz), 129.9 (s), 129.7 (s), 129.2–128.4 (m), 128.1 (d, J = 10 Hz), 127.8 (d, J = 10 Hz), 127.4 (d, J = 10 Hz), 124.1 (d, J = 19 Hz) and 123.1 (d, J = 18 Hz) (36C). Data for7-Hdmpz. Yield: 92%. IR (KBr, cm−1): 3194 (m), ν(N–H); 1642 (s), 1623 (s), ν(C[double bond, length as m-dash]O). Anal. Calcd for C43H36N2O2P2Rh; C 63.52, H 4.46, N 3.44; found C 62.70, H 4.24, N 3.53. 1H NMR (CDCl3): δ Hdmpz: 11.16 (s, 1H, NH); 5.41 (s, 1H, CH); 1.80 (s, 3H, CH3); 1.39 (s, 3H, CH3); phenyl: 8.09, 7.88, 7.67, 7.59, 7.47, 7.34, 7.21, 7.13, 6.97 and 6.83 (m, 28H). 31P{1H} NMR (CDCl3): δ 58.3 (dd, J(P,P) = 340 Hz, J(Rh,P) = 155 Hz, PPh2(o-C6H4CO)); 52.9 (dd, J(Rh,P) = 143 Hz, PPh2(o-C6H4CO)). 13C{1H} NMR (CDCl3): δ 229.3 (d, (Rh,C) = 34 Hz, C[double bond, length as m-dash]O); 227.4 (d, J(Rh,C) = 36 Hz, C[double bond, length as m-dash]O); Hdmpz: 106.3 (s, CH); 13.7 (s, CH3); 10.6 (s, CH3); phenyl: 136.2–135.4 (m), 133.7 (s), 133.5 (d, J = 9 Hz), 133.2 (s), 132.4 (d, J = 9 Hz), 132.2–131.2 (m), 130.3 (d, J = 11 Hz), 129.8 (s), 129.4, 129.1–128.3 (m), 127.8 (d, J = 9.2 Hz), 127.2 (d, J = 6.1 Hz), 125.2 (d, J = 15 Hz) and 125.0 (d, J = 16 Hz) (36C).

Crystallographic refinement and structure solution

X-ray quality crystals were obtained by vapour diffusion of diethyl ether onto dichloromethane (2-Hpz and 7-Hpz,) or chloroform (3a-Hdmpz) solutions at −20 °C. Single crystals of suitable dimensions were used for data collection. For compounds 2-Hpz and 3a-Hdmpz, diffraction intensities were collected on an Agilent Technologies Super-Nova diffractometer, which was equipped with monochromated Mo Kα radiation (λ = 0.71[thin space (1/6-em)]073 Å) and an Eos CCD detector at 100 K. For compound 7-Hpz, intensity data were collected on a Bruker AXS APEX CCD area detector equipped with graphite monochromated Mo Kα radiation (λ = 0.71[thin space (1/6-em)]073 Å) by applying the ω-scan method. The structures were solved by direct methods and refined with full-matrix least-squares calculations on F2 using the program SHELXL-97.35 Anisotropic temperature factors were assigned to all atoms except for hydrogen atoms, which ride their parent atoms with an isotropic temperature factor arbitrarily chosen as 1.2 times that of the respective parent. Hydrogens from nitrogen atoms pertaining to pyrazole rings were located and not fixed. Attempts to solve disorder problems with the crystallization molecule failed in compound 2Hpz. Instead, a new set of F2 (hkl) values with the contribution from solvent molecules withdrawn was obtained by the SQUEEZE procedure implemented in PLATON-94.36 Final R(F), wR(F2) and goodness of fit agreement factors, details of the data collection and analysis can be found in Table SI-3. CCDC reference numbers for the structures of 2-Hpz, 3a-Hdmpz, and 7-Hpz are 1062495, 1062496 and 1062497, respectively.

