Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2020 Apr 24;39(14):2594–2601. doi: 10.1021/acs.organomet.0c00167

Synthesis, Characterization, and Catalytic Reactivity of {CoNO}8 PCP Pincer Complexes

Jan Pecak , Wolfgang Eder , Berthold Stöger , Sara Realista §, Paulo N Martinho , Maria José Calhorda , Wolfgang Linert , Karl Kirchner †,*
PMCID: PMC7388324  PMID: 32742055

Abstract

graphic file with name om0c00167_0015.jpg

The reaction of coordinatively unsaturated Co(II) PCP pincer complexes with nitric oxide leads to the formation of new, air-stable, diamagnetic mono nitrosyl compounds. The synthesis and characterization of five- and four-coordinate Co(III) and Co(I) nitrosyl pincer complexes based on three different ligand scaffolds is described. Passing NO through a solution of [Co(PCPNMe-iPr)Cl], [Co(PCPO-iPr)Cl] or [Co(PCPCH2-iPr)Br] led to the formation of the low-spin complex [Co(PCP-iPr)(NO)X] with a strongly bent NO ligand. Treatment of the latter species with (X = Cl, Br) AgBF4 led to chloride abstraction and formation of cationic square-planar Co(I) complexes of the type [Co(PCP-iPr)(NO)]+ featuring a linear NO group. This reaction could be viewed as a formal two electron reduction of the metal center by the NO radical from Co(III) to Co(I), if NO is counted as NO+. Hence, these systems can be described as {CoNO}8 according to the Enemark–Feltham convention. X-ray structures, spectroscopic and electrochemical data of all nitrosyl complexes are presented. Preliminary studies show that [Co(PCPNMe-iPr)(NO)]+ catalyzes efficiently the reductive hydroboration of nitriles with pinacolborane (HBpin) forming an intermediate {CoNO}8 hydride species.

Introduction

The investigation of pincer ligands and their transition metal complexes has received significant attention during the past decades.1 Their adjustable steric and electronic properties enable the stabilization of various chemical systems and thus allow for the investigation of fundamental reactivity and for the design of innovative catalysts. Since their introduction by Moulton and Shaw in 1976, no pincer ligand has retained as much attention as the PCP scaffold based on the phosphinomethylated benzene.2 Related pincer ligands that feature an aromatic anionic benzene backbone connected to phosphine donors via CH2, O, or NR (R = H, alkyl or aryl) linkers provide for a very attractive class of ligands that are simple to prepare and modify. With respect to first-row transition metals, the chemistry of nickel PCP complexes is already quite comprehensive, but studies of cobalt, iron, manganese and chromium PCP pincer complexes are exceedingly rare. This may be attributed to the failure of many simple metal salts to cleave the C–H bonds of the arene moiety of the pincer ligands and/or the thermodynamic instability of the resulting complexes. Another important ancillary ligand in coordination chemistry is the nitrosyl (NO) ligand.3 Transition metal nitrosyl compounds often play an important role in enzymatic reactions but are relatively unexplored yet in homogeneous catalysis. Indeed, the redox noninnocent character of this ligand is one of its basic and eminent features. It can bind to a metal center in various bonding modes, i.e., the linear and the bent mode, and can be treated as either NO+ (nitrosonium), NO (neutral radical), or NO (nitroxyl). This could enable the stabilization of intermediates in catalytic cycles by supplying or detracting electron density or generating vacant coordination sites. Since different binding modes can lead to different formal oxidation states, Enemark and Feltham proposed a simple notation, {MNO}n (n = metal d + NO-π* electrons) to describe metal NO complexes unambiguously and ease their theoretical comparison.4 Surprisingly, as cobalt is concerned, no PCP pincer nitrosyl complexes featuring a direct cobalt–carbon single bond have been reported in the literature. An overview of existing group 9 nitrosyl complexes being supported by pincer type ligands is depicted in Chart 1.57 Recently our group reported a series of chromium, manganese, and iron nitrosyl complexes supported by aromatic PCP and PNP pincer ligands.8 In this account, we describe a series of well-defined four- and five-coordinate low-spin {CoNO}8 nitrosyl PCP pincer complexes based on three different ligand scaffolds. X-ray structures of representative complexes are presented.

Chart 1. Examples of Group 9 Nitrosyl Pincer Complexes.

Chart 1

Results and Discussion

The starting materials for the present study, [Co(PCPNMe-iPr)Cl] (1a) and [Co(PCPO-iPr)Cl] (1b), were obtained by the reaction of anhydrous CoCl2 and PCPNMe/O-iPr (1a, 1b) in the presence of a strong base as reported previously.9,10 The analogous complex [Co(PCPCH2-iPr)Br] (1c), where the PCP ligand features CH2-linkers, was prepared in analogous fashion by treatment of the ligand PCPCH2-iPr-Br with nBuLi and subsequent transmetalation in 65% isolated yield.11 In analogy to 1a and 1b, complex 1c is a d7 low spin complex with a solution magnetic moment μeff of 2.3(1) μB (Evans method). The solid-state structures of complexes 1b and 1c were determined by X-ray diffraction, and representations of the molecules are presented in Figures 1 and 2.

Figure 1.

Figure 1

Structural view of 1b showing 50% displacement ellipsoids (H atoms omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–C1 1.9131(9), Co1–Cl1 2.2258(4), Co1–P1 2.1879(4), Co1–P2 2.1898(3), P1–Co1–P2 162.90(1), C1–Co1–Cl1 178.88(5).

Figure 2.

Figure 2

Structural view of 1c showing 50% displacement ellipsoids (H atoms and a second independent complex omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–C1 1.955(1), Co1–Br1 2.3764(2), Co1–P1 2.2077(4), Co1–P2 2.2124(4), P1–Co1–P2 169.39(2), C1–Co1–Br1 177.49(4).

Selected metrical parameters are provided in the figure captions. The molecular structures of these compounds show the metal in a typical slightly distorted-square planar conformation with the PCP ligands coordinated to the metal center in a tridentate mode. In both complexes the C1–Co–Cl1 and C1–Co1–Br1 angles deviate slightly from linearity being 178.88(5)° and 177.49(4)°, respectively. The P1–M–P2 angles are 162.90(1) and 169.39(2)°, respectively. The Co1–C1 bond distances are 1.9131(9) and 1.955(1), respectively.

