Skip to main content
ACS Omega logoLink to ACS Omega
. 2024 Apr 16;9(17):19690–19699. doi: 10.1021/acsomega.4c02076

Mechanochemical Synthesis of Chromium(III) Complexes Containing Bidentate PN and Tridentate P-NH-P and P-NH-P′ Ligands

Tomilola J Ajayi 1, Alan J Lough 1, Robert H Morris 1,*
PMCID: PMC11064035  PMID: 38708235

Abstract

graphic file with name ao4c02076_0007.jpg

Chromium(III) complexes bearing bidentate {NH2(CH2)2PPh2: PN, (S,S)-[NH2(CHPh)2PPh2]: P’N} and tridentate [Ph2P(CH2)2N(H)(CH2)2PPh2: P-NH-P, (S,S)-(iPr)2PCH2CH2N(H)CH(Ph)CH(Ph)PPh2: P-NH-P′] ligands have been synthesized using a mechanochemical approach. The complexes {cis-[Cr(PN)Cl2]Cl (1), cis-[Cr(P’N)Cl2]Cl (2), mer-Cr(P-NH-P)Cl3 (3), and mer-Cr(P-NH-P′)Cl3 (4)} were obtained in high yield (95–97%) via the grinding of the respective ligands andthe solid Cr(III) ion precursor [CrCl3(THF)3] with the aid of a pestle and mortar, followed by recrystallization in acetonitrile. The isolated complexes are high spin. A single-crystal X-ray diffraction study of 2 revealed a cationic chromium complex with two P’N ligands in a cis configuration with P′ trans to P′ with chloride as the counteranion. The X-ray study of 4 shows a neutral Cr(III) complex with the P-NH-P′ ligand in a mer configuration. The difference in molecular structures and bulkiness of the ligands influence the electronic, magnetic, and electrochemical properties of the complexes as exhibited by the bathochromic shifts in the electronic absorption peaks of the complexes and the relative increase in the magnetic moment of 3 (4.19 μβ) and 4 (4.15 μβ) above the spin only value (3.88 μβ) for a d3 electronic configuration. Complexes 14 were found to be inactive in the hydrogenation of an aldimine [(E)-1-(4-fluorophenyl)-N-phenylmethanimine] under a variety of activating conditions. The addition of magnesium and trimethylsilyl chloride in THF did cause hydrogenation at room temperature, but this occurred even in the absence of the chromium complex. The hydrogen in the amine product came from the THF solvent in this novel reaction, as determined by deuterium incorporation into the product when deuterated THF was used.

Introduction

The industrial demand for cheaper and less toxic alternative metals in the design of catalysts with higher efficiency and selectivity for essential fine chemicals in the flavor, perfumery, agricultural, and pharmaceutical industries is on the increase.17 This has led to tremendous interest in the chemistry of first transition series elements which are more abundant and cheaper than the second and third transition series elements commonly given higher priority in catalytic design.2,3,79 Among the first transition series elements, chromium has shown great potential in catalysis, especially when coordinated to aminophosphine (PN) and aminodiphosphine (PNP) ligands for the production of hex-1-ene and octa-1ene via trimerization and tetramerization of ethene.1015 The combination of hard (N) and soft (P) donors of these ligands is suitable to form stable complexes with Cr(III). The presence of hydrogen on the amino nitrogen, the varying of the substituents on the phosphine(s), and the introduction of asymmetric carbon provide endless possibilities in the synthesis of such metal complexes.1619 Complexes of chromium in several oxidation states have been synthesized including those of Cr(0) with a bidentate PPh2CH2CH2NH2 ligand,20 Cr(III) with bidentate Ph2PNR’PPh2,21 2,6-R1-4-R2-C6H2–N=CH–C6H4–2-PPh213 or Ph2PNMeCH2CH2C5H4N22 ligands, and Cr(II) with tridentate [tBu2PCHC5H3NCHPtBu2]23 ligands, and their structures and catalytic properties were investigated. However, there is not a report on homochiral tridentate P-NH-P′ Cr complex with asymmetric carbon groups. Similarly, there are few reports of PN Cr complexes.13,20 It is therefore not surprising that applications of such Cr complexes have not been given much consideration. Unlike Cr, chiral PN and unsymmetrical P-NH-P′ complexes of Ru,24 Fe,2531 Mn,32,33 and Ir34 are known, and they have been explored in diverse applications.

The general routes adopted in the synthesis of these complexes often involve the use of solvents, which are often toxic and expensive. However, environmental safety concerns have given credence to the concept of green synthesis that minimizes or deliberately avoids the use of solvents. Mechanochemical techniques (dry grinding and liquid assisted grinding) are green routes to synthesize complexes without the use of solvent or with a few drops of solvent especially when the byproducts can be easily removed. This usually avoids the multiple purification and drying processes and also eliminates the generation of waste solvents.20 These methods have been previously highlighted.35 No reports of such a synthesis route have been adopted in the synthesis of metal complexes of chiral PN, P-NH-P, and P-NH-P′ ligands.

There are several protocols for the synthesis of chiral PN. However, the protocol developed by Guo et al.(36) provides a less complicated route with higher yield and purity. The unsymmetrical P-NH-P′ ligand can be derived from the chiral PN ligand via a condensation reaction with α-disubstituted phosphine acetaldehyde, as previously described by our group.37 The tridentate P-NH-P′ pincer ligands are known to adopt different coordination modes around metal centers. In an octahedral coordinate geometry, the P-NH-P′ ligand can adopt a mer or fac coordination mode depending on the bulkiness of the substituents on the phosphines. The mer mode is thermodynamically more stable and more prevalent with bulky substituents.30,32,38,39 Few examples of chromium complexes are known to adopt fac mode on the coordination of Cr(0) to tridentate PPh2CH2CH2N(C2H5)CH2CH2PPh240 and Cr(II) to PPh2CH2Si(CH3)2NCH2Si(CH3)2PPh241,42 ligand. There are several examples of chromium complexes that adopt mer mode with tridentate P-NH-P ethylene spacer groups,14,43 PNP with phenylene backbone,44,45 PNP disubstituted pyrrole backbone,46 PNP disubstituted pyridine backbone,4752 PNP methylene-dimethylsilane backbone,53,54 PNP disubstituted piperidine backbone,55 PNP disubstituted triazine backbone,56 and PNP substituted carbazole backbone.57

In the past, we have studied different Fe and Mn complexes of achiral and chiral PNNP5865 and unsymmetrical P-NH-P’33,38,39,66,67 ligands, which have been demonstrated as efficient catalysts in asymmetric transfer hydrogenation and asymmetric pressure hydrogenation for the synthesis of chiral alcohols with high demand in the perfumery and pharmaceutical industries (Figure 1). However, a dearth of reports on the Cr complexes of these ligands prompted us to investigate the possible interactions and coordination modes that may exist between the Cr(III) ion and these ligands. Moreover, we also explore the mechanochemical technique as a green route for the synthesis of the complexes. Herein, we describe the synthesis, characterization, electrochemical properties, and crystal structures of Cr(III) complexes of chiral PN and unsymmetrical P-NH-P′ ligands.

Figure 1.

Figure 1

Representative of previously reported PNNP, P-NH-P′, and (PN)2 complexes by our group and the new P-NH-P′ and PN chromium complexes in this work (see refs (17, 33, 38 and 39) for the structures shown).

Results and Discussion

The preparations of the ligands [2-(diphenylphosphino)ethylamine (PN), (1S,2S)-2-(diphenylphosphino)-1,2-diphenylethylamine (P’N), Ph2P(CH2)2N(H)(CH2)2PPh2 (P-NH-P), and (S,S)-(iPr)2PCH2CH2N(H)CH(Ph)CH(Ph)PPh2 (P-NH-P′)] have been previously reported by our group38,39 and others,14,43 and the details are provided in the Supporting Information. These pincer ligands were coordinated to Cr(III) using CrCl3(THF)3 as the Cr(III) precursor in an equimolar stoichiometric ratio via a mechanochemical synthesis. The synthesized complexes are listed in Figure 1. All attempts to synthesize CrLCl3(THF) (L: PN or P’N) by varying the stoichiometric ratios resulted in [CrL2Cl2]+ (1 and 2). This reflects the strong affinity of the ligands for the Cr(III) ion. The complexes were dissolved in acetonitrile and left to crystallize at room temperature. Single crystals suitable for X-ray crystallographic study were obtained for complexes 2 and 4. Complex 3 has been previously synthesized in THF, and the crystal structure was reported by McGuinness et al.(43) However, the use of a solvent was avoided in the method of synthesis adopted in this study. The similarity of the obtained product to the previously reported complex43 shows that the same mechanism of interaction occurs between the metal salt and the ligands, with or without the use of solvent. This shows the viability of the adopted mechanochemical synthesis.35,68,69 The report of McGuinness et al.(43) entails the crystal structure, mass spectra, and elemental analyses of complex 3. We provide more detailed characterization and electrochemical properties of complex 3 in this study for comparative analysis with the newly synthesized complexes 1, 2, and 4.