DFT calculations

Quantum chemical calculations were carried out with the Gaussian 09 series of programs.37 Full geometry optimizations of each species were performed in the gas phase by employing the hybrid density functional B3LYP38 with the 6-31G(d,p) basis set for nonmetal atoms together with the LANL2DZ39 for the metal. When available, X-ray structures were used as starting geometries. Energy values were then refined by single-point calculations on the B3LYP/LANL2DZ geometries at the M062X/6-311++G(2d,2p) level of theory and solvent (benzene) contributions to free energy were considered by means of polarizable continuum model40 (PCM) calculations. The nature of the stationary points was verified by analytical computations of harmonic vibrational frequencies at the M062X/6-311++G(2d,2p) level of theory. A pressure of 1 atm and a temperature of 298.15 K were assumed in the calculations. The theoretical ratio of the isomers/rotamers present in solution was predicted assuming a Boltzmann distribution.41

Dehydrogenation of RR′R′′N–BH3 with complexes 1 or 4

A typical dehydrogenation experiment is described here for the hydrolysis of H3N–BH3 using 2.70 ml of THF/H2O = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixtures and 0.5 mol% catalyst loading: a solution of 38.8 mg (1.25 mmol) of H3N–BH3 in H2O was prepared in a round bottom 40 ml flask fitted with a gas outlet and with a side arm sealed with a tight-fitting septum cap. The flask was connected via the gas outlet to a water-filled gas burette. A solution of 0.006 mmol of the corresponding catalyst in dry THF was then syringed through the septum, magnetic stirring was connected and timing started. Gas evolution began immediately and the amount of gas evolved was determined periodically by measuring the displacement of water in the burette. Volumes were measured at atmospheric pressure and 298 K.

Acknowledgements

Partial financial support from Ministerio de Economía y Competitividad (CTQ2011-24859), Gobierno Vasco, Diputación Foral de Gipuzkoa and Universidad del País Vasco (UPV/EHU) is gratefully acknowledged. SGI/IZO-SGIker UPV/EHU is gratefully acknowledged for a generous allocation of computational resources. Dr C. Ubide is acknowledged for helpful discussions. SA is grateful to Gobierno Vasco for a scholarship.