Direct nitrosylation of complexes 1a1c with nitric oxide at ambient pressure afforded the five-coordinate, green colored, diamagnetic compounds 2a2c in 91–98% isolated yield (Scheme 1). All three complexes were fully characterized by a combination of 1H, 13C{1H}, and 31P{1H} NMR spectroscopy, ATR-IR spectroscopy, and HR-MS analysis. In the proton decoupled 31P spectrum the compounds exhibit broad resonances at 139.1, 198.9, and 77.4 ppm, respectively, for the N, O, or CH2 linked systems. Moreover, one single strong absorption band for the characteristic NO stretching mode is detected at 1643, 1650, and 1639 cm–1, respectively, in the IR spectrum for 2a2c. These values are indicative for NO in a bent coordination mode and therefore advocate for a formal interpretation as NO anion and the assignment of the oxidation state + III for the cobalt center. In addition, they underline the different donor strengths (CH2 ≈ NR > O) of the ligands in use expressed via NO π-back-bonding, according to the series of Tolman’s electronic parameters.12

Scheme 1. Synthesis of Complexes 2ac.

Scheme 1

In order to unequivocally establish the ligand arrangement around the metal centers, the solid-state structures of complexes 2a and 2b were determined by X-ray diffraction. Suitable single crystals were grown for this purpose from saturated pentane solutions kept at −20 °C. A view of the molecular structures is depicted in Figures 3 and 4 with selected bond distances and angles reported in captions. Both complexes adopt a slightly distorted square pyramidal geometry (τ5 < 0.1)13 with the metal center surrounded by three meridionally placed donor atoms of the PCP ligand and the nitrosyl group occupying the apical position.

Figure 3.

Figure 3

Structural view of 2a showing 50% displacement ellipsoids (H atoms omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–Cl1 2.2947(12), Co1–C1 1.959(2), Co1–P1 2.227(1), Co1–P2 2.236(1), Co1–N3 1.736(2), N3–O1 1.164(3), P1–Co1–P2 155.88(2), C1–Co1–Cl1 151.25(6), Co1–N3–O1 140.1(2).

Figure 4.

Figure 4

Structural view of 2b showing 50% displacement ellipsoids (H atoms omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–Cl1 2.2823(5), Co1–C1 1.939(2), Co1–P1 2.2298(5), Co1–P2 2.2070(5), Co1–N1 1.726(1), N1–O3 1.176(2), P2–Co1–P1 154.58(2), C1–Co1–Cl1 150.69(5), Co1–N1–O3 140.2(1).

The N–O bond distances are 1.164(3) and 1.176(2) Å for 2a and 2b, respectively. The Co–N–O angles of 140.1(2) and 140.2(1)° are in the range expected for a bent configuration. Gaviglio et al. reported a N–O bond distance of 1.184(4) Å and a Rh–N–O angle of 127.4(3)° for the isoelectronic {RhNO}8 pincer complex [RhIII(PCP-tBu)(NO)Cl].5a

Cyclic voltammetry studies were conducted for compound 2a to gain a deeper insight into the electronic structure and properties. The cyclic voltammogram exhibits a strong redox process at E1/2 = 0.11 V in dichloromethane, and two consecutive reductions at E1/2 = −1.21 and −1.64 V (see Supporting Information). The reduction at −1.21 V is an irreversible process and probably metal based. In order to assign the redox processes involved in solution, spectro-electrochemical infrared (SEC-IR, Figure 5) experiments were carried out, and changes in the N–O stretching frequency were monitored.

Figure 5.

Figure 5

SEC-IR experiments with [Co(PCPNMe-iPr)(NO)Cl] (2a, 5 mM in CH2Cl2) at 1.00 V vs Ag pseudoreference electrode. The absorption at 1654 cm–1, typical for a bent NO alignment, gradually shifts to a new resonance at 1771 cm–1 typical of a linear NO coordination mode (OCP = Open Circuit Potential).

After applying a voltage of 1.00 V, the absorption at 1654 cm–1, typical for a bent NO arrangement, decreases with time, and a new peak appears at 1771 cm–1. The new frequency is characteristic of a linear NO+ stretching vibrational mode, therefore suggesting that oxidation took place at the NO moiety rather than at the Co(III) center. DFT (BP86/TZ2P) calculations (see Supporting Information) indicate that the HOMO has mainly dz2 character with both Co–NO σ* and Co–PCP π* antibonding character (Figure 6 left). The DFT-optimized geometry (S = 1/2) of the oxidized species therefore shows only a slight change of the Co–N–O angle from 143.5 (experimentally 140.1°) to 148.2°, despite the more significantly increased νNO frequency (from 1618 to 1722 cm–1, calculated). The spin density is predominantly localized on the dz2 orbital of the metal center (Figure 6, right) suggesting a d7 low-spin system. According to Hoffmann and co-workers penta-coordinate nitrosyls are in general very likely to bend when the basal ligands are strong σ- or π-donors.14 Since the assignment of an oxidation is not trivial owing to the delocalization of the HOMO, [Co(PCPNMe-iPr)(NO)Cl]+ is best described as {CoNO}7 using the Enemark–Feltham notation.

Figure 6.

Figure 6

HOMO of [Co(PCPNMe-iPr)(NO)Cl] (2a, left) and the calculated spin density in open-shell [Co(PCPNMe-iPr)(NO)Cl]+ obtained after electrochemical oxidation (right).

One electron reduction of 2a, on the other side, leads to a loss of the halide ligand (optimized Co–Cl distance 5.41 Å) and formation of a nearly linear NO moiety (calc. Co–N–O angle 176.8°). The Co–NO π bonding and antibonding orbitals are strongly involved in both processes.

Treatment of 2a and 2b with halogen scavengers such as AgBF4 or TlPF6 in CH2Cl2 afforded the square planar, diamagnetic cationic complexes 3a and 3b in 75 to 88% yield (Scheme 2). In the IR spectrum both compounds exhibit strong bands at 1806 and 1861 cm–1, which is a characteristic feature for a linear configuration of the NO moiety. In the 31P{1H} spectrum the associated resonances at 156.0 and 216.9 ppm are shifted downfield with respect to 2a and 2b.

Scheme 2. Synthesis of Complexes 3a and 3b.