Crystal Structure Analysis

The single-crystal X-ray diffraction study reveals that complex 2 crystallized in the orthorhombic space group P212121 (Table S1). Two P’N ligands are coordinated to the Cr(III) via the N (amino) and P (phosphino) binding sites. The P binding sites occupy the axial position in a trans-conformation, while the two N sites are in the equatorial region in a cis-conformation (Figure 2). Similarly, two chlorides also coordinate to Cr(III) in a cis-conformation to complete the octahedral geometry. The coordinated P’N ligands form two five-membered rings with the Cr(III) center. The bond lengths around the Cr(III) center follow the covalent radii [P (1.11 Å), Cl (0.99 Å), and N (0.71 Å)] of the coordinated atoms in the order Cr–P > Cr–Cl > Cr–N. The bond angles 174.88° [N(2)–Cr–Cl(1)], 175.57° [N(1)–Cr–Cl(2)], and 159.40° [P(1)–Cr–P(2)] are less than 180 expected for an ideal octahedral complex, which reflect the distortion in complex 2 along its linear planes and also account for the twist of the P’N ligands as they coordinate to the Cr(III) center. Similar angular deviations from 90° are observed in the bond angles 79.52° [(N(1)–Cr–P(1)], 87.76° [N(2)–Cr–P(1)], and 97.34° [Cl(1)–Cr–P(1)] of the coordinated groups between the axial and equatorial planes, which also reflects the observed distortion in the geometry of the complex.

Figure 2.

Figure 2

X-ray single-crystal structure of complex 2. The hydrogen atoms of the phenyl groups have been omitted for the sake of clarity. The atomic symbols and the corresponding numbering of the elements in the structures are indicated for carbon (C), nitrogen (N), phosphorus (P), and chromium (Cr). Selected bond lengths (Å) and angles (deg) of complex 2: Cr(1)–N(1) 2.086(3), Cr(1)–N(2) 2.104(3), Cr(1)–Cl(1) 2.3012(9), Cr(1)–Cl(2) 2.3053(9), Cr(1)–P(1) 2.4815(13), Cr(1)–P(2) 2.5030(13), N(1)–Cr(1)–N(2) 95.02(10), N(1)–Cr(1)–Cl(1) 86.32(8), N(2)–Cr(1)–Cl(1), 174.88(10), N(1)–Cr(1)–Cl(2) 175.57(10), N(2)–Cr(1)–Cl(2) 84.54(8), Cl(1)–Cr(1)–Cl(2) 94.50(3), P(1)–Cr(1)–P(2) 159.40(3), N(1)–Cr(1)–P(1) 79.52(10), N(2)–Cr(1)–P(1) 87.76(10), Cl(1)–Cr(1)–P(1) 97.34(5), Cl(2)–Cr(1)–P(1) 96.05(5), N(1)–Cr(1)–P(2) 85.96(10), N(2)–Cr(1)–P(2) 78.98(10), Cl(1)–Cr(1)–P(2) 96.21(5), and Cl(2)–Cr(1)–P(2) 98.27(5).

Unlike complex 2, complex 4 with a tridentate P-NH-P′ ligand crystallized in the triclinic space group P21 (Table S1). The ligand coordinates via the P and N moieties of the phosphine and amino moieties in a meridional mode. The coordination of three Cl ions to the Cr(III) center completed the octahedral geometry. The Cl ions are dispersed along the Meridional plane transversely positioned to the P-NH-P′ ligand (Figure 3). This arrangement minimizes the geometric constraints and enhances the thermodynamic stability of the complex in comparison to the facial mode.70,71 The P-NH-P′ ligand forms two fused five-membered bicyclic rings with the Cr(III) center, with the C(2) and C(3) occupying an endo position to the N of the amino moiety.32 The bond lengths Cr–Cl, Cr–N, and Cr–P are longer than those of complex 2 with the P’N ligand. This reflects the steric hindrance imposed by the bulkiness of the P-NH-P′,33,38 which induces distortion in the octahedral geometry with elongated axial bond lengths 2.5252(9) Å [Cr–P(1)] and 2.4734(9) Å [Cr–P(2)] in comparison with the bond lengths along the equatorial plane. The bond angles also deviate from 90 and 180°, which also indicates distortion in the octahedral geometry of complex 4. This can be attributed to the steric influence of the P-NH-P′ ligand due to its bulkiness. Complex 4 has a strong structural similarity with previously reported complex 3(43) as indicated by the closeness of the bond angles: P(1)–Cr–P(2) (3: 163.10°, 4: 163.14°), N–Cr–Cl(2) (3: 176.74°, 4: 177.00°), N–Cr–P(1) (3: 81.08°, 4: 81.85°), and N–Cr–P(2) (3: 82.07°, 4: 82.18°).

Figure 3.

Figure 3

X-ray single-crystal structure of complex 4. The hydrogen atoms of the phenyl and isopropyl groups have been omitted for the sake of clarity. The atomic symbols and the corresponding numbering of the elements in the structures are indicated for carbon (C), nitrogen (N), phosphorus (P), and chromium (Cr). Selected bond lengths (Å) and angles (deg) of complex 4: Cr(1A)–N(1A) 2.136(2), Cr(1A)–Cl(2A) 2.3032(8), Cr(1A)–Cl(3A) 2.3205(8), Cr(1A)–Cl(1A) 2.3345(8), Cr(1A)–P(2A) 2.4734(9), Cr(1A)–P(1A) 2.5252(9), N(1A)–Cr(1A)–Cl(2A) 177.00(7), N(1A)–Cr(1A)–Cl(3A) 86.05(8), Cl(2A)–Cr(1A)–Cl(3A) 91.47(3), N(1A)–Cr(1A)–Cl(1A) 84.80(8), Cl(2A)–Cr(1A)–Cl(1A) 97.72(3), Cl(3A)–Cr(1A)–Cl(1A) 170.67(3), N(1A)–Cr(1A)–P(2A) 82.18(7), Cl(2A)–Cr(1A)–P(2A) 99.49(3), Cl(3A)–Cr(1A)–P(2A) 89.22(3), Cl(1A)–Cr(1A)–P(2A) 87.70(3), N(1A)–Cr(1A)–P(1A) 81.85(7), Cl(2A)–Cr(1A)–P(1A) 96.70(3), Cl(3A)–Cr(1A)–P(1A) 95.06(3), Cl(1A)–Cr(1A)–P(1A) 85.49(3), and P(2A)–Cr(1A)–P(1A) 163.14(3).