References

  1. (a) A. H. Roy, C. P. Lenges and M. Brookhart, J. Am. Chem. Soc., 2007, 129, 2082–2093 CrossRef CAS PubMed; (b) Z. Shen, H. A. Khan and V. M. Dong, J. Am. Chem. Soc., 2008, 130, 2916–2917 CrossRef CAS PubMed; (c) S. Omura, T. Fukuyama, J. Horiguchi, Y. Murakami and I. Ryu, J. Am. Chem. Soc., 2008, 130, 14094–14095 CrossRef CAS PubMed; (d) G. L. Moxham, H. Randell-Sly, S. K. Brayshaw, A. S. Weller and M. C. Willis, Chem. – Eur. J., 2008, 14, 8383–8397 CrossRef CAS PubMed; (e) R. J. Pawley, M. A. Huertos, G. C. Lloyd-Jones, A. S. Weller and M. C. Willis, Organometallics, 2012, 31, 5650–5659 CrossRef CAS.
  2. (a) P. J. Alaimo, B. A. Arndtsen and R. G. Bergman, J. Am. Chem. Soc., 1997, 119, 5269–5270 CrossRef CAS; (b) M. A. Esteruelas, Y. A. Hernández, A. M. López, M. Oliván and L. Rubio, Organometallics, 2008, 27, 799–802 CrossRef CAS; (c) T. Iwai, T. Fujihara and Y. Tsuji, Chem. Commun., 2008, 6215–6217 RSC; (d) R. Beck, U. Flörke and H. F. Klein, Inorg. Chem., 2009, 48, 1416–1422 CrossRef CAS PubMed; (e) M. A. Garralda, Dalton Trans., 2009, 3635–3645 RSC; (f) J. J. Adams, N. Arulsamy and D. M. Roddick, Organometallics, 2012, 31, 1439–1447 CrossRef CAS; (g) A. E. Roa, V. Salazar, J. López-Serrano, E. Oñate, M. Paneque and M. L. Poveda, Organometallics, 2012, 31, 716–721 CrossRef CAS.
  3. (a) M. A. Garralda, R. Hernández, L. Ibarlucea, E. Pinilla and M. R. Torres, Organometallics, 2003, 22, 3600–3603 CrossRef CAS; (b) R. Ciganda, M. A. Garralda, L. Ibarlucea, E. Pinilla and M. R. Torres, Dalton Trans., 2010, 39, 7226–7229 RSC; (c) M. A. Garralda, C. Mendicute-Fierro, A. Rodríguez-Diéguez, J. M. Seco, C. Ubide and I. Zumeta, Dalton Trans., 2013, 42, 11652–11660 RSC; (d) V. San Nacianceno, L. Ibarlucea, C. Mendicute-Fierro, A. Rodríguez-Diéguez, J. M. Seco, I. Zumeta, C. Ubide and M. A. Garralda, Organometallics, 2014, 33, 6044–6052 CrossRef.
  4. S. Fustero, M. Sánchez-Roselló, P. Barrio and A. Simón-Fuentes, Chem. Rev., 2011, 111, 6984–7034 CrossRef CAS PubMed.
  5. (a) S. Trofimenko, Chem. Rev., 1972, 72, 497–509 CrossRef CAS; (b) S. Trofimenko, Prog. Inorg. Chem., 1986, 34, 115–210 CrossRef CAS PubMed.
  6. (a) D. H. Choi, J. H. Yoon, J. H. Lim, H. C. Kim and C. S. Hong, Inorg. Chem., 2006, 45, 5947–5952 CrossRef CAS PubMed; (b) J. Pérez and L. Riera, Eur. J. Inorg. Chem., 2009, 4913–4925 CrossRef PubMed.
  7. D. A. Johnson, W. C. Deese and A. W. Cordes, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1981, 37, 2220–2223 CrossRef.
  8. D. B. Grotjahn, S. Van, D. Combs, D. A. Lev, C. Schneider, C. D. Incarvito, K. C. Lam, G. Rossi, A. L. Rheingold, M. Rideout, C. Meyer, G. Hernández and L. Mejorado, Inorg. Chem., 2003, 42, 3347–3355 CrossRef CAS PubMed and ref. therein.
  9. (a) D. Carmona, L. A. Oro, M. P. Lamata, J. Elguero, M. C. Apreda, C. Foces-Foces and F. H. Cano, Angew. Chem., Int. Ed. Engl., 1986, 25, 1114–1115 CrossRef PubMed; (b) G. A. Ardizzoia, S. Brenna, G. LaMonica, A. Maspero, N. Masiocchi and M. Moret, Inorg. Chem., 2002, 41, 610–614 CrossRef CAS; (c) G. Albertin, S. Antoniutti, J. Castro and S. García-Fontán, Eur. J. Inorg. Chem., 2011, 510–520 CrossRef CAS PubMed.