Scheme 2

Single crystals of 3a and 3b were successfully grown by slow diffusion of pentane into a saturated solution in THF and analyzed by X-ray diffraction. A structural view is depicted in Figures 7 and 8 with selected bond distances and angles reported in captions. The molecular structures of both compounds show the metal center in a square planar coordination geometry (τ4 < 0.1)15 with the N–O group located trans to the ipso carbon atom. The N–O bond distances are 1.173(4) and 1.162(2) Å for 3a and 3b, respectively. The Co–N–O angles are 176.6(3) and 177.3(2)° showing a small deviation from linearity probably as a result of strong σ-donation by the benzene ring. The Cipso–Co bond length is also slightly shortened compared to their five-coordinate precursors.

Figure 7.

Figure 7

Structural view of 3a showing 50% displacement ellipsoids (H atoms and counterion omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–C1 1.929(4), Co1–P1 2.2086(11), Co1–P2 2.202(1), Co1–N3 1.631(3), N3–O1 1.173(4), P2–Co1–P1 164.57(4), C1–Co1–N3 176.0(2), Co1–N3–O1 176.6(3).

Figure 8.

Figure 8

Structural view of 3b showing 50% displacement ellipsoids (H atoms and counterion omitted for clarity). Selected bond lengths [Å] and angles [°]: Co1–C1 1.932(2), Co1–P1 2.2042(8), Co1–P2 2.2023(7), Co1–N1 1.635(2), N3–O1 1.162(2), P2–Co1–P1 159.68(3), C1–Co1–N1 175.1(1), Co1–N1–O3 177.3(2).

In this sense, the abstraction of a halide atom leads to formation of a complex with a formal oxidation state of + I and an electronic isomerization between two {CoNO}8 complexes: CoIII–NO ↔ CoI–NO+. It is therefore reasonable to compare these compounds to already literature known Co(I) PCP pincer complexes. Murugesan et al. reported on the synthesis of low-spin complexes [CoI(PCPNMe-iPr)(tBuNC)2] and [CoI(PCPNMe-iPr)(CO)2] with very similar shifts in the 31P{1H} spectrum (164.6 and 170.6 ppm, respectively) as compared to 3a (156.0 ppm).16 An alternative and effective approach to cationic 3a is therefore utilizing a simple and fast ligand exchange reaction starting from [CoI(PCPNMe-iPr)(CO)2] (4a, vide supra) and NOBF4 in dichloromethane (Scheme 3). Complex 3a readily reacts with CO in solution to afford the five-coordinate adduct [Co(PCPNMe-iPr)(CO)(NO)]+ (5) similar to the Rh complexes reported by Milstein and co-workers.5a In the IR spectrum this complex gives rise to strong bands at 2025 (νCO) and 1732 (νNO) cm–1. The latter may suggest a linear NO moiety coordinated in the equatorial position trans to the Cipso carbon atom. Compound 5 is inherently unstable and decomposes to 3a instantaneously when exposed to air or a vacuum.

Scheme 3. Alternative Synthesis of Complex 3a.

Scheme 3

The cyclic voltammogram of cationic 3a exhibits redox processes at 0.68 V and two consecutive reductions at −1.14 V and −1.35 V. The first oxidation wave at 0.68 V can tentatively be assigned to oxidation of the PCP ligand moiety. DFT calculations show that the HOMO of 3a is a delocalized π orbital on the pincer ligand.

Scan rate experiments show that the reduction peak at −1.14 V is an irreversible process, whereas the peak at −1.35 V is associated with a quasi-reversible reduction. DFT calculations reveal that the LUMO and LUMO+1 are both antibonding between Co and NO, and the reduction product has an elongated Co–NO bond. The loss of the NO ligand is thus likely to happen in the first irreversible step, and the second reduction wave might result from a secondary product.

Reaction of 3a with LiCl in MeOH-d4 leads to formation of 2a in quantitative yield. It must be mentioned that putative 3c can be prepared in the same manner as 3a and 3b, but it proved to be very difficult to afford an analytically pure compound (approximately 75% yield, νNO = 1823 cm–1).

Since various cobalt species including PNN and NNN pincer type complexes were shown to be catalytically active for the reductive hydroboration of ketones and nitriles, we began preliminary studies on the potential of the nitrosyl Co(I) complex 3a for this type of transformation.1719 Complex 3a (4 mol %) reacted with both aromatic and aliphatic nitriles and 4,4,5,5-tetramethyl-1,3,2-dioxaborolane (HBpin) in benzene at 25 °C (Table 1). After filtration and workup with HCl in dry diethyl ether all amines were isolated as ammonium salts and characterized by 1H and 13C{1H} NMR spectroscopy. Monitoring of the reaction by 1H NMR spectroscopy revealed the formation of a hydride intermediate giving rise to a triplet resonance at −9.3 ppm (JHP = 57 Hz), which may be involved in the catalytic cycle. This species prevails throughout the experiment and may be structurally related to the pyrrole-based hydride complex described by Tonzetich and co-workers.6

Table 1. Reductive Hydroboration of Nitriles Catalyzed by 3aa.

graphic file with name om0c00167_0013.jpg

graphic file with name om0c00167_0014.jpg

a

Reaction conditions: 0.33 mmol substrate, 0.72 mmol HBpin (2.2 equiv), 3a (4 mol %), 1 mL of benzene.

b

Isolated yields as hydrochloride.

c

Conversion determined by 19F{1H} NMR spectroscopy.

Conclusion

In sum we have prepared a series of cobalt nitrosyl complexes supported by PCP pincer ligands that are connected via NMe, O, or CH2 spacers to an aromatic backbone and fully characterized them by a combination of X-ray, NMR, CV, and FT-IR techniques. Five coordinate Co(III) complexes of the type [Co(PCP)(NO)X] (X = Cl or Br) can easily be prepared by direct nitrosylation of the related 15e Co(II) precursors [Co(PCP)X] (1a1c) using nitric oxide. Treatment of these compounds with halide scavengers (AgBF4 or TlPF6) leads to the formation of square planar cationic Co(I) complexes of the type [Co(PCP)(NO)]+ in good yields. Alternatively, [Co(PCPNMe-iPr)(NO)]+ was also obtained by reacting the literature known dicarbonyl complex [Co(PCPNMe-iPr)(CO)2] with NOBF4. Structural characterization by means of X-ray diffraction supports the IR measurements that show that the NO moiety is strongly bent in the Co(III) species (approximately 140°) and nearly linear for the Co(I) systems. Both classes of compounds are stable low-spin diamagnetic compounds that can be unambiguously described as {CoNO}8 systems according to the Enemark–Feltham convention. Preliminary studies showed that [Co(PCPNMe-iPr)(NO)]+ (3a) efficiently catalyzes the reductive hydroboration of nitriles with HBpin at room temperature, presumably via a hydride intermediate. These observations warrant further investigations of the catalytic activity of cobalt nitrosyl pincer complexes.