Electronic Absorption Spectra

The spectra of the complexes (14) in DMSO display an intense prominent absorption in the UV region (304–354 nm) (Figure 4A). This corresponds to intra-ligand transition which is strongly influenced by the nature of the ligand.72 A considerable redshift and broadening were observed in the peak absorption wavelength on going from the cationic complexes 1 and 2 to the neutral complexes 3 and 4. The order of the spin-allowed ligand field transitions characteristic of a Cr(III) complex with d3 electronic configurations is 4A2g4T2g (v1) < 4A2g4T1g (F) (v2) < 4A2g4T1g (P) (v3).73,74 We use these octahedral field symmetry labels in our work for comparison with assignments for literature complexes, even though our complexes are distorted from this ideal geometry. The low energy transition B in the region of 578–586 nm, the 4A2g4T2g (v1), has the energy of the ligand field splitting 10Dq (Table 1). Complexes 1 and 2 are thought to have similar cis geometries. However, the PN ligands in complex 1 have less hindered, more flexible backbones than the P’N ligands in complex 2, and this causes a slightly larger splitting and stabilization of complex 1 (Dq = 1712 cm–1) compared to complex 2 (Dq = 1701 cm–1). Similarly, the more flexible PNP ligand causes the larger splitting in complex 3 (Dq = 1730 cm–1) compared to complex 4 (Dq = 1707 cm–1) with the more rigid PNP’ ligand, even though the latter has a potentially more basic PiPr2 donor group. These Dq values are intermediate between those observed for [CrCl6]3– (1360) and [Cr(en)3]3+ (2188)75,76 as might be expected. The absorption peak B (Figure 4) in the region 439–468 nm can be attributed to 4A2g4T1g (F) v2 transition.77,78 The spacing between the v1 and v2 transitions of the neutral complexes (4485 cm–1 for 3 and 4303 cm–1 for 4) is smaller than that for the cationic complexes (5656 cm–1 for 1 and 5926 cm–1 for 2). This spacing is even larger for [Cr(OH2)6]3+ and [Cr(NH3)6]3+ (>6000 cm–1) due to an increase d electron repulsion as expressed by the Racah B parameter.75 The transition of the highest energy [4A2g4T1g (P)] is obscured by the more intense charge transfer transitions and would be expected in this case at approximately 40,000 cm–1 (12Dq + 15B + 9349) assuming the Racah parameter B is 680 cm–1 and the interaction energy between the two 4T1g states is 9349 cm–1 as determined from the position of v2.79 The molar extinction coefficients (ε) observed for v1 (60–301 M–1 cm–1) and v2 (55–214 M–1 cm–1) (Table 1) are within the expected range for spin-allowed transitions.74

Figure 4.

Figure 4

Electronic absorption spectra of complexes 14 in DMSO 1 mg/mL; [1.6 (1), 1.1 (2), 1.7 (3), and 1.5 (4) mM].

Table 1. Electronic Absorption Peak Wavelength (nm), the Corresponding Wavenumber (cm–1), Molar Extinction Coefficient (ε; M–1 cm–1), and Crystal Field Splitting Energy [Dq (cm–1)] of 14.

  π–π* ε (π–π*) 4A2g → 4T1g (v2) ε (v2) 4A2g → 4T2g (v1) ε (v1) Dq
1 304 (32,895) 1415 439 (22,779) 60 584 (17,123) 55 1712
2 311 (32,154) 2022 436 (22,936) 90 588 (17,007) 111 1701
3 321 (31,153) 1316 459 (21,786) 72 578 (17,301) 78 1730
4 331 (30,211) 1496 468 (21,368) 301 586 (17,065) 214 1707
  354 (28,249) sh 1362          

FTIR Vibrational Spectroscopy

The vibrational bands for v(N–H) stretching in the region 3171–3298 cm–1 of the spectra of films of the complexes (14) indicate that the amino protons were retained after the coordination of the ligands to the Cr(III) center.14,74 Hence, the amino ligands remained neutral, and the Cr(III) oxidation state was preserved. The vibrational bands in the region 400–500 cm–1 are characteristics of v(Cr–Cl), which accounts for the chlorides in the coordination sphere.74,77 It is intriguing to note that the influence of the difference in symmetric nature of the ligands becomes obvious as the primary amino protons in complex 2 display three vibrational bands (3201, 3242, and 3298 cm–1) in contrast to complex 1, where a single band (3173 cm–1) is indicated for its primary amino protons. This difference is, however, not observed in complexes 3 and 4 with tridentate ligands, which contain secondary amino protons with vibrational bands at 3171 and 3286 cm–1, respectively.

Magnetic Susceptibility and Electron Paramagnetic Resonance Spectroscopy

The effective magnetic moments (μeff) of complexes 14 were determined by the Evans’ method.80 The μeff for the complexes fall within the expected range of spin only magnetic moment (μs.o = 3.88 μβ) for a d3 electronic configuration.80,81 The X-band electron paramagnetic resonance (EPR) spectra of complexes 14 as powders (see the Supporting Information) were obtained at room temperature and display the broad band characteristics of Cr(III) species with a d3 electronic configuration.82 The broad peaks hide any hyperfine couplings to the 31P and 53Cr nuclei.21,83

Electrochemical Study

The cyclic voltammogram (CV) of complex 1 (Figure 5) shows a complex series of redox events including CrIII/II and CrII/I reduction processes with cathodic peak potentials (Epc) at −0.44 and −1.42 V, and anodic peak potentials (Epa) at −0.32 and −0.77 V. Reduction of substitution inert CrIII produces labile CrII, resulting in irreversible electrochemical behavior. The Epa at −0.12 V indicates an irreversible oxidation of Cl with a surge in current as the product of the oxidation (Cl2 gas) desorbed from the surface of the electrode.84,85 The redox activity associated with Cl is equally observed by measuring the CV of tetrabutyl ammonium chloride (TBA+ Cl) in DMF and MeCN, where a similar increase in current is indicated in the CVs at −0.18 and −0.09 V in DMF and MeCN, respectively (Figure S6). The high potential suggests the substitution of dissociated Cl with a solvent molecule (DMF) from reduced chromium species (eqs 1, 2).

graphic file with name ao4c02076_m001.jpg 1
graphic file with name ao4c02076_m002.jpg 2

Figure 5.

Figure 5

CVs of complexes 14 in 0.1 M [nBu4N][BF4]/DMF.

Similarly, an irreversible reduction at Epc of −1.96 V can be attributed to the substitution of the PN ligand with DMF.86,87 The CV of complex 2 also shows a two-step electron transfer processes with Epc at −0.41 and −1.33 V and corresponding Epa at −0.32 and −0.77 V. The rapid increase in current in the region where E > −0.15 V again suggests the contribution of the irreversible Cl ion oxidation/substitution with DMF. The CVs of the complexes with tridentate ligands (3 and 4) also reveal a two-step electron transfer characteristic of Cr ion. The influence of the Cl ion oxidation significantly increased the current in the region where E > −0.10 V (eq 3), and the involvement/substitution of the PNP and PNP’ ligands can be attributed to the Epc at −2.10 and −2.11 V in the CV of complex 3 and 4, respectively.

graphic file with name ao4c02076_m003.jpg 3

The likely substitution of the Cl and the pincer ligands with solvent molecules in the electrochemical process is further investigated by replacing DMF with MeCN.73 MeCN significantly shifts the redox potentials in the electrochemical process, especially for the potentials associated with the substitution of the Cl and pincer ligands (Figure S7).

Lack of Catalytic Activity of Complexes 1–4 in Hydrogenation Reactions

The catalytic activity of the chromium complexes (14) was investigated in the hydrogenation of the aldimine [(E)-1-(4-fluorophenyl)-N-phenylmethanimine] to the amine [N-(4-fluorobenzyl)aniline] using molecular H2 under 25 bar in THF at 60 °C with a 5 mol % catalytic loading. The reaction was monitored by 19F NMR spectroscopy. Complexes 14 showed no observable activity. Further catalytic investigations were carried out using complexes 14 (1 mol %) in the asymmetric transfer hydrogenation of diphenylphosphinoyl imines to their respective amines analogues using KOtBu (8 mol %) as a base and 2-PrOH as a solvent and as a reducing agent at 30 °C.62,88,89 After 24 h, no conversion was noted using 31P NMR spectroscopy to monitor the reaction. We envisaged the possible activation of the complexes by the generation of their corresponding hydrides using LiHBEt3 in THF at −30 to 28 °C to induce their catalytic activity in transfer hydrogenation.38,90 The formation of hydrides was not successful. We also attempted to aid the hydride formation of the complexes by first generating the triflate analogues of the complexes 14 but the complexes showed no reaction in the presence of AgOTf in toluene. On the basis of these negative results, we did not proceed further on investigating the complexes as catalysts in transfer hydrogenation.