  10. (a) G. W. Crabtree, M. S. Dresselhaus and M. V. Buchanan, Phys. Today, 2004, 57, 39–44 CrossRef CAS PubMed; (b) U. Eberle, M. Felderhoff and F. Schüth, Angew. Chem., Int. Ed., 2009, 48, 6608–6630 CrossRef CAS PubMed; (c) A. F. Dalebrook, W. Gan, M. Grasemann, S. Moret and G. Laurenczy, Chem. Commun., 2013, 49, 8735–8751 RSC.
  11. (a) T. B. Marder, Angew. Chem., Int. Ed., 2007, 46, 8116–8118 CrossRef CAS PubMed; (b) F. H. Stephens, V. Pons and R. T. Baker, Dalton Trans., 2007, 2613–2626 RSC; (c) P. Wang and X. D. Kang, Dalton Trans., 2008, 5400–5413 RSC; (d) C. W. Hamilton, R. T. Baker, A. Staubitz and I. Manners, Chem. Soc. Rev., 2009, 38, 279–293 RSC; (e) N. C. Smythe and J. C. Gordon, Eur. J. Inorg. Chem., 2010, 509–521 CrossRef CAS PubMed; (f) A. Staubitz, A. P. M. Robertson and I. Manners, Chem. Rev., 2010, 110, 4079–4124 CrossRef CAS PubMed; (g) T. Hügle, M. Hartl and D. Lentz, Chem. – Eur. J., 2011, 17, 10184–10207 CrossRef PubMed; (h) Z. Huang and T. Autrey, Energy Environ. Sci., 2012, 5, 9257–9268 RSC; (i) J. A. Buss, G. A. Edouard, C. Cheng, J. Shi and T. Agapie, J. Am. Chem. Soc., 2014, 136, 11272–11275 CrossRef CAS PubMed.
  12. For recent examples see: (a) H. L. Jiiang and Q. Xu, Catal. Today, 2011, 170, 56–63 CrossRef PubMed; (b) M. Yadav and Q. Xu, Energy Environ. Sci., 2012, 5, 9698–9725 RSC; (c) P. Z. Li, A. Aijaz and Q. Xu, Angew. Chem., Int. Ed., 2012, 51, 6753–6756 CrossRef CAS PubMed; (d) M. Kaya, M. Zahmakiran, S. Özkar and M. Volkan, ACS Appl. Mater. Interfaces, 2012, 4, 3866–3873 CrossRef CAS PubMed; (e) B. Kilic, S. Sencanli and O. Metin, J. Mol. Catal., 2012, 361–362, 104–110 CrossRef CAS PubMed; (f) S. Caliskan, M. Zahmakiran, F. Durap and S. Özkar, Dalton Trans., 2012, 41, 4976–4984 RSC; (g) G. Chen, S. Desinan, R. Rosei, F. Rosei and D. Ma, Chem. – Eur. J., 2012, 18, 7925–7930 CrossRef CAS PubMed; (h) Y. Du, N. Cao, L. Yang, W. Luo and G. Cheng, New J. Chem., 2013, 37, 3035–3042 RSC; (i) X. Wang, D. Liu, S. Song and H. Zhang, Chem. – Eur. J., 2013, 19, 8082–8086 CrossRef CAS PubMed; (j) S. Akbayrak and S. Ozkar, Dalton Trans., 2014, 43, 1797–1805 RSC; (k) F. Qiu, G. Liu, L. Li, Y. Wang, C. Xu, C. An, C. Chen, Y. Xu, Y. Huang, Y. Wang, L. Jiao and H. Yuan, Chem. – Eur. J., 2014, 20, 505–509 CrossRef CAS PubMed; (l) H. Cheng, T. Kamegawa, K. Mori and H. Yamashita, Angew. Chem., Int. Ed., 2014, 53, 2910–2914 CrossRef CAS PubMed.
  13. (a) M. Chandra and Q. Xu, J. Power Sources, 2006, 156, 190–194 CrossRef CAS PubMed; (b) M. Chandra and Q. Xu, J. Power Sources, 2007, 168, 135–142 CrossRef CAS PubMed; (c) T. J. Clark, G. R. Whittell and I. Manners, Inorg. Chem., 2007, 46, 7522–7527 CrossRef CAS PubMed; (d) C. W. Yoon and L. G. Sneddon, J. Am. Chem. Soc., 2006, 128, 13992–13993 CrossRef CAS PubMed; (e) M. Fetz, R. Gerber, O. Blacque and C. M. Frech, Chem. – Eur. J., 2011, 17, 4732–4736 CrossRef CAS PubMed.
  14. T. W. Graham, C. W. Tsang, X. Chen, R. Guo, W. Jia, S. M. Lu, C. Sui-Seng, C. B. Ewart, A. Lough, D. Amoroso and K. Abdur-Rashid, Angew. Chem., Int. Ed., 2010, 49, 8708–8711 CrossRef CAS PubMed.
  15. (a) G. C. Fortman, A. M. Z. Slawin and S. P. Nolan, Organometallics, 2011, 30, 5487–5492 CrossRef CAS; (b) D. J. Nelson, B. J. Truscott, J. D. Egbert and S. P. Nolan, Organometallics, 2013, 32, 3769–3772 CrossRef CAS.