Experimental Section

General Information

All manipulations were performed under an inert atmosphere of argon by using Schlenk techniques or in an MBraun inert-gas glovebox. The solvents were purified according to standard procedures.20 The deuterated solvents were purchased from Aldrich and dried over 4 Å molecular sieves. Nitric oxide (NO 2.5) was purchased from MESSER GmbH (Gumpoldskirchen, Austria). The synthesis of 2-bromo-1,3-phenylene)bis(methylene)bis(diisopropylphosphane) (P(C–Br)PCH2-iPr),11 [Co(PCPNMe-iPr)Cl] (1a),9 [Co(PCPO-iPr)Cl] (1b),10 and [Co(PCPNMe-iPr)(CO)2] (4a)16 was carried out according to the literature. 1H, 13C{1H}, and 31P{1H} NMR spectra were recorded on Bruker AVANCE-250 and AVANCE-400 spectrometers. 1H and 13C{1H} NMR spectra were referenced internally to residual protiosolvent and solvent resonances, respectively, and are reported relative to tetramethylsilane (δ = 0 ppm). 31P{1H} NMR spectra were referenced externally to H3PO4 (85%) (δ = 0 ppm). Infrared spectra were recorded in attenuated total reflection (ATR) mode on a PerkinElmer Spectrum Two FT-IR spectrometer.

High resolution-accurate mass spectra were recorded on a hybrid Maxis Qq-aoTOF mass spectrometer (Bruker Daltonics, Bremen, Germany) fitted with an ESI source. Measured accurate mass data of the [M]+ ions for confirming calculated elemental compositions were typically within ±5 ppm accuracy. The mass calibration was done with a commercial mixture of perfluorinated trialkyl-triazines (ES Tuning Mix, Agilent Technologies, Santa Clara, CA, USA).

Cyclic voltammograms were recorded in a three-electrode one compartment electrochemical cell with glassy carbon (3 mm diameter, CHI Instruments) and platinum wire as working and counter electrodes, respectively. The reference electrode used was a saturated Calomel electrode (SCE, −0.241 V vs. NHE); however, all potentials are referenced to the Fc+/Fc couple (0.44 V vs SCE). One mM solutions of the complexes in dichloromethane were used and TBAPF6 (0.2 M) as supporting electrolyte. The working electrode was polished with aluminum oxide, washed thoroughly with Milli-Q water and dried under a nitrogen flux. All experiments were accomplished using an Autolab PGSTAT128N potentiostat (Metrohm GmbH) controlled by NOVA software. SEC-IR experiments were performed using a cell acquired from LabOmak. This cell has two platinum grids as working and counter electrodes and an Ag wire as the pseudoreference electrode. A N2 saturated MeCN solution with 5 mM of complex and 0.2 M of TBAPF6 was injected in the SEC-IR cell. A linear voltammogram was performed using a scan rate of 10 mV s–1 to reach the suitable potential (+1.0 vs Ag pseudoreference) after the first oxidation peak. The IR spectra were obtained using a Nicolet Nexus 6700 FTIR spectrophotometer with a resolution of 4 cm–1 with 16 scans between 2200 and 1500 cm–1. The background spectrum was recorded with the sample at the open circuit potential.

Synthesis

[Co(PCPCH2-iPr)Br] (1c)

To a solution of P(C–Br)PCH2-iPr (100 mg, 0.24 mmol) in THF (10 mL), n-BuLi (0.32 mL, 1.6 M, 0.48 mmol) was slowly added at −78 °C and then stirred for 1 h. After allowing to warm to room temperature a suspension of anhydrous CoBr2 (55 mg, 0.25 mmol) in THF (5 mL) was added in dropwise fashion. The reaction mixture was stirred for 0.5 h and all volatiles were removed under reduced pressure. The remaining brownish solid was extracted into n-pentane and the yellow extract was filtered through a plug of Celite. The solvent was partially removed under a vacuum and the solution was placed in a freezer at −20 °C to obtain a yellow crystalline solid. Yield: 74 mg (65%). μeff = 2.3(1) μB (Evans, CH2Cl2). HR-MS (ESI+, CH3OH) m/z calcd for C20H35CoBrP2 [M]+ 475.0729 found 475.0728.

[Co(PCPNMe-iPr)(NO)Cl] (2a)

A solution of [Co(PCPNMe-iPr)Cl] (1a) (100 mg, 0.21 mmol) in THF (5 mL) was stirred for 15 min under a NO gas atmosphere (1 bar) whereupon the solution turned from red to green. Insoluble materials were removed by filtration through a syringe filter. After removal of the solvent under vacuum, the product was obtained as green solid. Yield: 104 mg (98%). 1H NMR (400 MHz, δ, CD2Cl2) 6.96 (t, J = 7.9 Hz, 1H, ph), 6.02 (d, J = 7.9 Hz, 2H, ph), 2.97 (s, 6H, CH3), 2.82–2.64 (m, 4H, CH), 1.55 (app. q, J = 7.5 Hz, 6H, CH3), 1.24 (app. q, J = 6.7 Hz, 6H, CH3), 0.85 (app. q, J = 7.2 Hz, 6H, CH3), 0.76 (app. q, J = 7.5 Hz, 6H, CH3). 13C{1H} NMR (101 MHz, δ, CD2Cl2) 160.3 (t, J = 14.3 Hz), 126.1 (s), 100.9 (t, J = 5.7 Hz), 32.5 (s), 18.3 (s), 17.4 (t, J = 2.9 Hz), 17.2 (t, J = 3.9 Hz), 16.9 (s). 31P{1H} NMR (162 MHz, δ, CD2Cl2) 139.1. IR (ATR, cm–1) 1643 (νNO). HR-MS (ESI+, CH3OH) m/z calcd for C20H37CoN3OP2 [M–Cl]+ 456.1744 found 456.1739.

[Co(PCPO-iPr)(NO)Cl] (2b)

This complex was prepared analogously to 2a with [Co(PCPO-iPr)Cl] (1b) (100 mg, 0.23 mmol) as starting material. Yield: 104 mg (98%). 1H NMR (400 MHz, δ, CD2Cl2) 6.94 (t, J = 7.9 Hz, 1H, ph), 6.54 (d, J = 7.9 Hz, 2H, ph), 2.70–2.51 (m, 4H, CH), 1.47 (app. q, J = 7.6 Hz, 6H, CH3), 1.18 (app. q, J = 7.0 Hz, 6H, CH3), 1.11 (app q, J = 6.9 Hz, 6H, CH3), 1.01 (q, J = 6.9 Hz, 6H, CH3). 13C{1H} NMR (101 MHz, δ, CD2Cl2) 169.9 (t, J = 8.1 Hz), 127.5 (s), 105.6 (t, J = 5.5 Hz), 30.2 (t, J = 8.3 Hz), 29.3 (t, J = 11.6 Hz), 17.7 (s), 16.4 (s), 16.3 (t, J = 2.7 Hz), 16.1 (s). 31P{1H} NMR (162 MHz, δ, CD2Cl2) 198.9. IR (ATR, cm–1) 1650 (νNO). HR-MS (ESI+, CH3OH) m/z calcd for C18H31CoNO3P2 [M–Cl]+ 430.1111 found 430.1124.

[Co(PCPCH2-iPr)(NO)Br] (2c)

A solution of 1c (30 mg, 0.06 mmol) in 5 mL of pentane reacted with nitric oxide whereupon the color changed from yellow to green. The obtained solution was filtered via syringe filter and the solvent was removed in vacuum. Yield: 29 mg (91%). 1H NMR (400 MHz, δ, CD2Cl2) 7.00 (t, J = 7.4 Hz, 2H, ph), 6.92 (d, J = 8.9 Hz, 1H, ph), 3.39 (t, J = 3.8 Hz, 1H, CH2), 3.35 (t, J = 3.8 Hz, 1H, CH2), 3.21 (t, J = 4.7 Hz, 1H, CH2), 4.17 (t, J = 5.5 Hz, 1H, CH2), 2.60–2.44 (m, 4H, CH), 1.34 (app. q, J = 7.1 Hz, 6H, CH3), 1.06 (app. q, J = 7.0 Hz, 6H, CH3), 1.03–0.92 (m, 12H, CH3). 13C{1H} NMR (101 MHz, δ, CD2Cl2) 149.7 (t, J = 9.9 Hz), 124.5 (s), 122.3 (t, J = 8.1 Hz), 31.3 (t, J = 13.0 Hz), 25.6 (t, J = 9.7 Hz), 24.3 (t, J = 10.5 Hz), 10.1 (s), 18.1 (s), 17.6 (s), 17.2 (s). 31P{1H} NMR (162 MHz, δ, CD2Cl2) 77.4. IR (ATR, cm–1) 1639 (νNO). HR-MS (ESI+, CH3OH) m/z calcd for C20H35CoNOP2 [M–Br]+ 426.1526 found 426.1529.

[Co(PCPNMe-iPr)(NO)]BF4 (3a)

Method A. To a solution of 2a (80 mg, 0.16 mmol) in CH2Cl2 (6 mL), AgBF4 (33 mg, 0.16 mmol) was added and the mixture was stirred for 2 h at room temperature. Insoluble materials were removed by filtration through a syringe filter and the solvent was removed under reduced pressure. The remaining red solid was washed twice with pentane (3 mL) and dried under a vacuum. Yield: 75 mg (85%). Method B. To a solution of [Co(PCPNMe-iPr)(CO)2] (4a) (40 mg, 0.08 mmol) in CH2Cl2 (3 mL), NOBF4 (10 mg, 0.08 mmol) was added and the reaction mixture was stirred for 30 min at room temperature whereupon the color changed from yellow to dark red. All volatiles were removed under reduced pressure; the product was washed with pentane (3 mL) and dried under a vacuum. Yield: 39 mg (88%). 1H NMR (400 MHz, δ, CD2Cl2) 6.97 (t, J = 8.2 Hz, 1H, ph), 5.91 (d, J = 8.1 Hz, 2H, ph), 3.57–3.38 (m, 4H, CH), 3.28 (app. t, J = 3.3 Hz, 6H, CH3) 1.52 (app. q, J = 7.4 Hz, 12H, CH3), 1.45 (app. q, J = 7.6 Hz, 12H, CH3). 13C{1H} NMR (101 MHz, δ, CD2Cl2) 164.0 (s), 135.9 (s), 102.1 (t, J = 6.6 Hz), 33.4 (s) 29.6 (t, J = 13.6 Hz), 18.9 (d, J = 11.7 Hz). 31P{1H} NMR (162 MHz, δ, CD2Cl2) 156.0. IR (ATR, cm–1) 1806 (νNO) HR-MS (ESI+, CH3OH) m/z calcd for C20H37CoN3OP2 [M]+ 456.1744 found 456.1736.

[Co(PCPO-iPr)(NO)]PF6 (3b)

This complex was prepared analogously to 3a (method A) with 2b (50 mg, 0.11 mmol) and TlPF6 (39 mg, 0.11 mmol) as starting materials. Yield: 41 mg (75%). 1H NMR (400 MHz, δ, CD2Cl2) 7.16 (t, J = 8.1 Hz, 1H, ph), 6.58 (d, J = 8.1 Hz, 2H, ph), 3.79–3.26 (m, 4H, CH), 1.58 (app. q, J = 7.6 Hz, 12H, CH3), 1.49 (app. q, J = 8.9 Hz, 12H, CH3). 13C{1H} NMR (101 MHz, δ, CD2Cl2) 170.5 (s), 137.6 (s), 106.8 (t, J = 6.8 Hz), 31.8 (t, J = 12.8 Hz), 18.0 (s), 17.5 (s). 31P{1H} NMR (162 MHz, δ, CD2Cl2) 216.9. IR (ATR, cm–1) 1861 (νNO). HR-MS (ESI+, CH3OH) m/z calcd for C18H31CoNO3P2 [M]+ 430.1111 found 430.1110.

Reaction of 3a with LiCl. Formation of 2a

To a solution of 3a (15 mg, 0.03 mmol) in methanol-d4 (1 mL), LiCl (in excess) was added and the suspension stirred for 10 min. The mixture was filtrated via syringe filter into an NMR tube and spectra were immediately measured revealing the quantitative formation of 2a. 1H, 31P{1H} NMR and ESI-MS spectra were identical to those of 2a.

Reaction of 3a with CO. Formation of [Co(PCPNMe-iPr)(CO)(NO)]+ (5)

CO (2 atm) was injected into the headspace of a solution of 3a (15 mg) in CD2Cl2 whereupon the color changed from magenta to brown. The new compound is inherently unstable and decomposes rapidly when the CO atmosphere is removed. Alternatively, the same compound can be prepared by a solid–gas reaction. IR (ATR, cm–1) 1732 (νNO), 2025 (νCO).

General Procedure for the Hydroboration of Nitriles

The nitrile substrates (0.33 mmol, 1 equiv), nitrosyl complex 3a (4 mol %), and pinacolborane (HBpin, 2.2 equiv) were mixed with 1 mL of benzene and stirred at room temperature for 24 h. The reaction mixture was then filtered through a PTFE syringe filter, diluted with Et2O (4 mL), and a solution of HCl in Et2O was added dropwise until the precipitation of the hydrochloride was complete. The obtained ammonium salts were washed with Et2O, dried under vacuum, and analyzed by 1H NMR spectroscopy.

X-ray Structure Determination

X-ray diffraction data for 1a, 1c, 2a, 2b, 3a, and 3b (CCDC 1953233 (1a), 1959138 (1c), 1953234–1953237 (2a, 3a, 3b)) were collected at T = 100 or 150 K in a dry stream of nitrogen on a Bruker Kappa APEX II diffractometer system using graphite-monochromatized Mo Kα radiation (λ = 0.71073 Å) and fine sliced φ- and ω-scans. Data were reduced to intensity values with SAINT and an absorption correction was applied with the multiscan approach implemented in SADABS.21 The structures were solved by the dual space method implemented in SHELXT22 and refined against F2 with SHELXL.23 Non-hydrogen atoms were refined with anisotropic displacement parameters. The H atoms were placed in calculated positions and thereafter refined as riding on the parent atoms. The Co–N and N–O distances of the minor (3%) orientation of a disordered NO ligand were restrained to 1.74(2) and 1.16(2) Å, respectively. Molecular graphics were generated with the program MERCURY.24

Computational Details

The Amsterdam Density Functional program (ADF)25 was used in the Density Functional Theory calculations.26 Geometries were optimized without symmetry constraints, considering solvent (acetonitrile), with gradient correction, using the Vosko–Wilk–Nusair27 Local Density Approximation of the correlation energy and the Generalized Gradient Approximation with Becke’s exchange28 and Perdew’s29 correlation functionals. Unrestricted calculations were carried out for open shell complexes. The solvent correction was taken into account using the COSMO approach implemented in ADF. Relativistic effects were treated with the ZORA approximation.30 Triple ζ Slater-type orbitals (STO) were used to describe all the valence electrons of H, C, B, and N, augmented with a set of two polarization functions (H, single ζ 2s, 2p, C, N, O, and P single ζ, 3d, 4f, and Co single ζ, 4p, 4f) and with a frozen core (1s) for C, O, N, and P, and for Co (1s 2p). Orbitals and three-dimensional structures were drawn with Chemcraft.31

Acknowledgments

Financial support by the Austrian Science Fund (FWF) is gratefully acknowledged (Project No. P29584-N28). MJC, PNM, and SR acknowledge the financial support of Fundação para a Ciência e a Tecnologia (UIDB/04046/2020 and (UIDP/04046/2020).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.organomet.0c00167.

  • 1H, 13C{1H}, and 31P{1H} NMR spectra of all complexes and organic products (PDF)

  • Optimized Cartesian coordinates for DFT-calculated structures (XYZ)

Accession Codes

CCDC 1953233–1953237 and 1959138 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, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

The authors declare no competing financial interest.

This paper was published ASAP on April 24, 2020, with several incorrect structures. In Scheme 1, compounds 1c and 2c should include Br instead of Cl. In Table 1, the bottom two structures should be benzyl amines. The corrected version was reposted on May 19, 2020.

Supplementary Material

om0c00167_si_001.pdf (4.1MB, pdf)
om0c00167_si_002.xyz (21.3KB, xyz)

References

  1. For reviews on pincer complexes, see:; a Gossage R. A.; van de Kuil L. A.; van Koten G. Diaminoarylnickel(II) “pincer” complexes: mechanistic consideration in the Kharasch addition reaction, controlled polymerization, and dendrimeric transition metal catalysis. Acc. Chem. Res. 1998, 31, 423–431. 10.1021/ar970221i. [DOI] [Google Scholar]; b Albrecht M.; van Koten G. Platinum group organometallics based on “Pincer” complexes: sensors, switches, and catalysts in memory of Prof. Dr. Luigi M. Venanzi and his pioneering work in organometallic chemistry, particularly in PCP pincer chemistry. Angew. Chem., Int. Ed. 2001, 40, 3750–3781. . [DOI] [PubMed] [Google Scholar]; c van der Boom M. E.; Milstein D. Cyclometalated phosphine-based pincer complexes: mechanistic insight in catalysis, coordination, and bond activation. Chem. Rev. 2003, 103, 1759–1792. 10.1021/cr960118r. [DOI] [PubMed] [Google Scholar]; d Singleton J. T. The uses of pincer complexes inorganic synthesis. Tetrahedron 2003, 59, 1837–1857. 10.1016/S0040-4020(02)01511-9. [DOI] [Google Scholar]; e Liang L. C. Metal complexes of chelating diarylamido phosphine ligands. Coord. Chem. Rev. 2006, 250, 1152–1177. 10.1016/j.ccr.2006.01.001. [DOI] [Google Scholar]; f The Chemistry of Pincer Compounds; Morales-Morales D., Jensen C. M.; Eds.; Elsevier: Amsterdam, 2007. [Google Scholar]; g Nishiyama H. Synthesis and use of bisoxazolinyl-phenyl pincers. Chem. Soc. Rev. 2007, 36, 1133–1141. 10.1039/b605991k. [DOI] [PubMed] [Google Scholar]; h Benito-Garagorri D.; Kirchner K. Modularly Designed Transition Metal PNP and PCP Pincer Complexes Based on Aminophosphines: Synthesis and Catalytic Applications. Acc. Chem. Res. 2008, 41, 201–213. 10.1021/ar700129q. [DOI] [PubMed] [Google Scholar]; i Choi J.; MacArthur A. H. R.; Brookhart M.; Goldman A. S. Dehydrogenation and Related Reactions Catalyzed by Iridium Pincer Complexes. Chem. Rev. 2011, 111, 1761–1779. 10.1021/cr1003503. [DOI] [PubMed] [Google Scholar]; j Selander N.; Szabo K. J. Catalysis by Palladium Pincer Complexes. Chem. Rev. 2011, 111, 2048–2076. 10.1021/cr1002112. [DOI] [PubMed] [Google Scholar]; k Bhattacharya P.; Guan H. Synthesis and Catalytic Applications of Iron Pincer Complexes. Comments Inorg. Chem. 2011, 32, 88–112. 10.1080/02603594.2011.618855. [DOI] [Google Scholar]; l Schneider S.; Meiners J.; Askevold B. Cooperative Aliphatic PNP Amido Pincer Ligands - Versatile Building Blocks for Coordination Chemistry and Catalysis. Eur. J. Inorg. Chem. 2012, 2012, 412–429. 10.1002/ejic.201100880. [DOI] [Google Scholar]; m Organo-metallic Pincer Chemistry; van Koten G., Milstein D., Eds.; Springer: Berlin, 2013; Topics in Organometallic Chemistry, Vol. 40. [Google Scholar]; n Szabo K. J.; Wendt O. F.. Pincer and Pincer-Type Complexes: Applications in Organic Synthesis and Catalysis; Wiley-VCH: Weinheim, Germany, 2014. [Google Scholar]; o Asay M.; Morales-Morales D. Non-symmetric pincer ligands: complexes and applications in catalysis. Dalton Trans. 2015, 44, 17432–17447. 10.1039/C5DT02295A. [DOI] [PubMed] [Google Scholar]; p Murugesan S.; Kirchner K. Non-precious metal complexes with an anionic PCP pincer architecture. Dalton Trans. 2016, 45, 416–439. 10.1039/C5DT03778F. [DOI] [PubMed] [Google Scholar]
  2. Moulton C. J.; Shaw B. L. Transition metal-carbon Bonds, Part XLII. Complexes of Nickel, Palladium, Platinum, Rhodium and Iridium with the Tridentate Ligand 2,6-bis[(di-t-butylphosphino)]phenyl. J. Chem. Soc., Dalton Trans. 1976, 1020–1024. 10.1039/DT9760001020. [DOI] [Google Scholar]
  3. For reviews on nitric oxide as ligand:; a McCleverty J. A. Chemistry of Nitric Oxide Relevant to Biology. Chem. Rev. 2004, 104, 403–418. 10.1021/cr020623q. [DOI] [PubMed] [Google Scholar]; b Roncaroli F.; Videla M.; Slep L. D.; Olabe J. A. New features in the redox coordination chemistry of metal nitrosyls {M-NO+; M-NO; M-NO (HNO)}. Coord. Chem. Rev. 2007, 251, 1903–1930. 10.1016/j.ccr.2007.04.012. [DOI] [Google Scholar]; c Hayton T. W.; Legzdins P.; Sharp W. B. Coordination and Organometallic Chemistry of Metal-NO Complexes. Chem. Rev. 2002, 102, 935–992. 10.1021/cr000074t. [DOI] [PubMed] [Google Scholar]; d Ford P. C.; Lorkovic I. M. Mechanistic Aspects of the Reactions of Nitric Oxide with Transition-Metal Complexes. Chem. Rev. 2002, 102, 993–1018. 10.1021/cr0000271. [DOI] [PubMed] [Google Scholar]; e Kaim W.; Schwederski B. Non-innocent ligands in bioinorganic chemistry – An overview. Coord. Chem. Rev. 2010, 254, 1580–1588. 10.1016/j.ccr.2010.01.009. [DOI] [Google Scholar]
  4. Enemark J. H.; Feltham R. D. Principles of structure, bonding, and reactivity for metal nitrosyl complexes. Coord. Chem. Rev. 1974, 13, 339–406. 10.1016/S0010-8545(00)80259-3. [DOI] [Google Scholar]
  5. a Gaviglio C.; Ben-David Y.; Shimon L. J. W.; Doctorovich F.; Milstein D. Synthesis, Structure, and Reactivity of Nitrosyl Pincer-Type Rhodium Complexes. Organometallics 2009, 28, 1917–1926. 10.1021/om8011536. [DOI] [Google Scholar]; b Pellegrino J.; Gaviglio C.; Milstein D.; Doctorovich F. Electron Transfer Behavior of Pincer-Type {RhNO}8 Complexes: Spectroscopic Characterization and Reactivity of Paramagnetic {RhNO}9 Complexes. Organometallics 2013, 32, 6555–6564. 10.1021/om4008746. [DOI] [Google Scholar]
  6. Krishnan V. M.; Arman H. D.; Tonzetich Z. J. Preparation and reactivity of a square-planar PNP cobalt (II)-hydrido complex: isolation of the first {CoNO}8-hydride. Dalton Trans. 2018, 47, 1435–1441. 10.1039/C7DT04339B. [DOI] [PubMed] [Google Scholar]
  7. Camp C.; Naested L. C. E.; Severin K.; Arnold J. N-N bond cleavage in a nitrous oxide-NHC adduct promoted by a PNP pincer cobalt(I) complex. Polyhedron 2016, 103, 157–163. 10.1016/j.poly.2015.09.001. [DOI] [Google Scholar]
  8. a Himmelbauer D.; Mastalir M.; Stoger B.; Veiros L. F.; Pignitter M.; Somoza V.; Kirchner K. Iron PCP Pincer Complexes in Three Oxidation States: Reversible Ligand Protonation To Afford an Fe(0) Complex with Agostic C-H Arene Bond. Inorg. Chem. 2018, 57, 7925–7931. 10.1021/acs.inorgchem.8b01018. [DOI] [PubMed] [Google Scholar]; b Himmelbauer D.; Stöger B.; Veiros L. F.; Kirchner K. Reversible Ligand Protonation of a Mn(I) PCP Pincer Complex to Afford a Complex with an η2-Caryl-H Agostic Bond. Organometallics 2018, 37, 3475–3479. 10.1021/acs.organomet.8b00193. [DOI] [Google Scholar]; c Himmelbauer D.; Stöger B.; Veiros L. F.; Pignitter M.; Kirchner K. Cr(II) and Cr(I) PCP Pincer Complexes: Synthesis, Structure, and Catalytic Reactivity. Organometallics 2019, 38, 4669–4678. 10.1021/acs.organomet.9b00651. [DOI] [Google Scholar]; d Pecak J.; Stöger B.; Mastalir M.; Veiros L. F.; Ferreira L. P.; Pignitter M.; Linert W.; Kirchner K. Five-Coordinate Low-Spin {FeNO}7 PNP Pincer Complexes. Inorg. Chem. 2019, 58, 4641–4646. 10.1021/acs.inorgchem.9b00235. [DOI] [PubMed] [Google Scholar]
  9. a Murugesan S.; Stöger B.; Carvalho M. D.; Ferreira L. P.; Pittenauer E.; Allmaier G.; Veiros L. F.; Kirchner K. Synthesis and Reactivity of Four and Five Coordinate Low-spin Co(II) PCP Pincer Complexes and some Ni(II) Analogs. Organometallics 2014, 33, 6132–6140. 10.1021/om5007769. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Murugesan S.; Stoger B.; Weil M.; Veiros L. F.; Kirchner K. Synthesis, Structure, and Reactivity of Co(II) and Ni(II) PCP Pincer Borohydride Complexes. Organometallics 2015, 34, 1364–1372. 10.1021/acs.organomet.5b00075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Foley B. J.; Palit C. M.; Timpa S. D.; Ozerov O. V. Synthesis of (POCOP)Co(Ph)(X) Pincer Complexes and Observation of Aryl–Aryl Reductive Elimination Involving the Pincer Aryl. Organometallics 2018, 37, 3803–3812. 10.1021/acs.organomet.8b00511. [DOI] [Google Scholar]
  11. Eder W.; Stöger B.; Kirchner K. Synthesis and Characterization of Xylene-based Group Six Metal PCP Pincer Complexes. Monatsh. Chem. 2019, 150, 1235–1240. 10.1007/s00706-019-02422-6. [DOI] [Google Scholar]
  12. Tolman C. A. Steric effects of phosphorous ligands in organometallic chemistry and homogenous catalysis. Chem. Rev. 1977, 77, 313–348. 10.1021/cr60307a002. [DOI] [Google Scholar]
  13. Addison A. W.; Rao T. N.; Reedijk J.; van Rijn J.; Verschoor G. C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen-sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2’-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349–1356. 10.1039/DT9840001349. [DOI] [Google Scholar]
  14. Hoffmann R.; Chen M. M. L.; Elian M.; Rossi A. R.; Mingos D. M. P. Pentacoordinate nitrosyls. Inorg. Chem. 1974, 13, 2666–2675. 10.1021/ic50141a024. [DOI] [Google Scholar]
  15. Yang L.; Powell D. R.; Houser R. P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 955–964. 10.1039/B617136B. [DOI] [PubMed] [Google Scholar]
  16. Murugesan S.; Stöger B.; Pittenauer E.; Allmaier G.; Veiros L. F.; Kirchner K. A Cobalt(I) Pincer Complex with an η2-Caryl-H Agostic Bond - Facile C-H Bond Cleavage through Deprotonation, Radical Abstraction, and Oxidative Addition. Angew. Chem., Int. Ed. 2016, 55, 3045–3048. 10.1002/anie.201510145. [DOI] [PubMed] [Google Scholar]
  17. Ibrahim A. D.; Entsminger S. W.; Fout A. R. Insights into a Chemoselective Cobalt Catalyst for the Hydroboration of Alkenes and Nitriles. ACS Catal. 2017, 7, 3730–3734. 10.1021/acscatal.7b00362. [DOI] [Google Scholar]
  18. Zhang H.; Lu Z. Dual-Stereocontrol Asymmetric Cobalt-Catalyzed Hydroboration of Sterically Hindered Styrenes. ACS Catal. 2016, 6, 6596–6600. 10.1021/acscatal.6b02278. [DOI] [Google Scholar]
  19. Ghosh C.; Kim S.; Mena M. R.; Kim J.; Pal R.; Rock C. L.; Groy T. L.; Baik M.; Trovitch R. J. Efficient Cobalt Catalyst for Ambient-Temperature Nitrile Dihydroboration, the Elucidation of a Chelate-Assisted Borylation Mechanism, and a New Synthetic Route to Amides. J. Am. Chem. Soc. 2019, 141, 15327–15337. 10.1021/jacs.9b07529. [DOI] [PubMed] [Google Scholar]
  20. Perrin D. D.; Armarego W. L. F.. Purification of Laboratory Chemicals, 3rd ed.; Pergamon: New York, 1988. [Google Scholar]
  21. Bruker Computer Programs: APEX2, SAINT, and SADABS; Bruker AXS Inc.: Madison, WI, 2018.
  22. Sheldrick G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, A71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Sheldrick G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, C71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Macrae C. F.; Edgington P. R.; McCabe P.; Pidcock E.; Shields G. P.; Taylor R.; Towler M.; van de Streek J. Mercury: visualization and analysis of crystal structures. J. Appl. Crystallogr. 2006, 39, 453–457. 10.1107/S002188980600731X. [DOI] [Google Scholar]
  25. Theoretical Chemistry, Vrije Universiteit, Amsterdam, T. N. ADF2013 SCM http://www.scm.com (last accessed March 2020).
  26. Parr R. G.; Yang W.. Density-Functional Theory of Atoms and Molecules; Oxford University Press, New York, 1989. [Google Scholar]
  27. Vosko S. H.; Wilk L.; Nusair M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: a critical analysis. Can. J. Phys. 1980, 58, 1200–1211. 10.1139/p80-159. [DOI] [Google Scholar]
  28. Becke A. D. A new inhomogeneity parameter in density-functional theory. J. Chem. Phys. 1998, 109, 2092–2098. 10.1063/1.476722. [DOI] [Google Scholar]
  29. Perdew J. P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B: Condens. Matter Mater. Phys. 1986, 33, 8822–882. 10.1103/PhysRevB.33.8822. [DOI] [PubMed] [Google Scholar]
  30. van Lenthe E.; Ehlers A.; Baerends E. Geometry optimizations in the zero order regular approximation for relativistic effects. J. Chem. Phys. 1999, 110, 8943–8953. 10.1063/1.478813. [DOI] [Google Scholar]
  31. Chemcraft Program. http://www.chemcraftprog.com/index.html (accessed March 2020).

Associated Data

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

Supplementary Materials

om0c00167_si_001.pdf (4.1MB, pdf)
om0c00167_si_002.xyz (21.3KB, xyz)

Articles from Organometallics are provided here courtesy of American Chemical Society

RESOURCES