Following a recent report on the chromium complexes of cyclic (alkyl)(amino) carbene (CAAC)-catalyzed hydrogenation of alkynes for the selective generation of E- and Z- olefins with high yield in the presence of Mg and trimethylsilyl chloride (TMSCl),91 we further investigated the possibility that our complexes 14 are catalysts for the conversion of imine [(E)-1-(4-fluorophenyl)-N-phenylmethanimine] to amine [N-(4-fluorobenzyl)aniline] using H2. However, our preliminary investigation showed that the hydrogenation to the amine proceeded in the presence of Mg and TMSCl with or without our complexes as catalysts (Table 2, entries 1–5). We wondered whether the hydrogenation proceeded via the in situ generation of a Grignard-like reagent from Mg and TMSCl. However, when only the Grignard reagent (PhCH2MgBr) (entry 6) or activated Mg (entry 7) was used under H2 in the absence of a chromium complex, lower conversion to the amine was observed. Furthermore, when we excluded H2 (entry 8) and conducted the reaction under Ar at room temperature (entry 9), the hydrogenation still proceeded. This implicates the solvent (THF) as being the hydrogen source. We confirmed this hypothesis by conducting two separate experiments. In the first, the dideuterated amine PhNDCDHC6H4F was generated by the reaction of the imine with NaBD4 in CD3OD in order to obtain its 19F NMR and mass spectrum. Then, deuterated THF-D8 (entry 10) was used as the solvent in the Mg/TMSCl reaction under the same conditions as in entry 9 but with a longer duration (48 h). The mass spectrum of the isolated crude amine product was consistent with partial conversion to the deuterated amine PhNDCDHC6H4F. However, the conversion was less than that of entry 9, presumably because of the significant kinetic isotope effect of breaking the carbon-deuterium bond of the THF-D8 [BDE(C–H) 92 kcal/mol92. Thus, we suggest that the hydrogenation proceeds by magnesium reduction of the imine or TMS-iminium chloride with hydrogen atom abstraction from the THF. This side reaction was not reported for the Cr(CAAC) complexes that showed catalytic activity and selectivity in the hydrogenation of alkynes in the presence of Mg/TMSCl in THF.91

Table 2. Catalytic Study of Hydrogenation of an Imine (0.1 M) to an Amine.

graphic file with name ao4c02076_0006.jpg

entry [Cr] Mg (equiv) TMSCl (equiv) yield (%)
1 1 2 0.5 98
2 2 2 0.5 100
3 3 2 0.5 96
4 4 2 0.5 97
5   2 0.5 99
6a       32
7b       33
8c   2 0.5 99
9c,d   2 0.5 90
10c,d,e   2 0.5 20
a

Grignard reagent (PhCH2MgCl).

b

Activated Mg.

c

Under Ar.

d

22 °C.

e

THF-D8 was used as the solvent.

Conclusions

We have demonstrated the preparation of new Cr(III) complexes in high-yield (95–97%) with a solventless mechanochemical approach and studied the electronic, magnetic, and electrochemical properties of these complexes. The structural analysis revealed that the bidentate P’N ligands form a cis complex with trans Cr–P bonds. The P-NH-P′ ligand coordinates in a mer arrangement about the Cr(III) center with three mer Cl ions. In the case of complex 1, no suitable crystals were obtained for the X-ray diffraction study. However, we envisage that it will adopt a structural configuration similar to that of complex 2 based on similarity in properties just as complex 4 shows strong structural resemblance to the previously reported complex 3. The structural differences in the ligands induce variations in the electronic, magnetic, and electrochemical properties of the complexes, which influence their stability. The only conditions discovered for the hydrogenation of an aldimine did not involve the chromium complexes but instead involved the novel reduction of the imine by a Mg/TMSCl mixture under Ar at room temperature with hydrogen abstraction from the THF solvent.

Experimental Section

General Experimental Details

All the procedures and modifications reported in this work were conducted under a N2 of Ar atmosphere using standard Schlenk line and glovebox facilities, unless otherwise stated. The reagents were used as received from Sigma-Aldrich, Alfa Aesar, and ACROS Organics. All solvents were predegassed and dried based on standard procedures before the commencement of each experiment. The 1H (500 MHz) NMR spectra were recorded on an Agilent DD2 500 MHz. The FTIR spectra were acquired on a Bruker Alpha spectrometer equipped with an ATR platinum-diamond attachment. X-ray crystallographic data for 2 and 4 were collected on a Bruker Kappa APEX-DUO diffractometer equipped with a PHOTON II CMOS detector and were measured using a combination of ϕ scans and ω scans. The data were processed using APEX3 and SAINT. Mass spectra were obtained with a JEOL AccuTOF Plus 4G equipped with a direct analysis in real time ion source. The absorption spectra were measured by using DMSO solutions of the complexes (14). For the Evans’ method,80 the NMR frequency was 500 MHz, and the reference solvent used for 1 was C4H8O/C4D8O, while CH3CN/CD3CN was used for 24. The X-band EPR spectra were obtained by using a Bruker CW X-band ECS-EMXplus EPR spectrometer at 9.5 GHz. The electrochemical study was conducted using a three-electrode setup with a platinum wire as a working electrode, a tungsten auxiliary electrode, and a Ag/AgCl reference electrode. The electrolyte was a DMF solution with 0.1 M [nBu4N][BF4]. All potentials were referenced to the ferrocenium/ferrocene (Fc+/Fc) couple (0.72 V versus the standard hydrogen electrode, SHE).67 All characterizations were conducted using powder products of the complexes, except for the X-ray diffraction studies using crystals obtained from MeCN.

Mechanochemical Synthesis

[Cr(PN)2Cl2]Cl (1)

Purple CrCl3(THF)3 (0.24 g, 0.65 mmol) and white PN (0.30 g, 1.3 mmol) were ground with a pestle and mortar in a glovebox for 5 min to a dark blue powder and left overnight at room temperature. A solid product was obtained. Yield: 97% (0.39 g, 0.63 mmol). HRMS: calcd for C28H35Cl3CrN2P2, 619.90 ([M]+); found, (ESI+): 619.06. EA: % Anal. Calcd (found) for C28H35Cl3CrN2P2: 619.90 g/mol: C, 54.25 (54.19); H, 5.69 (5.62); N, 4.52 (4.49). FT-IR (ATR, v/cm–1): 3173 (m, NH2), 3053, (m, CH), 2949 (m, CH), 2924 (m, CH), 2870 (m, CH), 459 (s, Cr–Cl), 437 (s, Cr–Cl). μeff: 3.31 μβ.

[Cr(P’N)2Cl2]Cl (2)

CrCl3(THF)3 (0.24 g, 0.65 mmol) and white P’N (0.50 g, 1.3 mmol) were ground with a pestle and mortar in a glovebox for 5 min to a dark blue powder and left overnight at room temperature. A solid product was obtained, which was dissolved in acetonitrile and left to recrystallize at room temperature for use in the X-ray diffraction study. Yield: 95% (0.57 g, 0.62 mmol). HRMS: calcd for C52H51Cl3CrN2P2, 922.20 ([M]+); found, (ESI+): 922.01. EA: % Anal. Calcd (found) for C52H51Cl3CrN2P2: 922.20 g/mol: C, 67.57 (67.54); H, 5.56 (5.57); N, 3.03 (3.07). FT-IR (ATR, v/cm–1): 3298 (w, NH2), 3242 (w, NH2), 3201 (w, NH2), 3062, (w, CH), 2980 (w, CH), 459 (m, Cr–Cl), 432 (m, Cr–Cl). μeff: 3.23 μβ.

[Cr(P-NH-P)Cl3] (3)

CrCl3(THF)3 (0.48 g, 1.3 mmol) and white P-NH-P (0.57 g, 1.3 mmol) were ground with a pestle and mortar in a glovebox for 5 min to a dark blue powder and left overnight at room temperature. A solid product was obtained. Yield: 96% (0.75 g, 1.25 mmol). HRMS: calcd for C28H29Cl3CrNP2, 598.02 ([M]+); found, (ESI+): 598.04. EA: % Anal. Calcd (found) for with Mol wt for C28H29Cl3CrNP2: 598.02 g/mol: C, 56.07 (56.12); H, 4.87 (4.9); N, 2.34 (2.36). FT-IR (ATR, v/cm–1): 3171 (m, NH), 3053 (w, CH), 2950 (w, CH), 2924, (w, CH), 2870 (w, CH), 458 (m, Cr–Cl), 436 (m, Cr–Cl). μeff 4.19 μβ.

[Cr(P-NH-P′)Cl3] (4)

CrCl3(THF)3 (0.48 g, 1.3 mmol) and P-NH-P’ (0.77 g, 1.3 mmol) were mixed with a pestle and mortar in a glovebox for 5 min and left overnight at room temperature. A dark blue solid product was obtained. Crystals suitable for X-ray diffraction were obtained by recrystallization from acetonitrile at room temperature. Yield: 95% (0.85 g, 1.24 mmol). HRMS: calcd for C34H41Cl3CrNP2, 647.15 ([M – Cl]+); found, (ESI+): 647.21. EA: % Anal. Calcd for C34H41Cl3CrNP2: 683.97 g/mol: C, 59.41 (59.47); H, 6.26 (6.26); N, 1.96 (1.95). FT-IR (ATR, v/cm–1): 3286 (m, NH), 3056 (w, CH), 3030 (w, CH), 2918, 497 (m, Cr–Cl), 477 (m, Cr–Cl), 456 (w, Cr–Cl), 433 (w, Cr–Cl). μeff: 4.15 μβ.