  16. C. Boulho and J. P. Djukic, Dalton Trans., 2010, 39, 8893–8905 RSC.
  17. M. Muñoz-Olasagasti, A. Telleria, J. Pérez-Miqueo, M. A. Garralda and Z. Freixa, Dalton Trans., 2014, 42, 11652–11660 Search PubMed.
  18. P. E. Garrou, Chem. Rev., 1981, 81, 229–266 CrossRef CAS.
  19. (a) G. A. Ardizzoia, G. LaMonica, A. Maspero, M. Moret and N. Massiocchi, Eur. J. Inorg. Chem., 1998, 1503–1512 CrossRef CAS; (b) J. Pérez, D. Morales, S. Nieto, L. Riera, V. Riera and D. Miguel, Dalton Trans., 2005, 884–888 RSC.
  20. T. G. Appleton, H. C. Clark and L. I. Manzer, Coord. Chem. Rev., 1973, 10, 335–422 CrossRef CAS.
  21. E. Peris, J. C. Lee Jr., J. R. Rambo, O. Eisenstein and R. H. Crabtree, J. Am. Chem. Soc., 1995, 117, 3485–3491 CrossRef CAS.
  22. See for example: (a) M. Cano, J. A. Campo, J. V. Heras, J. Lafuente, C. Rivas and E. Pinilla, Polyhedron, 1995, 14, 1139–1147 CrossRef CAS; (b) M. Cano, J. V. Heras, M. Maeso, M. Alvaro, R. Fernández, E. Pinilla, J. A. Campo and A. Monge, J. Organomet. Chem., 1997, 534, 159–172 CrossRef CAS; (c) S. Nieto, J. Pérez, V. Riera, D. Miguel and C. Alvarez, Chem. Commun., 2005, 546–548 RSC; (d) S. Nieto, J. Pérez, L. Riera, D. Miguel, J. A. Golen and A. L. Rheingold, Inorg. Chem., 2007, 46, 3407–3418 CrossRef CAS PubMed; (e) C. J. Carrasco, F. Montilla, E. Álvarez, C. Mealli, G. Mancac and A. Galindo, Dalton Trans., 2014, 43, 13711–13730 RSC.
  23. (a) J. P. Wibaut and J. W. P. Boon, Helv. Chim. Acta, 1961, 44, 1171–1190 CrossRef CAS PubMed; (b) J. Catalán, M. Sánchez-Cabezudo, J. L. G. de Paz, J. Elguero, R. W. Taft and F. Anvia, J. Comput. Chem., 1989, 10, 426–433 CrossRef PubMed; (c) C. Cativiela, J. L. García-Laureiro, J. Elguero and E. Elguero, Gazz. Chim. Ital., 1991, 121, 477–481 CAS; (d) J. A. Jiménez, R. M. Claramunt, C. Escolástico and J. Elguero, Struct. Chem., 2000, 11, 77–83 CrossRef.
  24. A. V. Khripun, N. A. Bokach, S. I. Selivanov, M. Haukka, M. D. Revenco and V. Y. Kukushkin, Inorg. Chem. Commun., 2008, 11, 1352–1355 CrossRef CAS PubMed.
  25. (a) M. J. Decker, D. O. K. Fjeldsted, S. R. Stobart and M. J. Zaworoko, J. Chem. Soc., Chem. Commun., 1983, 1525–1527 RSC; (b) G. W. Bushnell, K. R. Dixon, D. T. Eadie and S. R. Stobart, Inorg. Chem., 1981, 20, 1545–1552 CrossRef CAS; (c) R. Giménez, A. Elduque, J. A. López, J. Barberá, E. Cavero, I. Lantero, L. A. Oro and J. L. Serrano, Inorg. Chem., 2006, 45, 10363–10370 CrossRef PubMed.
  26. Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, 215–241 CrossRef CAS.
  27. (a) C. Tejel, M. A. Ciriano, M. Millaruelo, J. A. López, F. J. Lahoz and L. A. Oro, Inorg. Chem., 2003, 42, 4750–4758 CrossRef CAS PubMed; (b) J. P. Rourke, G. Stringer, P. Chow, R. J. Deeth, D. S. Yufit, J. A. K. Howard and T. B. Marder, Organometallics, 2002, 21, 429–437 CrossRef CAS; (c) E. Gutiérrez-Puebla, A. Monge, M. C. Nicasio, P. J. Pérez, M. L. Poveda, L. Rey, C. Ruiz and E. Carmona, Inorg. Chem., 1998, 37, 4538–4546 CrossRef PubMed; (d) M. L. H. Green and D. J. Jones, Adv. Inorg. Radiochem., 1965, 7, 115–183 CrossRef CAS; (e) M. A. Garralda, R. Hernández, L. Ibarlucea, E. Pinilla, M. R. Torres and M. Zarandona, Organometallics, 2007, 26, 1031–1038 CrossRef CAS.
  28. (a) R. El Mail, M. A. Garralda, R. Hernández, L. Ibarlucea, E. Pinilla and M. R. Torres, Organometallics, 2000, 19, 5310–5317 CrossRef CAS; (b) G. Brockaart, R. El Mail, M. A. Garralda, R. Hernández, L. Ibarlucea and J. I. Santos, Inorg. Chim. Acta, 2002, 338, 249–254 CrossRef CAS.
  29. (a) L. B. Han, N. Choi and M. Tanaka, Organometallics, 1996, 15, 3259–3261 CrossRef CAS; (b) A. Christiansen, D. Selent, A. Spannenberg, W. Baumann, R. Franke and A. Börner, Organometallics, 2010, 29, 3139–3145 CrossRef CAS.
  30. G. D. Vaughn, C. E. Strouse and J. A. Gladysz, J. Am. Chem. Soc., 1986, 108, 1462–1472 CrossRef CAS.
  31. (a) B. J. Coe and S. J. Glenwright, Coord. Chem. Rev., 2000, 203, 5–80 CrossRef CAS; (b) B. de Bruin, J. A. Brands, J. J. J. Donners, M. P. J. Donners, R. de Gelder, J. M. M. Smits, A. W. Gal and A. L. Spek, Chem. – Eur. J., 1999, 5, 2921–2936 CrossRef CAS; (c) M. A. Garralda, R. Hernández, L. Ibarlucea, E. Pinilla, M. R. Torres and M. Zarandona, Organometallics, 2007, 26, 5369–5376 CrossRef CAS.
  32. (a) T. Steiner, Angew. Chem., Int. Ed., 2002, 41, 48–76 CrossRef CAS; (b) R. Bieda, M. C. Winter and W. S. Sheldrick, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2007, 63, m2473 CAS; (c) F. P. Pruchnik, P. Jakimowicz, Z. Ciunik, J. Zakrzewska-Czerwinska, A. Opolski, J. Wietrzyk and E. Wojdat, Inorg. Chim. Acta, 2002, 334, 59–66 CrossRef CAS.
  33. J. Chatt and L. M. Venanzi, J. Chem. Soc. A, 1957, 4735–4740 RSC.
  34. S. Laue, L. Greiner, J. Wöltinger and A. Liese, Adv. Synth. Catal., 2001, 343, 711–720 CrossRef CAS.
  35. G. M. Sheldrick, Acta Crystallogr., Sect. A: Fundam. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
  36. A. L. Spek, PLATON-4, A Multipurpose Crystallographic Tool, University of Utrecht, The Netherlands, 1994 Search PubMed.
  37. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.
  38. (a) A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS PubMed; (b) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter, 1988, 37, 785–789 CrossRef CAS; (c) S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys., 1980, 58, 1200–1211 CrossRef CAS PubMed; (d) P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627 CrossRef CAS.
  39. (a) P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270–283 CrossRef CAS PubMed; (b) W. R. Wadt and P. J. Hay, J. Chem. Phys., 1985, 82, 284–298 CrossRef CAS PubMed; (c) P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 299–310 CrossRef CAS PubMed.
  40. (a) J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999–3093 CrossRef CAS PubMed; (b) M. Cossi, V. Barone, R. Cammi and J. Tomasi, Chem. Phys. Lett., 1996, 255, 327–335 CrossRef CAS; (c) J. Tomasi and M. Persico, Chem. Rev., 1994, 94, 2027–2094 CrossRef CAS.
  41. H. Rzepa , “Conformational Analysis”. Imperial College London. Retrieved 11 November 2013.

Footnote

Electronic supplementary information (ESI) available: Cartesian coordinates of optimized DFT structures, images generated from those coordinate files, computationally obtained energy values associated with each structure, and crystallographic data file of complexes 2-Hpz, 3a-Hdmpz and 7-Hpz. CCDC 1062495–1062497. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5dt01705j

This journal is © The Royal Society of Chemistry 2015