General Procedure for Catalytic Study

The catalytic study was conducted based on a modified procedure.91 Briefly, the imine (0.2 mmol, 40 mg), the synthesized complex (5 mol %), Mg (10 mg), TMSCl (13 μL), and THF (2 mL) were stirred continuously in a vial under an argon atmosphere. The reaction mixture was transferred to a 50 mL stainless steel Parr hydrogenation reactor at constant temperatures (aided with an oil bath) and pressures (23 atm of H2(g)) with continuous stirring. The crude product was extracted in ethyl acetate (3 × 5 mL) after quenching the reaction with aqueous HCl (1 M, 2 mL). This was followed by drying the product in ethyl acetate over anhydrous Na2SO4 before removing the solvent in vacuo. The conversion was determined from the 19F NMR signal of the crude product (−115.6 ppm, 400 MHz, CDCl3) relative to that of the starting imine (−108.02 ppm, 400 MHz, CDCl3).

Acknowledgments

T.J.A. thanks the South African government for a Fellowship. R.H.M. thanks the NSERC (Canada) for a Discovery grant.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c02076.

  • Synthesis of the ligands and complexes, FTIR spectra, selected X-ray diffraction information, CVs, synthesis of the imine, spectra corresponding to the entries of Table 2, and synthesis of the dideuterated amine (PDF)

  • Crystal structure data for complex 2 (CIF)

  • Crystal structure data for complex 4 (CIF)

The authors declare no competing financial interest.

Supplementary Material

ao4c02076_si_001.pdf (2.4MB, pdf)
ao4c02076_si_002.cif (1.9MB, cif)
ao4c02076_si_003.cif (2.6MB, cif)

References

  1. Morris R. H. Asymmetric hydrogenation, transfer hydrogenation and hydrosilylation of ketones catalyzed by iron complexes. Chem. Soc. Rev. 2009, 38, 2282–2291. 10.1039/b806837m. [DOI] [PubMed] [Google Scholar]
  2. Bullock R. M.; Chen J. G.; Gagliardi L.; Chirik P. J.; Farha O. K.; Hendon C. H.; Jones C. W.; Keith J. A.; Klosin J.; Minteer S. D.; et al. Using nature’s blueprint to expand catalysis with earth-abundant metals. Science 2020, 369, eabc3183 10.1126/science.abc3183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Seo C. S.; Morris R. H. Catalytic homogeneous asymmetric hydrogenation: successes and opportunities. Organometallics 2019, 38, 47–65. 10.1021/acs.organomet.8b00774. [DOI] [Google Scholar]
  4. Manßen M.; Scott S. S.; Deng D.; Zheng C. H.; Schafer L. L. Accessing secondary amine containing fine chemicals and polymers with an earth-abundant hydroaminoalkylation catalyst. Green Chem. 2023, 25, 2629–2639. 10.1039/D3GC00011G. [DOI] [Google Scholar]
  5. Wen J.; Wang F.; Zhang X. Asymmetric hydrogenation catalyzed by first-row transition metal complexes. Chem. Soc. Rev. 2021, 50, 3211–3237. 10.1039/D0CS00082E. [DOI] [PubMed] [Google Scholar]
  6. Rana S.; Biswas J. P.; Paul S.; Paik A.; Maiti D. Organic synthesis with the most abundant transition metal-iron: from rust to multitasking catalysts. Chem. Soc. Rev. 2021, 50, 243–472. 10.1039/D0CS00688B. [DOI] [PubMed] [Google Scholar]
  7. Liu Y.; You T.; Wang H.-X.; Tang Z.; Zhou C.-Y.; Che C.-M. Iron-and cobalt-catalyzed C(sp 3)-H bond functionalization reactions and their application in organic synthesis. Chem. Soc. Rev. 2020, 49, 5310–5358. 10.1039/D0CS00340A. [DOI] [PubMed] [Google Scholar]
  8. Lu C. C.; Bill E.; Weyhermüller T.; Bothe E.; Wieghardt K. Neutral bis (α-iminopyridine) Metal complexes of the first-row transition ions (Cr, Mn, Fe, Co, Ni, Zn) and their monocationic analogues: mixed valency involving a redox noninnocent ligand system. J. Am. Chem. Soc. 2008, 130, 3181–3197. 10.1021/ja710663n. [DOI] [PubMed] [Google Scholar]
  9. Yang P.; Xu H.; Zhou J. Nickel-Catalyzed asymmetric transfer hydrogenation of olefins for the synthesis of α-and β-Amino acids. Angew. Chem., Int. Ed. 2014, 53, 12210–12213. 10.1002/anie.201407744. [DOI] [PubMed] [Google Scholar]
  10. Nifant’ev I. E.; Vinogradov A. A.; Vinogradov A. A.; Roznyatovsky V. A.; Grishin Y. K.; Ivanyuk A. V.; Sedov I. V.; Churakov A. V.; Ivchenko P. V. 5, 6-Dihydrodibenzo [c, e] [1, 2] azaphosphinine-based PNP ligands, Cr (0) coordination, and Cr(III) precatalysts for ethylene oligomerization. Organometallics 2018, 37, 2660–2664. 10.1021/acs.organomet.8b00427. [DOI] [Google Scholar]
  11. Blann K.; Bollmann A.; de Bod H.; Dixon J. T.; Killian E.; Nongodlwana P.; Maumela M. C.; Maumela H.; McConnell A. E.; Morgan D. H. Ethylene tetramerisation: subtle effects exhibited by N-substituted diphosphinoamine ligands. J. Catal. 2007, 249, 244–249. 10.1016/j.jcat.2007.04.009. [DOI] [Google Scholar]
  12. Suttil J. A.; Wasserscheid P.; McGuinness D. S.; Gardiner M. G.; Evans S. J. A Survey of pendant donor-functionalised (N, O) phosphine ligands for Cr-catalysed ethylene tri-and tetramerisation. Catal. Sci. Technol. 2014, 4, 2574–2588. 10.1039/C4CY00457D. [DOI] [Google Scholar]
  13. Zheng Q.; Zheng D.; Han B.; Liu S.; Li Z. Chromium complexes supported by the bidentate PN ligands: synthesis, characterization and application for ethylene polymerization. Dalton Trans. 2018, 47, 13459–13465. 10.1039/C8DT02834F. [DOI] [PubMed] [Google Scholar]
  14. McGuinness D. S.; Brown D. B.; Tooze R. P.; Hess F. M.; Dixon J. T.; Slawin A. M. Ethylene trimerization with Cr- PNP and Cr- SNS complexes: effect of ligand structure, metal oxidation state, and role of activator on catalysis. Organometallics 2006, 25, 3605–3610. 10.1021/om0601091. [DOI] [Google Scholar]
  15. Peitz S.; Peulecke N.; Müller B. H.; Spannenberg A.; Drexler H.-J.; Rosenthal U.; Al-Hazmi M. H.; Al-Eidan K. E.; Wöhl A.; Müller W. Heterobimetallic Al- Cl- Cr intermediates with relevance to the selective catalytic ethene trimerization systems consisting of CrCl3 (THF) 3, the aminophosphorus ligands Ph2PN(R)P(Ph)N(R)H, and triethylaluminum. Organometallics 2011, 30, 2364–2370. 10.1021/om200100g. [DOI] [Google Scholar]
  16. Zuo W.; Tauer S.; Prokopchuk D. E.; Morris R. H. Iron catalysts containing amine (imine) diphosphine P-NH-NP ligands catalyze both the asymmetric hydrogenation and asymmetric transfer hydrogenation of ketones. Organometallics 2014, 33, 5791–5801. 10.1021/om500479q. [DOI] [Google Scholar]
  17. Zuo W.; Prokopchuk D. E.; Lough A. J.; Morris R. H. Details of the mechanism of the asymmetric transfer hydrogenation of acetophenone using the amine (imine) diphosphine iron precatalyst: the base effect and the enantiodetermining step. ACS Catal. 2016, 6, 301–314. 10.1021/acscatal.5b01979. [DOI] [Google Scholar]
  18. Younus H. A.; Su W.; Ahmad N.; Chen S.; Verpoort F. Ruthenium pincer complexes: synthesis and catalytic applications. Adv. Synth. Catal. 2015, 357, 283–330. 10.1002/adsc.201400777. [DOI] [Google Scholar]
  19. Wang H.; Wen J.; Zhang X. Chiral tridentate ligands in transition metal-catalyzed asymmetric hydrogenation. Chem. Rev. 2021, 121, 7530–7567. 10.1021/acs.chemrev.1c00075. [DOI] [PubMed] [Google Scholar]
  20. Vielhaber T.; Faust K.; Topf C. Group 6 metal carbonyl complexes supported by a bidentate PN ligand: syntheses, characterization, and catalytic hydrogenation activity. Organometallics 2020, 39, 4535–4543. 10.1021/acs.organomet.0c00612. [DOI] [Google Scholar]
  21. Venderbosch B.; Wolzak L. A.; Oudsen J.-P. H.; de Bruin B.; Korstanje T. J.; Tromp M. role of the ligand and activator in selective Cr-PNP ethene tri-and tetramerization catalysts-a spectroscopic study. Catal. Sci. Technol. 2020, 10, 6212–6222. 10.1039/D0CY01168A. [DOI] [Google Scholar]
  22. Shaikh Y.; Gurnham J.; Albahily K.; Gambarotta S.; Korobkov I. Aminophosphine-based chromium catalysts for selective ethylene tetramerization. Organometallics 2012, 31, 7427–7433. 10.1021/om3007135. [DOI] [Google Scholar]
  23. Simler T.; Frison G.; Braunstein P.; Danopoulos A. A. Direct synthesis of doubly deprotonated, dearomatised lutidine PNP Cr and Zr pincer complexes based on isolated K and Li ligand transfer reagents. Dalton Trans. 2016, 45, 2800–2804. 10.1039/C6DT00144K. [DOI] [PubMed] [Google Scholar]
  24. Chiu P. L.; Lee H. M. Chemistry of the PCNHCP ligand: silver and ruthenium complexes, facial/meridional coordination, and catalytic transfer hydrogenation. Organometallics 2005, 24, 1692–1702. 10.1021/om049070v. [DOI] [Google Scholar]
  25. Mikhailine A.; Lough A. J.; Morris R. H. Efficient asymmetric transfer hydrogenation of ketones catalyzed by an iron complex containing a P- N- N- P tetradentate ligand formed by template synthesis. J. Am. Chem. Soc. 2009, 131, 1394–1395. 10.1021/ja809493h. [DOI] [PubMed] [Google Scholar]
  26. Mikhailine A. A.; Maishan M. I.; Lough A. J.; Morris R. H. The mechanism of efficient asymmetric transfer hydrogenation of acetophenone using an iron(II) complex containing an (S, S)-Ph2PCH2CH = NCHPhCHPhN = CHCH2PPh2 ligand: partial ligand reduction is the key. J. Am. Chem. Soc. 2012, 134, 12266–12280. 10.1021/ja304814s. [DOI] [PubMed] [Google Scholar]
  27. Sonnenberg J. F.; Coombs N.; Dube P. A.; Morris R. H. Iron nanoparticles catalyzing the asymmetric transfer hydrogenation of ketones. J. Am. Chem. Soc. 2012, 134, 5893–5899. 10.1021/ja211658t. [DOI] [PubMed] [Google Scholar]
  28. Lagaditis P. O.; Lough A. J.; Morris R. H. Iron complexes for the catalytic transfer hydrogenation of acetophenone: steric and electronic effects imposed by alkyl substituents at Phosphorus. Inorg. Chem. 2010, 49, 10057–10066. 10.1021/ic101366z. [DOI] [PubMed] [Google Scholar]
  29. Meyer N.; Lough A. J.; Morris R. H. Iron(II) complexes for the efficient catalytic asymmetric transfer hydrogenation of ketones. Chem.—Eur. J. 2009, 15, 5605–5610. 10.1002/chem.200802458. [DOI] [PubMed] [Google Scholar]
  30. Huber R.; Passera A.; Gubler E.; Mezzetti A. P-Stereogenic PN (H) P iron(II) catalysts for the asymmetric hydrogenation of ketones: the importance of non-covalent interactions in rational ligand design by computation. Adv. Synth. Catal. 2018, 360, 2900–2913. 10.1002/adsc.201800433. [DOI] [Google Scholar]
  31. Uchiike C.; Ouchi M.; Ando T.; Kamigaito M.; Sawamoto M. Evolution of iron catalysts for effective living radical polymerization: P-N chelate ligand for enhancement of catalytic performances. J. Polym. Sci. 2008, 46, 6819–6827. 10.1002/pola.22990. [DOI] [Google Scholar]
  32. Passera A.; Mezzetti A. Mn(I) and Fe(II)/PN(H)P catalysts for the hydrogenation of ketones: a comparison by experiment and calculation. Adv. Syn. Catal. 2019, 361, 4691–4706. 10.1002/adsc.201900671. [DOI] [Google Scholar]
  33. Seo C. S.; Tsui B. T.; Gradiski M. V.; Smith S. A.; Morris R. H. Enantioselective direct, base-free hydrogenation of ketones by a manganese amido complex of a homochiral, unsymmetrical P-N-P′ ligand. Catal. Sci. Technol. 2021, 11, 3153–3163. 10.1039/D1CY00446H. [DOI] [Google Scholar]
  34. Wang H.; Zhang Y.; Yang T.; Guo X.; Gong Q.; Wen J.; Zhang X. Chiral electron-rich PNP ligand with a phospholane motif: structural features and application in asymmetric hydrogenation. Org. Lett. 2020, 22, 8796–8801. 10.1021/acs.orglett.0c03159. [DOI] [PubMed] [Google Scholar]
  35. DeGroot H. P.; Hanusa T. P. Solvate-assisted grinding: metal solvates as solvent sources in mechanochemically driven organometallic reactions. Organometallics 2021, 40, 3516–3525. 10.1021/acs.organomet.1c00316. [DOI] [Google Scholar]
  36. Guo R.; Lu S.; Chen X.; Tsang C.-W.; Jia W.; Sui-Seng C.; Amoroso D.; Abdur-Rashid K. Synthesis of chiral aminophosphines from chiral aminoalcohols via cyclic sulfamidates. J. Org. Chem. 2010, 75, 937–940. 10.1021/jo902302c. [DOI] [PubMed] [Google Scholar]
  37. Mikhailine A. A.; Lagaditis P. O.; Sues P. E.; Lough A. J.; Morris R. H. New cyclic phosphonium salts derived from the reaction of phosphine-aldehydes with acid. J. Organomet. Chem. 2010, 695, 1824–1830. 10.1016/j.jorganchem.2010.04.016. [DOI] [Google Scholar]
  38. Smith S. A.; Lagaditis P. O.; Lüpke A.; Lough A. J.; Morris R. H. Unsymmetrical iron P-NH-P′ catalysts for the asymmetric pressure hydrogenation of aryl ketones. Chem.—Eur. J. 2017, 23, 7212–7216. 10.1002/chem.201701254. [DOI] [PubMed] [Google Scholar]
  39. Sonnenberg J. F.; Lough A. J.; Morris R. H. Synthesis of iron PNP′ and P-NH-P′ asymmetric hydrogenation catalysts. Organometallics 2014, 33, 6452–6465. 10.1021/om5008083. [DOI] [Google Scholar]
  40. Cotton F. A.; LaPrade M. D. Crystal and molecular structures of [N,N-bis(2-diphenylphosphinoethyl) ethylamine] tricarbonylchromium. J. Am. Chem. Soc. 1969, 91, 7000–7005. 10.1021/ja01053a018. [DOI] [Google Scholar]
  41. Fryzuk M. D.; Leznoff D. B.; Rettig S. J.; Thompson R. C. Magnetic exchange in dinuclear chromium(II) complexes: effect of bridging chlorides and bridging hydrides in antiferromagnetic coupling. Inorg. Chem. 1994, 33, 5528–5534. 10.1021/ic00102a029. [DOI] [Google Scholar]
  42. Fryzuk M. D.; Leznoff D. B.; Rettig S. J. Synthesis and structure of even-electron paramagnetic chromium(II) complexes. Organometallics 1995, 14, 5193–5202. 10.1021/om00011a042. [DOI] [Google Scholar]
  43. McGuinness D. S.; Wasserscheid P.; Keim W.; Hu C.; Englert U.; Dixon J. T.; Grove C. novel Cr-PNP complexes as catalysts for the trimerisation of ethylene. Chem. Commun. 2003, 334–335. 10.1039/b210878j. [DOI] [PubMed] [Google Scholar]
  44. Bluhm M. E.; Walter O.; Döring M. Chromium imine and amine complexes as homogeneous catalysts for the trimerisation and polymerisation of ethylene. J. Organomet. Chem. 2005, 690, 713–721. 10.1016/j.jorganchem.2004.09.080. [DOI] [Google Scholar]
  45. Liu Q.; Gao R.; Hou J.; Sun W. Tridentate P̂N̂P Chromium Complexes: Synthesis, Character-ization and Their Ethylene Oligomerization and Polymerization. Chin. J. Org. Chem. 2013, 33, 808. 10.6023/cjoc201302022. [DOI] [Google Scholar]
  46. Kuriyama S.; Arashiba K.; Nakajima K.; Tanaka H.; Yoshizawa K.; Nishibayashi Y. Azaferrocene-based PNP-Type pincer ligand: synthesis of molybdenum, chromium, and iron complexes and reactivity toward nitrogen fixation. Eur. J. Inorg. Chem. 2016, 2016, 4856–4861. 10.1002/ejic.201601051. [DOI] [Google Scholar]
  47. Alzamly A.; Gambarotta S.; Korobkov I. Polymer-free ethylene oligomerization using a pyridine-based pincer PNP-type of ligand. Organometallics 2013, 32, 7204–7212. 10.1021/om4008828. [DOI] [Google Scholar]
  48. Alzamly A.; Gambarotta S.; Korobkov I. Reactivity with alkylaluminum of a chromium complex of a pyridine-containing PNP ligand: redox N-P bond cleavage. Organometallics 2014, 33, 1602–1607. 10.1021/om401129w. [DOI] [Google Scholar]
  49. Mastalir M.; Glatz M.; Stöger B.; Weil M.; Pittenauer E.; Allmaier G.; Kirchner K. Synthesis, characterization and reactivity of vanadium, chromium, and manganese PNP pincer complexes. Inorg. Chim. Acta 2017, 455, 707–714. 10.1016/j.ica.2016.02.064. [DOI] [Google Scholar]
  50. Gong D.; Liu W.; Chen T.; Chen Z.-R.; Huang K.-W. Ethylene polymerization by PN3-type pincer chromium(III) complexes. J. Mol. Catal. A: Chem. 2014, 395, 100–107. 10.1016/j.molcata.2014.08.005. [DOI] [Google Scholar]
  51. Alzamly A.; Gambarotta S.; Korobkov I. Synthesis, structures, and ethylene oligomerization activity of bis(phosphanylamine) pyridine chromium/aluminate complexes. Organometallics 2013, 32, 7107–7115. 10.1021/om4008289. [DOI] [Google Scholar]
  52. Mastalir M.; De Aguiar S. R.; Glatz M.; Stöger B.; Kirchner K. A Convenient solvothermal synthesis of group 6 PNP pincer tricarbonyl complexes. Organometallics 2016, 35, 229–232. 10.1021/acs.organomet.5b00940. [DOI] [Google Scholar]
  53. Fryzuk M. D.; Leznoff D. B.; Rettig S. J. Synthesis and structure of a five-coordinate organochromium(III) thiolate complex from a chromium(II) alkyl precursor. Organometallics 1997, 16, 5116–5119. 10.1021/om970574l. [DOI] [Google Scholar]
  54. Fryzuk M. D.; Leznoff D. B.; Rettig S. J.; Young Jr V. G. One-electron oxidation of Paramagnetic chromium(II) alkyl complexes with alkyl halides: synthesis and structure of five-coordinate chromium(III) complexes. J. Chem. Soc., Dalton Trans. 1999, 147–154. 10.1039/a806099a. [DOI] [Google Scholar]
  55. Simler T.; Braunstein P.; Danopoulos A. A. Chromium(II) pincer complexes with dearomatized PNP and PNC ligands: a comparative study of their catalytic ethylene oligomerization activity. Organometallics 2016, 35, 4044–4049. 10.1021/acs.organomet.6b00685. [DOI] [Google Scholar]
  56. Kallmeier F.; Fertig R.; Irrgang T.; Kempe R. Chromium-catalyzed alkylation of amines by alcohols. Angew. Chem., Int. Ed. 2020, 59, 11789–11793. 10.1002/anie.202001704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Ott J. C.; Isak D.; Melder J. J.; Wadepohl H.; Gade L. H. Single or paired? Structure and reactivity of PNP-Chromium(II) hydrides. Inorg. Chem. 2020, 59, 14526–14535. 10.1021/acs.inorgchem.0c02315. [DOI] [PubMed] [Google Scholar]
  58. Prokopchuk D. E.; Morris R. H. Inner-sphere activation, outer-sphere catalysis: theoretical study on the mechanism of transfer hydrogenation of ketones using iron(II) PNNP eneamido complexes. Organometallics 2012, 31, 7375–7385. 10.1021/om300572v. [DOI] [Google Scholar]
  59. Demmans K. Z.; Olson M. E.; Morris R. H. Asymmetric transfer hydrogenation of ketones with well-defined manganese(I) PNN and PNNP complexes. Organometallics 2018, 37, 4608–4618. 10.1021/acs.organomet.8b00625. [DOI] [Google Scholar]
  60. Sui-Seng C.; Freutel F.; Lough A. J.; Morris R. H. Highly efficient catalyst systems using iron complexes with a tetradentate PNNP ligand for the asymmetric hydrogenation of polar bonds. Angew. Chem., Int. Ed. 2008, 47, 940–943. 10.1002/anie.200705115. [DOI] [PubMed] [Google Scholar]
  61. Lagaditis P. O.; Lough A. J.; Morris R. H. Low-valent ene-amido iron complexes for the asymmetric transfer hydrogenation of acetophenone without base. J. Am. Chem. Soc. 2011, 133, 9662–9665. 10.1021/ja202375y. [DOI] [PubMed] [Google Scholar]
  62. Mikhailine A. A.; Maishan M. I.; Morris R. H. Asymmetric transfer hydrogenation of ketimines using well-defined iron(II)-based precatalysts containing a PNNP ligand. Org. Lett. 2012, 14, 4638–4641. 10.1021/ol302079q. [DOI] [PubMed] [Google Scholar]
  63. Sues P. E.; Lough A. J.; Morris R. H. Stereoelectronic factors in iron catalysis: synthesis and characterization of aryl-substituted iron(II) carbonyl P-N-N-P complexes and their use in the asymmetric transfer hydrogenation of ketones. Organometallics 2011, 30, 4418–4431. 10.1021/om2005172. [DOI] [Google Scholar]
  64. Prokopchuk D. E.; Sonnenberg J. F.; Meyer N.; Zimmer-De Iuliis M.; Lough A. J.; Morris R. H. A spectroscopic and DFT study of ferraaziridine complexes formed in the transfer hydrogenation of acetophenone catalyzed using trans-[Fe(CO)(NCMe)(PPh2C6H4CH = NCH2-)24P,N,N,P](BF4)2. Organometallics 2012, 31, 3056–3064. 10.1021/om201170f. [DOI] [Google Scholar]
  65. Mikhailine A. A.; Morris R. H. Effect of the structure of the diamine backbone of P- N- N- P ligands in iron(II) complexes on catalytic activity in the transfer hydrogenation of acetophenone. Inorg. Chem. 2010, 49, 11039–11044. 10.1021/ic101548j. [DOI] [PubMed] [Google Scholar]
  66. Zuo W.; Lough A. J.; Li Y. F.; Morris R. H. Amine (imine) diphosphine iron catalysts for asymmetric transfer hydrogenation of ketones and imines. Science 2013, 342, 1080–1083. 10.1126/science.1244466. [DOI] [PubMed] [Google Scholar]
  67. Lagaditis P. O.; Mikhailine A. A.; Lough A. J.; Morris R. H. Template synthesis of iron(II) complexes containing tridentate P-N-S, P-N-P, P-N-N, and tetradentate P-N-N-P ligands. Inorg. Chem. 2010, 49, 1094–1102. 10.1021/ic901945c. [DOI] [PubMed] [Google Scholar]
  68. Ajayi T. J.; Shapi M. Solvent-free mechanochemical synthesis, Hirshfeld surface analysis, crystal structure, spectroscopic characterization and NBO analysis of bis (ammonium) bis ((4-methoxyphenyl) phosphonodithioato)-nickel(II) dihydrate with DFT studies. J. Mol. Struct. 2020, 1202, 127254. 10.1016/j.molstruc.2019.127254. [DOI] [Google Scholar]
  69. Ajayi T. J.; Pillay M. N.; van Zyl W. E. Solvent-free mechanochemical synthesis of dithiophosphonic acids and corresponding nickel(II) complexes. Phosphorus, Sulfur Silicon Relat. Elem. 2017, 192, 1205–1211. 10.1080/10426507.2017.1358717. [DOI] [Google Scholar]
  70. Maser L.; Vogt M.; Langer R. Facial vs. meridional coordination modes in ReI tricarbonyl complexes with a carbodiphosphorane-based tridentate ligand. Anorg. Allg. Chem. 2021, 647, 1518–1523. 10.1002/zaac.202100151. [DOI] [Google Scholar]
  71. Tamayo A. B.; Alleyne B. D.; Djurovich P. I.; Lamansky S.; Tsyba I.; Ho N. N.; Bau R.; Thompson M. E. Synthesis and characterization of facial and meridional tris-cyclometalated iridium(III) complexes. J. Am. Chem. Soc. 2003, 125, 7377–7387. 10.1021/ja034537z. [DOI] [PubMed] [Google Scholar]
  72. McDaniel A. M.; Tseng H.-W.; Damrauer N. H.; Shores M. P. Synthesis and solution phase characterization of strongly photooxidizing heteroleptic Cr(III) tris-dipyridyl complexes. Inorg. Chem. 2010, 49, 7981–7991. 10.1021/ic1009972. [DOI] [PubMed] [Google Scholar]
  73. Lever A. P.Inorganic Electronic Spectroscopy; Elsevier Publishing Company: Amsterdam, The Netherlands, 1984; Vol. 33. [Google Scholar]
  74. Moulin J. O.; Evans J.; McGuinness D. S.; Reid G.; Rucklidge A. J.; Tooze R. P.; Tromp M. Probing the effects of ligand structure on activity and selectivity of Cr(III) complexes for ethylene oligomerisation and polymerisation. Dalton Trans. 2008, 1177–1185. 10.1039/b716078j. [DOI] [PubMed] [Google Scholar]
  75. Figgis B.Introduction to Ligand Fields″; Interscience Publ: New York, USA, 1966; Vol. 5. [Google Scholar]
  76. Sinha N.; Yaltseva P.; Wenger O. S. The nephelauxetic effect becomes an important design factor for photoactive first-Row transition metal complexes. Angew. Chem., Int. Ed. 2023, 62, e202303864 10.1002/anie.202303864. [DOI] [PubMed] [Google Scholar]
  77. Sharma A. K.; Chandra S. Complexation of nitrogen and sulphur donor Schiff’s base ligand to Cr(III) and Ni(II) metal ions: synthesis, spectroscopic and antipathogenic studies. Spectrochim. Acta A Mol. 2011, 78, 337–342. 10.1016/j.saa.2010.10.017. [DOI] [PubMed] [Google Scholar]
  78. Chandra S.; Sharma A. K. Applications of several spectral techniques to characterize coordination compounds derived from 2, 6-diacetylpyridine derivative. Spectrochim. Acta A Mol. 2009, 74, 271–276. 10.1016/j.saa.2009.06.014. [DOI] [PubMed] [Google Scholar]
  79. Porterfield William W.Inorganic Chemistry, A Unified Approach; Addison Wesley Pub. Co: Indiana, USA, 1984. [Google Scholar]
  80. De Buysser K.; Herman G.; Bruneel E.; Hoste S.; Van Driessche I. Determination of the number of unpaired electrons in metal-complexes. A comparison between the Evans’ method and susceptometer results. J. Chem. Phys. 2005, 315, 286–292. 10.1016/j.chemphys.2005.04.022. [DOI] [Google Scholar]
  81. Su J.; Duan L.; Zheng W. Chromium(III) complexes with 1, 2, 4-diazaphospholide and 2, 6-bis (N-1, 2, 4-diazaphosphol-1-yl) pyridine ligands: synthesis, X-ray structural characterization, EPR spectroscopy analysis, and magnetic susceptibility studies. Z. Naturforsch. B Chem. Sci. 2016, 71, 795–802. 10.1515/znb-2015-0216. [DOI] [Google Scholar]
  82. Gabriel C.; Raptopoulou C. P.; Terzis A.; Lalioti N.; Salifoglou A. Synthesis, structural, spectroscopic and magnetic susceptibility studies of a soluble Cr(III)-heida (2-hydroxyethyliminodiacetic acid) complex. Relevance to aqueous chromium(III)-heida speciation. Inorg. Chim. Acta 2007, 360, 513–522. 10.1016/j.ica.2006.07.097. [DOI] [Google Scholar]
  83. Clément N.; Toussaint C.; Rogez G.; Loose C.; Kortus J.; Brelot L.; Choua S.; Dagorne S.; Turek P.; Welter R. Novel CrIII dinuclear complexes supported by salicyloylhydrazono dithiolane and dithiane ligands: synthesis, stability, crystal structures and magnetic properties. Dalton Trans. 2010, 39, 4579–4585. 10.1039/c001057j. [DOI] [PubMed] [Google Scholar]
  84. Kim D.; Chang J. Determination of stannous chloride and stannic bromide speciation in concentrated Cl and Br media by cyclic voltammetry. J. Electroanal. Chem. 2017, 785, 20–25. 10.1016/j.jelechem.2016.12.004. [DOI] [Google Scholar]
  85. Murugappan K.; Arrigan D. W.; Silvester D. S. Electrochemical behavior of chlorine on platinum microdisk and screen-printed electrodes in a room temperature ionic liquid. J. Phys. Chem. C 2015, 119, 23572–23579. 10.1021/acs.jpcc.5b07753. [DOI] [Google Scholar]
  86. Popov I. A.; Mehio N.; Chu T.; Davis B. L.; Mukundan R.; Yang P.; Batista E. R. Impact of ligand substitutions on multielectron redox properties of Fe complexes supported by nitrogenous chelates. ACS Omega 2018, 3, 14766–14778. 10.1021/acsomega.8b01921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. King A. P.; Gellineau H. A.; MacMillan S. N.; Wilson J. J. Physical properties, ligand substitution reactions, and biological activity of Co(III)-Schiff base complexes. Dalton Trans. 2019, 48, 5987–6002. 10.1039/C8DT04606A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Schnider P.; Koch G.; Prétôt R.; Wang G.; Bohnen F. M.; Krüger C.; Pfaltz A. Enantioselective hydrogenation of imines with chiral (phosphanodihydrooxazole) iridium catalysts. Chem.—Eur. J. 1997, 3, 887–892. 10.1002/chem.19970030609. [DOI] [Google Scholar]
  89. Aranda B.; Valdebenito G.; Parra-Melipán S.; López V.; Moya S.; Vega A.; Aguirre P. Hydrogenation of imines catalyzed by ruthenium(II) complexes containing phosphorus-nitrogen ligands via hydrogen transfer reaction. Mol. Catal. 2022, 526, 112374. 10.1016/j.mcat.2022.112374. [DOI] [Google Scholar]
  90. Bagh B.; Stephan D. W. Half sandwich ruthenium(II) hydrides: hydrogenation of terminal, internal, cyclic and functionalized olefins. Dalton Trans. 2014, 43, 15638–15645. 10.1039/C4DT02407A. [DOI] [PubMed] [Google Scholar]
  91. Ling L.; Hu C.; Long L.; Zhang X.; Zhao L.; Liu L. L.; Chen H.; Luo M.; Zeng X. chromium-catalyzed stereodivergent E-and Z-selective alkyne hydrogenation controlled by cyclic (alkyl)(amino) carbene ligands. Nat. Commun. 2023, 14, 990. 10.1038/s41467-023-36677-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Kariofillis S. K.; Doyle A. G. Synthetic and mechanistic implications of chlorine photoelimination in nickel/photoredox C(sp(3))-H cross-coupling. Acc. Chem. Res. 2021, 54, 988–1000. 10.1021/acs.accounts.0c00694. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

ao4c02076_si_001.pdf (2.4MB, pdf)
ao4c02076_si_002.cif (1.9MB, cif)
ao4c02076_si_003.cif (2.6MB, cif)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES