Skip to main content
ACS Omega logoLink to ACS Omega
. 2019 Oct 31;4(20):18833–18845. doi: 10.1021/acsomega.9b02828

XAS Analysis of Reactions of (Arylimido)vanadium(V) Dichloride Complexes Containing Anionic NHC That Contains a Weakly Coordinating B(C6F5)3 Moiety (WCA-NHC) or Phenoxide Ligands with Al Alkyls: A Potential Ethylene Polymerization Catalyst with WCA-NHC Ligands

Kotohiro Nomura †,*, Go Nagai , Itsuki Izawa , Takato Mitsudome , Matthias Tamm §,*, Seiji Yamazoe †,*
PMCID: PMC6854829  PMID: 31737845

Abstract

graphic file with name ao9b02828_0006.jpg

(Arylimido)vanadium(V) dichloride complexes containing anionic N-heterocyclic carbene (NHC) ligands that contain weakly coordinating B(C6F5)3 moieties (WCA-NHC) of the type [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar = 2,6-Me2C6H3, Ar′ = 2,6-iPr2C6H3) showed significant catalytic activity for ethylene polymerization in the presence of Al cocatalysts (MAO and AliBu3); the activity by the 5–MAO catalyst (19 500 kg-PE/mol-V·h; TOF 11 600 min–1) is the highest among those reported using the other (imido)vanadium(V) complexes in the presence of MAO, and the 5–AliBu3 catalyst showed higher activity (66 000 kg-PE/mol-V·h; TOF 39 200 min–1). The V K-edge X-ray absorption near-edge structure (XANES) analyses (in toluene) strongly suggest the formation of certain vanadium(III) species by reduction with AliBu3 accompanying structural changes; the EXAFS analysis suggests the presence of the arylimido ligand and one V–Cl bond (2.34 ± 0.04 Å), which is longer than those [2.1901(8)–2.2462(8) Å] in the reported (imido)vanadium(V) complexes. The XANES analysis of [V(NAr)Cl2(OAr)] strongly suggests the formation of the other vanadium(III) species by reduction with Me2AlCl or Et2AlCl, and the EXAFS analysis suggests the presence of the arylimido ligand and two V–Cl bonds (2.45 ± 0.03 Å). The XANES spectra showed no significant changes in both the pre-edge peak(s) and the edge peak when these complexes were treated with MAO, suggesting that the basic geometry and the high oxidation state were preserved under these conditions.

Introduction

Transition-metal-catalyzed olefin polymerization is the core method in commercial production of polyolefin [exemplified as polyethylene (PE) and polypropylene (PP)], and the design of the molecular catalysts has been considered as an attractive subject especially in terms of synthesis of new polymers.117 Classical Ziegler-type vanadium catalyst systems (e.g., VOCl3, VCl4, VCl3–AlBr3, AlCl3–AlPh3, AliBu3, and SnPh4) are known to exhibit high reactivity toward olefins (notable propagation for synthesis of ultrahigh-molecular-weight polymers with narrow distributions)1823 and are applied for commercial production of EPDM (ethylene propylene diene monomer) rubber.8,17,2426,28

Although potentially high initial catalytic activities were exhibited using these catalyst systems, their rapid catalyst deactivation [probably associated with conversion into the inactive species by reduction, from vanadium(III) to vanadium(II)] causes very poor overall productivities. The concern, short catalyst life, could be overcome partly by addition of ethyl trichloroacetate (ETA),27 and this would be assumed as due to the catalyst re-activation by re-oxidation with ETA (Scheme 1);27 ethylene polymerizations with most of these catalysts were thus performed in the presence of large excess of ETA.8,17,2426 Therefore, vanadium(III) species have been assumed to play a role as the active species on the basis of electron spin resonance (ESR) spectra and titration study,2932 as shown in Scheme 1.21,22,25 However, as described below, ESR spectroscopy, widely employed for analysis of paramagnetic compounds,3238 faces concerns for the observation of “ESR-silent” vanadium(III) species and the quantitative analysis and for obtainment of structural information.

Scheme 1. Assumed Ethylene Polymerization Mechanism in the Presence of Ethyl Trichloroacetate (ETA)25.

Scheme 1

Studies on synthesis of the efficient vanadium complex catalysts and the related organometallic chemistry have been considered as important subjects in the fields of catalysis, organometallic chemistry, and polymer chemistry.8,16,17,28 (Adamantylimido)vanadium(V) complexes containing 2-anilidemethyl-pyridine, [V(NAd)X2(2-ArNCH2C5H4N)], [A; Ad = 1-adamantyl; Ar = 2,6-Me2C6H3; X = Cl, Me; Scheme 2],39,40 or 8-anilide-5,6,7-trihydroquinoline41 ligands have been known to exhibit significant catalytic activities for selective ethylene dimerization in the presence of MAO cocatalysts, whereas the related complexes containing 2-(2′-benzimidazolyl)pyridine ligands (B) showed the notable activities for ethylene polymerization in the presence of Me2AlCl cocatalysts.42 The (arylimido)vanadium(V) complexes that contain a monodentate anionic ancillary donor ligand (Y), [V(NAr)Cl2(Y)] [Y = phenoxide (C),4345 iminoimidazolide (D),46 and iminoimidazolidide (E),46Scheme 2], showed the high activities for ethylene polymerization and the copolymerization with norbornene (NBE) in the presence of MAO or Me2AlCl and Et2AlCl. Moreover, the complexes containing anionic N-heterocyclic carbenes (NHCs) that contain a weakly coordinating borate moiety (WCA-NHC, F) exhibited high catalytic activities for ethylene polymerization in the presence of AliBu3,47 which has been considered as an ineffective Al cocatalyst in the metal-catalyzed olefin polymerization.117,2426,28,3946

Scheme 2. Selected (Imido)vanadium(V) Dichloride Complex Catalysts for Ethylene Polymerization and Dimerization3947.

Scheme 2

It was demonstrated that cationic alkyl species play a role in the ethylene dimerization using the (A)–MAO catalyst, on the basis of (i) 51V NMR and electron spin resonance (ESR) spectra,40 (ii) the effect of the Al cocatalyst and the ethylene pressure dependence,40 (iii) syntheses of the cationic (and the dimethyl) complexes and their reaction chemistry,48 and (iv) analysis through solution V K-edge XANES (XANES = X-ray absorption near-edge structure) and the FT-EXAFS (EXAFS = extended X-ray absorption fine structure) spectra.48 Analysis by synchrotron X-ray absorption spectroscopy (XAS) has been known to provide important information of the oxidation states (through XANES analysis; pre-edge and edge peaks) and their coordination atoms around the centered metal (through FT-EXAFS analysis).49,50 XAS analysis has been widely used in the field of heterogeneous catalysis,4953 and the reports for studies in the field of molecular catalysis are still limited.42,48,5460

We recently introduced that the method should be useful for analysis of the vanadium catalyst solution, especially for ESR-silent paramagnetic species, which cannot be observed by both NMR and ESR spectra.42,55,56,58,59 For example, 51V NMR (in toluene-d8) and ESR (in toluene) spectra showed negligible resonances when complex B was treated with Me2AlCl (10.0 equiv) or when the toluene solution of complex F was treated with AliBu3 (10.0 equiv). In contrast, as shown in Figure 1 (XANES spectra), apparent changes in the oxidation state were observed when the toluene solution containing B was treated with Me2AlCl (10.0 equiv, Figure 1a)42 or F (complex 3 in Scheme 3) was treated with AliBu3 (10.0 equiv, Figure 1b);59 the spectra did not change when these complexes were treated with MAO. The low-energy shifts in the edge absorption accompanied by changes in the pre-edge peak strongly suggest that these complexes (B and F) were reduced by Al alkyls to form certain vanadium(III) species. These results are thus consistent with those observed in the 51V NMR and ESR spectra.42,58

Figure 1.

Figure 1

V K-edge XANES spectra (in toluene at 25 °C; 5.46 keV) for (a) [V(NAd)Cl2(L)] [B; Ad = 1-adamantyl; L = 2-(2′-benzimidazolyl)-6-methylpyridine]42 and (b) [V(N-2,6-Me2C6H3)Cl2(WCA-NHC)] (F)58,59 in the presence of cocatalysts (MAO, Me2AlCl, AliBu3, 10.0 equiv).

Scheme 3. (Arylimido)vanadium(V) Complexes Coordinated with Anionic N-Heterocyclic Carbenes (NHCs) Containing a Weakly Coordinating tris(Pentafluorophenyl) Borane Moiety [WCA-NHC-Ar (13), Ar = 2,6-Me2C6H3; WCA-NHC-Ar′ (4, 5), Ar′ = 2,6-iPr2C6H3].

Scheme 3

The above results clearly suggest the formation of certain vanadium(III) species by treatment of Me2AlCl or AliBu3,42,58,59 whereas the high oxidation state was preserved by addition of MAO.42,58 Moreover, it is clear that the XANES spectrum of the toluene solution consisting of B-Me2AlCl should be different from that consisting of F-AliBu3, especially in the edge region (Figure 1). In contrast, the oxidation state was preserved when the dichloro complex containing the 2-anilidemethylpyridine ligand (A) was treated with Me2AlCl (10.0 equiv), which afforded ultrahigh-molecular-weight polymers in the reaction with ethylene; the formation of another vanadium(V) species was thus suggested from 51V NMR spectra and V K-edge XANES spectra.48 Since, as described below, a new (arylimido)vanadium(V) dichloride complex containing the WCA-NHC-Ar′ ligand (5, Scheme 3, Ar′ = 2,6-iPr2C6H3) showed promising characteristics as the ethylene polymerization catalyst, we thus explored the effect of Al cocatalysts on the oxidation state of the formed vanadium species by the solution V K-edge XANES and EXAFS analyses (especially with complexes 3 and 5). Moreover, the studies with the phenoxide analogues [exemplified as [V(NAr)Cl2(OAr)] (C), Scheme 2] were also explored because the effect of the Al cocatalyst (MAO vs Me2AlCl) on both the activity and comonomer incorporation by C was significant in the ethylene/NBE copolymerization.44,45 Through this study, we explored for the obtainment of the information of the active species, especially vanadium(III), formed by reaction with Al alkyls, which should be important for understanding the catalysis mechanism in the ethylene polymerization.

Results and Discussion

Synthesis of (Arylimido)vanadium(V) Complexes Containing Anionic N-Heterocyclic Carbenes That Contain a Weakly Coordinating B(C6F5)3 (WCA-NHC), [V(NR′)Cl2(WCA-NHC)], and Their Use as the Catalyst Precursors for Ethylene Polymerization

Synthesis and Structural Analysis of [V(NR′)Cl2(WCA-NHC-Ar′)] (Ar′ = 2,6-iPr2C6H3)

Anionic N-heterocyclic carbene–borate (WCA-NHC) ligands have been promising anionic ligands47,6166 because it has been postulated that electron-deficient neutral or cationic early transition-metal-alkyl species with high oxidation states can be stabilized with the WCA-NHC ligands as anionic donor ligands through σ- and/or π-donation and/or by delocalization.47,66

(Arylimido)vanadium(V) dichloride complexes containing 2,6-diisopropylphenyl-modified WCA-NHC ligands of type [V(NR′)Cl2(WCA-NHC-Ar′)] [R′ = C6H5 (4), Ar (5); Ar = 2,6-Me2C6H3; Ar′ = 2,6-iPr2C6H3] were prepared according to the reported procedure for syntheses of [V(NR′)Cl2(WCA-NHC-Ar)] [R′ = Ad (1), C6H5 (2), Ar (3); Scheme 3]47 by treating V(NR′)Cl3 with lithium salts of the 2,6-diisopropylphenyl analogue64,65 in toluene. The resultant complexes were isolated as microcrystals grown from cold CH2Cl2n-hexane (−30 °C); the complexes were identified by NMR spectra and elemental analysis, and their structures were determined by X-ray diffraction analysis (Figure 2).a It was revealed that the resonance ascribed to 5 (410.5 ppm) in the 51V NMR spectrum was shifted low field compared to those in 1 (45.4 ppm), 2 (185.6 ppm), 3 (302.7 ppm), and 4 (256.6 ppm).

Figure 2.

Figure 2

ORTEP drawings of [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, left, Ar′ = 2,6-iPr2C6H3) and [V(NAr)Cl2(WCA-NHC-Ar′)] (5, right, Ar = 2,6-Me2C6H3). Thermal ellipsoids were drawn at 30% probability level, and H atoms were omitted for clarity. Selected bond lengths (Å) for 4: V–Ccarbene 2.060(2), V–N(1) 1.646(3), V–Cl(1) 2.1641(7), V–Cl(2) 2.1655(7), and N(1)–C(1) 1.380(4). Selected bond angles (°) for 4: Ccarbene–V–Cl(1) 113.65(6), Ccarbene–V–Cl(2) 113.55(7), Ccarbene–V–N(1) 103.36(10), Cl(1)–V–Cl(2) 114.92(3), and V–N(1)–C(1) 165.31(18). Selected bond lengths (Å) for 5: V–Ccarbene 2.076(3), V–N(1) 1.654(3), V–Cl(1) 2.1620(10), V–Cl(2) 2.1559(9), and N(1)–C(1) 1.386(5). Selected bond angles (deg) for 5: Ccarbene–V–Cl(1) 106.53(8), Ccarbene–V–Cl(2) 117.61(8), Ccarbene–V–N(1) 103.66(12), Cl(1)–V–Cl(2) 114.68(4), and V–N(1)–C(1) 169.1(2). Detailed data are shown in the Supporting Information (see footnote a).

These complexes showed a distorted tetrahedral geometry around vanadium (Figure 2); the V–Ccarbene bond distances [2.060(2) and 2.076(3) Å for 4 and 5, respectively] are longer than those in complexes 13 [2.039(3)–2.049(2) Å]47 but are shorter than those in the other complexes such as [VOCl3(NHC)]67 and [V(CHSiMe3)(NAd)(CH2SiMe3)(NHC)]68 [V–Ccarbene = 2.137 and 2.172(2) Å, respectively]. The results thus indicate that the WCA-NHC-Ar′ ligands in 4 and 5 also act as a σ-donor toward vanadium. It seems that the degree of the σ-donation especially in 5 would be rather weak compared to that in 13, as also seen from the chemical shift in the 51V NMR spectra (described above). The V–Cl bond distances in 4 and 5 [2.1559(9)–2.1655(7) Å] are within the range of those in 13 [2.1544(14)–2.1747(15) Å],47 but, as discussed previously for 13,47 the distances are shorter than those in the related dichloride complexes [2.1901(8)–2.2462(8) Å] containing monodentate anionic donor ligands (exemplified as complexes CE in Scheme 2).46,6971 The Cl(1)–V–Cl(2) bond angles in 4 and 5 [114.92(3) and 114.68(4)° for 4 and 5, respectively] are smaller than those in 13 [116.26(5)–117.74(6)°],47 probably due to a steric bulk on the diisopropylphenyl substituent on the NHC-B(C6F5)3 ligand. The V–N(1) distances in 4 [1.646(3) Å] and 5 [1.654(3) Å] are relatively close to those in 2 and 3 [1.634(3) and 1.648(2) Å for 2 and 3, respectively],47 whereas the V–N(1)–C(1) bond angle in 5 [169.1(2)°] is larger than that in 3 [164.14(17)°] and the angle in 4 [165.31(18)°] is smaller than that in 2 [170.8(3)°].47 Taking into account these analysis data, as discussed for 13,47 the WCA-NHC-Ar′ ligands play a role as σ donors in electron-deficient, cationic 12-electron complexes in 4 and 5.

Ethylene Polymerization by [V(NR′)Cl2(L)] in the Presence of Al Cocatalysts

Table 1 summarizes the results of ethylene polymerization conducted in toluene at 25 °C using [V(NR′)Cl2(WCA-NHC-Ar′)] [R′ = C6H5 (4), 2,6-Me2C6H3 (5)] in the presence of MAO (d-MAO white solid, prepared by removal of toluene and AlMe3 from TMAO, commercially available from Tosoh Finechem Co.)7274 or AliBu3 cocatalysts. The results obtained using [V(NR′)Cl2(WCA-NHC-Ar)] [R′ = Ad (1), C6H5 (2), and Ar (3)]47 are also shown for comparison.

Table 1. Ethylene Polymerization Catalyzed by Vanadium(V) Dichloro Complexes Containing WCA-NHC Ligands (1–5)–Al Cocatalysts (Ethylene 8 atm, 25 °C, 10 Min)a.
run V cat. (μmol) Al cocat. (μmol) Al/Vb yield/mg activityc TOFd/min Mne × 10–4 Mw/Mne
1f 1 (1.0) d-MAO (500) 500 6 35 21    
2e 2 (1.0) d-MAO (100) 100 8 48 29    
3f 3 (1.0) d-MAO (200) 200 264 1580 939 1.53 1.93
4f 3 (1.0) d-MAO (500) 500 137 824 490    
5 4 (1.0) d-MAO (100) 500 17 102 61    
6 5 (1.0) d-MAO (200) 200 895 4210 2500    
7 5 (0.2) d-MAO (100) 500 14 420 250    
8 5 (0.2) d-MAO (200) 1000 136 4080 2420    
9 5 (0.2) d-MAO (400) 2000 664 19 900 11 800    
10 5 (0.2) d-MAO (600) 3000 645 19 500 11 600 9.46 1.56
11g C (1.0) d-MAO (2500) 2500 488 2930 1740 175 1.64
12f 3 (0.2) AliBu3 (10.0) 50 365 11 000 6540 1.80 1.76
13f 3 (0.2) AliBu3 (20.0) 100 161 4830 2870    
14f,h 3 (0.2) Et2AlCl (500) 2500 65.8 1970 1170    
15 4 (1.0) AliBu3 (100) 100 95 570 339    
16 4 (1.0) AliBu3 (200) 200 99 594 353    
17 5 (0.2) AliBu3 (10.0) 50 13 390 232 5.11 1.42
18 5 (0.2) AliBu3 (20.0) 100 67 2010 1190    
19 5 (0.2) AliBu3 (40.0) 200 187 5610 3330    
20 5 (0.2) AliBu3 (200) 100 223 6690 3970    
21 5 (0.2) AliBu3 (400) 200 390 11 700 6950    
22 5 (0.1) AliBu3 (400) 400 263 15 780 9380    
23 5 (0.1) AliBu3 (600) 600 373 22 380 13 300    
24 5 (0.02) AliBu3 (400) 2000 157 47 100 28 000    
25 5 (0.02) AliBu3 (500) 2500 220 66 000 39 200 13.7 2.35
26 5 (0.02) AliBu3 (550) 2750 187 56 100 33 300    
27 5 (0.02) AliBu3 (600) 3000 59 17 700 10 500    
28g,h C (0.05) Me2AlCl (250) 5000 229 27 500 16 300 (898)i  
29g,h C (0.01) iBu2AlCl (250) 25 000 108 64 800 38 500 (1250)i  
a

Conditions: toluene 30 mL, d-MAO [used as a white solid prepared from TMAO (Tosoh Finechem Co., Ltd.) by removing toluene and AlMe3] or AliBu3.

b

Al/V molar ratio.

c

Activity in kg-PE/mol-V·h.

d

TOF (min–1) = (molar amount of ethylene reacted)/(mol-V)·(min).

e

GPC data in o-dichlorobenzene vs polystyrene standards.

f

Cited from ref (47).

g

Result by V(NAr)Cl2(OAr) cited from refs (43) (run 11) or45(runs 28, 29).

h

Conducted at 0 °C.

i

Molecular weight by viscosity.45

It should be noted that the 5–MAO catalyst exhibited 10 times higher catalytic activity than the 3–MAO catalyst [e.g., activity: 1580 kg-PE/mol-V·h (by 3, run 3)47 vs 19 500 kg-PE/mol-V·h (by 5, run 10)]. The catalyst afforded high-molecular-weight polymers with unimodal molecular weight distributions, suggesting the formation of uniform catalytically active species in the solution. The activity by 5 was affected by the amount of the Al cocatalyst employed (runs 7–10), and the observed trend was somewhat different from that observed in 3 (runs 3, 4). Importantly, the activity by 5 (19 500 kg-PE/mol-V·h, run 10) is higher than those reported by [V(NAr)Cl2(OAr)] (C, 2930 kg-PE/mol-V·h, run 11),44 [V(NAr)Cl2{1,3-Ar2(CHN)2C=N}] (507 kg-PE/mol-V·h),46 [V(NAr)Cl2{1,3-Ar2(CH2N)2C=N}] (627 kg-PE/mol-V·h),46 and [V(NAd)Cl2{N=CtBu2}] (516 kg-PE/mol-V·h).68 The results thus demonstrate that 5 should be the most active (imido)vanadium(V) complex catalyst in the presence of the MAO cocatalyst, although the Mn values in the resultant polymers were lower than those by C (Table 1). It also turned out that the activity by the phenylimido analogue (4)–MAO catalyst was low (run 5), as observed in 2.

It was revealed that the 5–AliBu3 catalyst exhibited higher activities than the 5–MAO catalyst under the similar conditions [e.g., activity: 19 500 kg-PE/mol-V·h (run 10) vs 66 000 kg-PE/mol-V·h (run 25)]. Note that the 5–AliBu3 catalyst showed higher activity than the 3–AliBu3 catalyst [e.g., activity: 11 000 kg-PE/mol-V·h (by 3, run 12)47 vs 66 000 kg-PE/mol-V·h (by 5, run 25)], affording high-molecular-weight polymers with unimodal molecular weight distributions. The activity by 5 was affected by Al/V molar ratio (runs 17–27), and the observed trend was somewhat different from that observed in 3 (runs 12, 13).47 It also turned out that the activities by the 4–AliBu3 catalyst were low (runs 15, 16), as seen in the presence of MAO.

The activity by the 5–AliBu3 catalyst (66 000 kg-PE/mol-V·h, run 25) was higher than that by the [V(NAr)Cl2(OAr)]–Me2AlCl catalyst (C, 27 500 kg-PE/mol-V·h, run 28)45 but was comparable to that by the CiBu2AlCl catalyst (64800 kg-PE/mol-V·h, run 29), which afforded ultrahigh-molecular-weight polymers under these conditions.45 However, these polymerizations by C were conducted at 0 °C, and the activity decreased significantly at 25 °C.43 A similar effect of retardations (by increasing the temperature from 0 to 25 °C) was also observed in the (arylimido)vanadium(V) complexes containing iminoimidazolide or iminoimidazolidide ligands.46 As described in the Introduction section, AliBu3 has been considered as an ineffective Al cocatalyst in the metal-catalyzed olefin polymerization,117,2426,28,3946 but the use of AliBu3 should be favored in terms of practical viewpoints, especially for the establishment of a halogen-free process (without using halogenated Al) and for using an inexpensive Al cocatalyst compared to MAO.

Solution V K-Edge XANES and EXAFS Analyses of Reactions of (Arylimido)vanadium(V) Dichloride Complexes Containing Phenoxide and WCA-NHC Ligands with Al Alkyls. Effects of Anionic Ancillary Donor Ligands and Al Cocatalysts

As described in the Introduction section, metal-alkyl species play an important role in olefin polymerization/dimerization. It was demonstrated that cationic vanadium(V) alkyl species play a role in the ethylene dimerization using [V(NAd)X2(2-ArNCH2C5H4N)] (A, X = Cl, Me; Ar = 2,6-Me2C6H3)–MAO catalysts, whereas another cationic vanadium(V) alkyl species plays a role in the ethylene polymerization using A–Me2AlCl or Et2AlCl catalysts.40,48 The different catalyst behaviors in the reactions with ethylene (in the presence of MAO or Me2AlCl cocatalysts) could be thus explained as due to the catalyst/cocatalyst nuclearity effect (the formation of isolated or associated cationic alkyl species due to different cation/anion interactions).10,12,40,56,75,76 In contrast, as also described in the Introduction section (Figure 1), certain vanadium(III) species were formed when [V(NAd)Cl2(L)] [B, L = 2-(2′-benzimidazolyl)-6-methylpyridine] was treated with Me2AlCl (Figure 1a)42 or [V(NAr)Cl2(WCA-NHC-Ar)] (3) was treated with AliBu3 (Figure 1b).58,59 The high oxidation states were preserved when complex B or 3 was treated with MAO, suggesting that vanadium(V) species play a role in the ethylene polymerization.42,48 Since the XANES spectra, especially in the edge region (shoulder-edge), are apparently different between B–Me2AlCl and 3–AliBu3 catalysts (Figure 1), we thus herein explore solution XANES and EXAFS analyses of complexes 3 and 5 to obtain more information concerning the active species.

Figure 3 shows V K-edge XANES spectra of [V(NAr)Cl2(WCA-NHC-Ar)] (3) and [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar′ = 2,6-iPr2C6H3) in the presence of MAO or AliBu3 (in toluene at 25 °C, 50 μmol-V/mL). As reported previously,58 the XANES spectrum of 3 shows sharp pre-edge peak(s), which are typically observed in the complex with the tetrahedral geometry, at 5466.6 eV (and 5465.1 eV) ascribed to a transition from 1s to 3d + 4p.55,56,7779 A shoulder-edge peak, ascribed to an absorption of the V–Cl bond,55,56 was also observed at 5478.1 eV. The diisopropylphenyl analogue (5) also shows the pre-edge peak(s) at 5466.9 eV (and 5465.1 eV) and the shoulder-edge peak at 5478.0 eV. It turned out that the pre-edge peaks in 5 were shifted slightly at 5464.5 eV with a decrease in the intensity of the shoulder-edge peak by addition of MAO (10 equiv, Figure 3), suggesting that the oxidation state and the basic framework (geometry) were preserved. In contrast, as observed in 3 (Figures 1a, 3), significant changes in the XANES spectrum (pre-edge and edge peaks) were observed when 5 was treated with AliBu3 (100 equiv, Figure 3). An apparent decrease in the pre-edge intensity (observed as a small shoulder at 5465.1 eV) upon addition of AliBu3 clearly suggests a structural change, and the large low-energy shift in the edge absorption, which is very close to that observed in the 3–AliBu3 system, strongly suggests the formation of vanadium(III) species by reduction.b

Figure 3.

Figure 3

V K-edge XANES spectra of [V(NAr)Cl2(WCA-NHC-Ar)] (3, Ar = 2,6-Me2C6H3) and [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar′ = 2,6-iPr2C6H3) in the presence of MAO or AliBu3 (in toluene at 25 °C).

Figure 4a shows V K-edge XANES spectra (in toluene at 25 °C) of [V(N-2,6-R21C6H3)Cl2(O-2,6-R22C6H3)] [R1, R2 = Me, Me (6a);43 Me, Ph (6b);44 and iPr, iPr (6c)43] and [V(NAr)(OAr)3] (7).80 The dichloride complexes containing phenoxide ligands (6ac) have been chosen in this study because 6a (corresponds to C in Scheme 2) showed the highest activities for ethylene polymerization4345 and ethylene copolymerization with norbornene (NBE)44,45 in the presence of Al cocatalysts. It was demonstrated that the activities and the NBE incorporation in the (co)polymerization were highly affected by the Al cocatalyst employed (MAO vs Me2AlCl or Et2AlCl);44,45 the 6a,b–Et2AlCl catalyst showed the higher catalytic activities but less NBE incorporation than the 6a,b–MAO catalyst in the ethylene/NBE copolymerization.44

Figure 4.

Figure 4

V K-edge XANES spectra (in toluene at 25 °C) of (a) [V(N-2,6-R21C6H3)Cl2(O-2,6-R22C6H3)] [R1, R2 = Me, Me (6a); Me, Ph (6b); and iPr, iPr (6c)], [V(NAr)(OAr)3] (7, Ar = 2,6-Me2C6H3), and 6a,b in the presence of MAO (10 equiv), (b) 6a in the presence of Me2AlCl or Et2AlCl (50 equiv) and norbornene (NBE, 50 equiv), (c) 6a,c in the presence of Me2AlCl (50 equiv), and (d) [V(NAr)Cl2(WCA-NHC-Ar)] (3) or [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar′ = 2,6-iPr2C6H3) in the presence of AliBu3 (100 equiv).

As reported previously (for 6a,b),59 the XANES spectra of 6ac showed pre-edge peaks (and shoulder peaks) [5466.8 (and 5465.3) eV (6a), 5467.4 (and 5465.2) eV (6b), and 5466.9 (and 5465.7) eV (6c), respectively], which are similar to those in 3, 5, and [V(NAr)Cl3] [5466.8 (and 5465.0) eV]48 and shoulder-edge peaks (marked in the dashed circle in Figure 4a; 5478.2, 5478.2, and 5478.9 eV, respectively). The tris(phenoxide) analogue (7) showed only a sharp pre-edge peak at 5467.1 eV, clearly suggesting that the shoulder-edge peaks in 6ac are ascribed to an absorption of the V–Cl bond.55,56 The pre-edge peak positions in the spectra of 6a,b did not change (5466.8 and 5467.0 eV, respectively) upon addition of MAO (10 equiv) with an increase in the intensity, and the edge absorptions shifted slightly (Figures 4a and S2–1).c The results thus suggest that the oxidation state and the basic framework were preserved even upon addition of MAO; the facts also suggest that cationic alkyl species with vanadium(V) play a role in this catalysis, as demonstrated previously.42,48,55,56,58

Note that notable changes in the edge peaks (shoulder-edge at 5475.7 eV) were observed when 6a was treated with Me2AlCl or Et2AlCl (50 equiv, Figure 4b), and the spectra did not change upon the further addition of NBE (50 equiv). It seems that the pre-edge peak in 6a [5466.8 (and 5465.3) eV] shifted to low energy or became one absorption band upon addition of Me2AlCl (5465.7 eV) or Et2AlCl (5465.9 eV). A similar XANES spectrum was observed when 6c was treated with Me2AlCl (50 equiv, Figure 4c), and these edge peaks (absorptions) were observed in a low-photon-energy region compared to that in the (imido)vanadium(IV) complex. The results thus suggest that 6a,c were reduced by Me2AlCl to afford certain vanadium(III) species, as suggested by the reaction of 3 and 5 with AliBu3 (as shown in Figures 1b and 3).

It should also be noted that the observed XANES spectra (in both intensities in the pre-edge peaks and edge peaks) of 6a with addition of Me2AlCl or Et2AlCl are apparently different from those of 3 or 5 with AliBu3. The spectra (6a–Me2AlCl), especially in the pre-edge intensities, are also different from those of [V(NAd)Cl2(L)] (B) containing 2-(2′-benzimidazolyl)-6-methylpyridine ligands with addition of Me2AlCl (Figure 1a);42 the pre-edge intensity by B–Me2AlCl further decreased upon addition of ETA (Cl3CCO2Et), whereas the intensity decreased slightly by the reaction of 6a with Me2AlCl.42 It has been known that pre-edge intensities are affected by the basic geometry around vanadium; a compound in Td symmetry generally shows much higher pre-edge peak intensity than that in Oh symmetry due to a difference in the possibility of a p–d orbital hybridization.77 These results thus suggest that different vanadium(III) species (geometry and ligands) play roles in the ethylene (co)polymerization (see footnote b).

Moreover, as shown in Figure S2-2 (see footnote c), the shoulder-edge intensity (at 5475.7 eV) decreased when ETA (50 equiv) was added into a toluene solution of 6a, Me2AlCl (50 equiv), and NBE (50 equiv), whereas the intensity (at 5475.7 eV) increased with a decrease in the pre-edge intensity (at 5465.5 eV) when a toluene solution of B and Me2AlCl was added with ETA.42 The observed facts also correspond to the facts that the activity in ethylene polymerization by 6a decreased upon addition of ETA,45 whereas the activity by B increased upon addition of ETA,42 as seen in most vanadium complex and classical Ziegler-type vanadium catalyst systems (as described in the Introduction section).8,16,17,2328 These also suggest that different catalytically active vanadium(III) species would play roles in these catalyses.

Figure 5 shows V K-edge EXAFS oscillations and FT-EXAFS spectra (in toluene at 25 °C) of 3 with addition of AliBu3 (100 equiv) and 6a with addition of MAO (10 equiv) or Me2AlCl (50 equiv) (see footnote c). Tables 2 and 3 summarize the analysis results for observed neighboring atoms and bond distances around vanadium (see footnote c). The result for B with addition of Me2AlCl is also shown for comparison.42

Figure 5.

Figure 5

V K-edge (a, c) EXAFS oscillations and (b, d) FT-EXAFS spectra (in toluene at 25 °C) for reactions of (a, b) [V(NAr)Cl2(WCA-NHC-Ar)] (3, Ar = 2,6-Me2C6H3) in the presence of AliBu3 (100 equiv) and (c, d) [V(NAr)Cl2(OAr)] (6a) in the presence of MAO (10 equiv) and Me2AlCl (50 equiv). Additional spectra including fitting curves are shown in the Supporting Information (see footnote c).

Table 2. Summary of Data for [V(NAr)Cl2(WCA-NHC-Ar)] (3) in the Presence of AliBu3 (100 Equiv)a.

  complex 3 3 + AliBu3 (100 equiv)
atom C.N. r (Å) C.N. r (Å)
N(C) 2.1 ± 0.2 1.6 2± 0.03 0.8 ± 0.3 1.66 ± 0.17
Cl 1.0 ± 0.2 2.16 ± 0.04 1.0 ± 0.2 2.34 ± 0.04
Cl 1.0 ± 0.2 2.34 ± 0.05    
a

Atom: neighbor atom, C.N.: coordination number, r: bond length.

Table 3. Summary of Data for the Complex [V(NAr)Cl2(OAr)] (6a) or [V(NAd)Cl2(L)] [B, L = 2-(2′-Benzimidazolyl)-6-methylpyridine] in the Presence of Me2AlCl or MAOa.

  complex 6a 6a + MAO (10 equiv) 6a + Me2AlCl (50 equiv) complex Bb B + Me2AlCl (10 equiv)b
atom C.N. r (Å) C.N. r (Å) C.N. r (Å) C.N. r (Å) C.N. r (Å)
N(O) 2.4 ± 0.3 1.80 ± 0.05 1.8 ± 0.2 1.73 ± 0.04 1.3 ± 0.2 1.64 ± 0.04 1.7(2) 1.683(5) 0.9(3) 1.64(2)
N             1.2(8) 2.290(42)    
Cl 1.9 ± 0.2 2.1 8± 0.03 1.4 ± 0.2 2.17 ± 0.03 2.0 ± 0.2 2.45 ± 0.03 1.6(2) 2.293(3) 2.6(1) 2.455(7)
a

Atom: neighbor atom, C.N.: coordination number, r: bond length.

b

Data cited from ref (42).

As summarized in Table 2, the observed V–N distance by the EXAFS analysis (1.62 ± 0.03 Å) in 3 is close to that determined by X-ray crystallography [1.654(3) Å], and one of the observed V–Cl bond distances by the EXAFS analysis (2.16 ± 0.04 Å) is close to those determined by X-ray crystallography [2.1559(9) and 2.1620(10) Å]. Another V–Cl bond distance (2.34± 0.05 Å) was much longer than those in the related dichloride complexes [2.1901(8)–2.2462(8) Å, by X-ray crystallography]46,6971 containing monodentate anionic donor ligands (such as phenoxides, ketimides, iminoimidazolides, etc.). However, the V–Ccarbene bond [2.076(3) Å] was not observed or may be overlapped with the V–N bond (coordination number, 2.1 ± 0.2, Table 2). This could be assumed as due to the flexibility of the anionic WCA-NHC ligand in solution or explained by a simple speculation that the anionic NHC ligand binds to vanadium strongly in solution (the short V–Ccarbene bond distance compared to that in the solid state).

It should be noted that both EXAFS oscillation and the FT-EXAFS spectrum of 3 changed drastically by treatment with AliBu3 (100 equiv, Figure 5a,b), as observed in the XANES spectra (Figures 1b and 3). It turned out that the imido ligand (V–N bond, 1.66 ± 0.17 Å) was preserved, and one V–Cl bond (2.34 ± 0.04 Å), which is longer than those in the reported (imido)vanadium complexes with anionic donor ligands (shown above),46,6971,81 was also observed. However, the V–Calkyl bond, which should be present in catalysis, could not be defined in this analysis. Since the formed species are considered as vanadium(III) species on the basis of 51V NMR and ESR spectra, and the XANES spectra in solution,58 it is thus suggested that the Cl atom would be coordinated to vanadium (containing imido and alkyl ligands) as a neutral donor ligand probably bridged through Al.

It should also be noted that the imido ligand (V–N bond, 1.64 ± 0.04 Å, Table 3) was preserved after treatment of 6a with Me2AlCl. It also turned out that two V–Cl bonds were observed as the coordinating atoms to vanadium, and the long V–Cl bond distances (2.45 ± 0.03 Å) compared to those in 6a (2.18 ± 0.03 Å) would suggest that these Cl ligands coordinated to vanadium (containing imido and alkyl ligands) as neutral (or weak anionic) donor ligands (bridged through Al).81 In contrast, no significant changes in both the EXAFS oscillation and the FT-EXAFS spectrum were observed when 6a was treated with MAO, as observed in the XANES spectra; no significant differences in the both V–N and V–Cl bond distances were observed in the EXAFS analysis, except a decrease in the coordination number of Cl by treatment with MAO. These results thus suggest that the basic framework including the oxidation state was preserved by treatment of 6a with MAO.

Taking into account these analysis results, one V–Cl bond (2.34 ± 0.04 Å) was observed upon treatment of 3 with AliBu3 and two V–Cl bonds (2.45 ± 0.03 Å) were observed in the reaction of 6a with Me2AlCl; three V–Cl bonds [2.455(7) Å] were observed in the reaction of [V(NAd)Cl2(L)] [B, L = 2-(2′-benzimidazolyl)-6-methylpyridine] with Me2AlCl.42 In all cases, vanadium(III) species containing the imido V–N bond were formed on the basis of XANES and EXAFS analyses. Although we could not define the V–Calkyl bond, which should play a role in the ethylene (co)polymerization, the observed differences strongly suggest that three different active vanadium(III) species were formed depending upon the anionic donor and Al cocatalyst employed.

Concluding Remarks

(Arylimido)vanadium(V) dichloride complexes that contain anionic N-heterocyclic carbenes (NHCs) coordinated with a weakly coordinating B(C6F5)3 moiety, [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar = 2,6-Me2C6H3, Ar′ = 2,6-iPr2C6H3), exhibited remarkable catalytic activity for ethylene polymerization in the presence of Al cocatalysts (MAO and AliBu3). The activity by the 5–MAO catalyst (19 500 kg-PE/mol-V·h) was much higher than that by the dimethylphenyl analogue, [V(NAr)Cl2(WCA-NHC-Ar)] (3) and those reported using the reported (imido)vanadium(V) complexes containing monodentate anionic ligands.44,46,67 The 5–AliBu3 catalyst showed higher activity (66 000 kg-PE/mol-V·h) than 5–MAO and 3–AliBu3 catalysts, affording rather high-molecular-weight polymers with unimodal molecular weight distributions (Mn = 1.37 × 105, Mw/Mn = 2.35).

The solution V K-edge XANES analyses of 3 and 5 strongly suggest the formation of vanadium(III) species by reduction with AliBu3, accompanying structural changes (geometry) around vanadium (supported by changes in the pre-edge peak and decrease in the intensity) (see footnote b). The solution EXAFS analysis suggests the formation of vanadium species containing arylimido ligands and one V–Cl bond (2.34 ± 0.04 Å), which is longer than those in the reported (imido)vanadium(V) dichloride complexes that contain monodentate anionic ligands [2.1901(8)–2.2462(8) Å, by X-ray crystallography];46,6971the V–Calkyl bond, which should be present in catalysis, could not be defined in this analysis.

Moreover, the solution XANES analysis of [V(NAr)Cl2(OAr)] (6a) also suggests the formation of vanadium(III) species by reduction with Me2AlCl or Et2AlCl (supported by changes in the edge peaks). The observed XANES spectra especially in the edge region (shoulder-edge at 5475.7 eV) are different from those observed in the toluene solution containing 3 or 5 and AliBu3. The solution EXAFS analysis suggests the formation of vanadium species containing arylimido ligands and two V–Cl bonds (2.45 ± 0.03 Å), although the presence of the V–Calkyl bond could not be defined.

In contrast, no significant changes in both XANES and EXAFS spectra were observed by treatment of these complexes (3, 6a) with MAO, suggesting the preservation of the basic framework (a distorted 4-coordinate tetrahedral geometry) and the oxidation state; these results thus suggest that cationic (arylimido)vanadium(V) species play a role in the ethylene polymerization in the presence of the MAO cocatalyst. Results concerning the effect of the Al cocatalyst in ethylene polymerization and XANES analysis of 3, 5, and 6a are summarized in Scheme 4.

Scheme 4. Summary of the Effect of the Al Cocatalyst and the Anionic Donor Ligand in Ethylene Polymerization Using [V(NAr)Cl2(WCA-NHC-Ar)] (3, Ar = 2,6-Me2C6H3), [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar′ = 2,6-iPr2C6H3), and [V(NAr)Cl2(OAr)] (6a), and Reaction Chemistry through Solution XANES Analysis.

Scheme 4

Results for [V(NAd)Cl2(L)] [B, L = 2-(2′-benzimidazolyl)-6-methylpyridine]42 are also shown for comparison.

The facts observed for 3, 5, and 6a by treatment with AliBu or Me2AlCl demonstrate a unique contrast to the fact in the EXAFS analysis that more than two (three) V–Cl bonds [2.455(7) Å] were observed when [V(NAd)Cl2(L)] [B, L = 2-(2′-benzimidazolyl)-6-methylpyridine] was treated with Me2AlCl, which showed remarkable activity for ethylene polymerization.42 In addition to the observed difference in the XANES spectra as well as the catalyst behavior (the effect of Cl3CCO2Et in 6a and B in ethylene polymerization, Scheme 4), it should be considered that three different vanadium(III) species play roles in the ethylene polymerization depending upon the anionic donor and the Al cocatalyst employed. Moreover, we can at least say that these vanadium(III) species could be stabilized by the coordination of neutral donor ligands exemplified as Cl.

As described in the Introduction section,42,55,56,58,59 reported examples of analysis of (NMR and ESR-silent) vanadium(III) species as well as the study of homogeneous catalysis using solution XANES and EXAFS spectra are limited. As far as we know, this should be the first demonstration of the formation of different vanadium(III) species generated by reactions of vanadium complexes with Al alkyls through XAS analysis. We highly believe that the information should be potentially important for better understanding the study of the catalysis mechanism and organo-vanadium chemistry. Moreover, these data also demonstrate that the solution XAS analysis should be a powerful method for the obtainment of information concerning the catalytically active species in solution. More detailed analysis including the observation of V–Calkyl and V–CNHC bonds will be expected by the development of the analysis method (or measurements at low temperature), and the simulation of the active species on the basis of EXAFS spectra might also be possible; we thus expect further progress in this project.

Experimental Section

General Procedure

All experiments were conducted under a nitrogen atmosphere in a vacuum atmospheres dry box. Toluene, dichloromethane, and n-hexane of anhydrous grades (Kanto Kagaku Co., Ltd.) were transferred into a bottle containing molecular sieves (a mixture of 3A 1/16, 4A 1/8, and 13 × 1/16) in the dry box; solvents were passed through an alumina short column prior to use. V(NC6H5)Cl339 and V(N-2,6-Me2C6H3)Cl382 were prepared according to a published method. WCA-NHC-Ar′ (NHC-Ar′ = 1,3-bis(2,6-diisopropyl)-imidazolin-2-ylidene] and Li(WCA-NHC-Ar′)(toluene) were prepared according to the reported procedure.64,83 AliBu3, Me2AlCl, and Et2AlCl (Kanto Kagaku Co., Ltd) and ethylene (polymerization grade, purity >99.9%, Sumitomo Seika Co. Ltd.) were used as received. Solid MAO (d-MAO) samples were prepared by the removal of toluene and AlMe3 from the commercially available TMAO [9.5 wt % (Al) toluene solution, Tosoh Finechem Co.] in the dry box under reduced pressure (at ca. 50 °C for removing toluene, AlMe3, and then heated at >100 °C for 1 h for the completion).7274

All 1H, 13C, 19F, and 51V NMR spectra were recorded on a Bruker AV500 spectrometer (500.13 MHz for 1H, 125.77 MHz for 13C, 470.59 MHz for 19F, and 131.55 MHz for 51V). All spectra were recorded in the solvent indicated at 25 °C unless otherwise noted. Chemical shifts are given in ppm and are referenced to SiMe4 (δ 0.00 ppm, 1H, 13C), CFCl3 (δ 0.00 ppm, 19F), and VOCl3 (δ 0.00 ppm, 51V). Coupling constants and half-width values, Δν1/2, are given in hertz. Elemental analyses were performed by an EAI CE-440 CHN/O/S elemental analyzer (Exeter Analytical, Inc.).

Synthesis of [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, Ar′ = 2,6-iPr2C6H3)

The basic synthetic procedure of 4 was analogous to that of [V(NC6H5)Cl2(WCA-NHC-Ar)] (2, Ar = 2,6-Me2C6H3),47 except that Li(WCA-NHC-Ar′) was used in place of Li(WCA-NHC-Ar). Into a toluene solution (40 mL) containing V(NC6H5)Cl3 (177 mg, 0.713 mmol), Li(WAC-NHC-Ar′)(toluene) (715 mg, 0.716 mmol) was added at −30 °C. The reaction mixture was warmed slowly to room temperature, and the mixture was continuously stirred at room temperature overnight. The resultant mixture was passed through a Celite pad; the filter cake was washed with toluene. After the removal of volatiles from the combined filtrate and the wash, the resultant solid was dissolved in a minimum amount of dichloromethane. The chilled solution layered with n-hexane afforded red microcrystals in the freezer (−30 °C, 83 mg, 74.6 μmol, yield: 10%). 1H NMR (500 MHz, C6D6 at 25 °C): δ 7.19 (t, J = 7.9 Hz, 1H), 7.03 (d, J = 7.8 Hz, 2H), 6.80 (s, 1H), 6.61 (d, J = 7.6 Hz, 2H), 6.51–6.44 (m, 6H), 2.95 (sep, J = 6.6 Hz, 2H), 2.52 (sep, J = 6.5 Hz, 2H), 1.25 (d, J = 6.6 Hz, 6H), 1.17 (d, J = 6.4 Hz, 6H), 0.91 (d, J = 5.7 Hz, 12H). 13C NMR (125 MHz, C6D6 at 25 °C): δ 149.6 (dm, 1JC–F = 233 Hz, C6F5), 147.7 (s, N–C–N), 146.4 (br, CH=C–B), 140.0 (dm, 1JC–F = 255 Hz, C6F5), 137.6 (dm, 1JC–F = 249 Hz, C6F5), 137.3 (s, Ar), 133.0 (s, Ar), 132.1 (s, Ar), 131.3 (s, CH=C–B), 130.9 (s, Ar), 129.4 (br, ipso-C6F5), 127.5 (s, Ar), 127.1 (s, Ar), 125.1 (s, Ar), 124.4 (br, Ar), 29.6 (s), 28.4 (s), 27.8 (s), 26.9 (s), 22.8 (br), 21.2 (s). 19F NMR (470 MHz, C6D6 at 25 °C): δ −158.7, −164.7. 51V NMR (131 MHz, C6D6 at 25 °C): δ 256.6 (Δν1/2 = 1121.0 Hz). Anal. calcd for C51H40BCl2F15N3V: C, 55.06; H, 3.62; N, 3.78; found: C, 53.18; H, 3.81; N, 3.51. The structure of 4 could be determined by X-ray crystallographic analysis (CCDC 1949138).

Synthesis of [V(NAr)Cl2(WCA-NHC-Ar′)] (5, Ar = 2,6-Me2C6H3)

The basic synthetic procedure of 5 was analogous to that of [V(NAr)Cl2(WCA-NHC-Ar)] (3),47 except that Li(WCA-NHC-Ar′) was used in place of Li(WCA-NHC-Ar). Into a toluene solution (40 mL) containing V(NAr)Cl3 (200 mg, 0.723 mmol), Li(WAC-NHC-Ar′)(toluene) (722 mg, 0.723 mmol) was added at −30 °C. The reaction mixture was continuously stirred at room temperature overnight. The resultant mixture was passed through a Celite pad; the filter cake was washed with toluene. After the removal of volatiles from the combined filtrate and the wash, the resultant solid was dissolved in a minimum amount of dichloromethane. The chilled solution layered with n-hexane afforded red microcrystals in the freezer (−30 °C, 116 mg, 102 μmol, yield: 14%). Yield 116 mg (0.102 mmol, 14%). 1H NMR (500 MHz, C6D6 at 25 °C): δ 7.22 (t, J =7.8 Hz,1H), 6.96 (t, J = 7.5 Hz, 3H), 6.74 (s, 1H), 6.71 (d, J = 7.7 Hz, 2H), 6.32 (t, J = 7.6 Hz, 1H), 6.22 (d, J = 7.6 Hz, 2H), 3.27 (sep, J = 6.2 Hz, 2H), 2.66 (sep, J = 6.5 Hz, 2H), 1.92 (m, 6H), 1.30 (s, 6H), 1.02 (s, 6H), 0.95 (d, J = 6.5, 6H), 0.76 (s, 6H). 13C NMR (125 MHz, C6D6 at 25 °C): δ 149.6 (dm, 1JC–F = 239 Hz, C6F5), 146.9 (s, N–C–N), 146.5 (br, CH=C–B), 142.5 (s, Ar), 139.8 (dm, 1JC–F = 245 Hz, C6F5), 137.6 (dm, 1JC–F = 253 Hz, C6F5), 135.0 (s, Ar), 134.8 (s, Ar), 132.4 (s, Ar), 132.4 (br, CH=C–B), 132.2 (s, Ar), 131.8 (s, Ar), 127.6 (s, Ar), 125.6 (s, ipso-C6F5), 124.8 (s, Ar), 28.9 (s), 28.6 (s), 26.8 (s), 23.5 (br), 21.4 (s), 18.8 (s). 51V NMR (131 MHz, C6D6 at 25 °C): δ 410.5 (Δν1/2 = 1450.0 Hz). Anal. calcd for C53H44BCl2F15N3V (+0.6 × n-hexane): C, 56.97; H, 4.40; N, 3.52; found: C, 57.14; H, 4.54; N, 3.46. The structure of 5 could be determined by X-ray crystallographic analysis (CCDC 1949093).

Typical Ethylene Polymerization Procedure

Into a stainless-steel autoclave (100 mL scale), toluene and the Al cocatalyst (total 29 mL) were added in the dry box. A toluene solution (1.0 mL) containing a prescribed amount of the catalyst (1.0 or 0.2 μmol/mL) was then added into the autoclave filled with ethylene (1 atm); the reactor was then pressurized immediately to 7 atm (total 8 atm), and the mixture was stirred magnetically for 10 min. After the reaction, the chilled mixture in the autoclave was then poured into MeOH containing HCl. The resultant polymer (white precipitate) was collected on a filter paper by filtration and was adequately washed with MeOH. The resultant polymer was then dried in vacuo at 60 °C for 2 h.

Analysis of Catalyst Solution by Solution-Phase X-ray Absorption Spectroscopy (XAS)

V K-edge X-ray absorption near-edge structure (XANES) and X-ray absorption fine structure (XAFS) measurements were carried out at the BL01B1 beam line at the SPring-8 facility of the Japan Synchrotron Radiation Research Institute (proposal nos. 2016A1455, 2016B1509, 2017A1512, 2018A1245, and 2018B1335). V K-edge XAFS spectra of V complex samples [toluene solution, 50 μmol/mL, at 25 °C; a Si (111) two-crystal monochromator was used for the incident beam] were recorded in the fluorescence mode using an ionization chamber as the I0 detector and 19 solid-state detectors as the I detector. The X-ray energy was calibrated using V2O5, and the data analysis was performed with the REX2000 Ver. 2.5.9 software package (Rigaku Co.). The XANES data was analyzed using a cubic spline from the χ spectra with the removal of the atomic absorption background and their normalization to the edge height.

Crystallographic Analysis

The measurements for [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, CCDC 1949138) and [V(N-2,6-Me2C6H3)Cl2(WCA-NHC-Ar′)] (5, CCDC 1949093) were performed using a Rigaku XtaLAB P200 diffractometer with multilayer mirror monochromated Mo Kα radiation. The crystal collection parameters are given in the Supporting Information (see footnote a). The data were collected at 180 °C and processed using CrystalClear (Rigaku)84 or CrysAlisPro (Rigaku Oxford Diffraction),85 and the structure was solved by direct methods86 and expanded using Fourier techniques. The nonhydrogen atoms were refined anisotropically, and some hydrogen atoms were refined using the riding model. All calculations were performed using the Crystal Structure87 crystallographic software package, except for refinement, which was performed using SHELXL Version 2014/7.88,89 Cif and xyz files are shown in the Supporting Information, and the crystallographic data was deposited at the Cambridge Crystallographic Data Centre (CCDC 1949093 and 1949138).

Acknowledgments

This project was partly supported by the Grant-in-Aid for Scientific Research on Innovative Areas (“3D Active-Site Science”, No. 26105003) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan; the Grant-in-Aid for Scientific Research from the Japan Society for the Promotion of Science (JSPS, nos. 15H03812 and 18H01982); and the Fund for the Promotion of Joint International Research (Fostering Joint International Research, 19KK0139). The authors express their thanks to Profs. S. Komiya and A. Inagaki (Tokyo Metropolitan University, TMU) for discussion; to Tosoh Finechem Co. for donating MAO; and to Dr. K. Tsutsumi, Dr. A. Igarashi, T. Omiya, M. Kuboki, H. Harakawa, K. Inoue, K. Kawamura, and Y. Kawamoto (TMU) for technical assistance with the XAS analysis at the BL01B1 beam line at the SPring-8 facility of Japan Synchrotron Radiation Research Institute (JASRI, proposal nos. 2016A1455, 2016B1509, 2017A1512, 2018A1245, and 2018B1335).

Supporting Information Available

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b02828.

  • Selected NMR spectra of [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, Ar′ = 2,6-iPr2C6H3) and [V(N-2,6-Me2C6H3)Cl2(WCA-NHC-Ar′)] (5); additional XANES spectra and EXAFS and FT-EXAFS spectra (in toluene at 25 °C) including fitting curves and the analysis data of [V(N-2,6-Me2C6H3)Cl2(WCA-NHC)] (3) and [V(NAr)Cl2(OAr)] (6a) (PDF)

  • NMR spectra table for crystal data and collection parameters in structural analyses for [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, Ar’ = 2,6-iPr2C6H3,CCDC 1949138) and [V(N-2,6-Me2C6H3)Cl2(WCA-NHC-Ar′)] (5, CCDC 1949093) (PDF), and the CIF files

  • Collection parameter structural analyses (XYZ)

The authors declare no competing financial interest.

Footnotes

a

NMR spectra and the table for crystal data and collection parameter structural analyses of [V(NC6H5)Cl2(WCA-NHC-Ar′)] (4, Ar′ = 2,6-iPr2C6H3, CCDC 1949138) and [V(N-2,6-Me2C6H3)Cl2(WCA-NHC-Ar′)] (5, CCDC 1949093) including cif and xyz files are shown in the Supporting Information.

b

The reviewer commented that reduction of [V(NAr)Cl2(WCA-NHC-Ar′)] (5) or [V(NAr)Cl2(WCA-NHC-Ar)] (3) occurred by the reaction with AliBu3 via β-hydrogen elimination, and this is the most probable route. However, the mechanism for the formation of (arylimido)vanadium(III) species is unclear at this moment.

c

Additional XANES spectra and EXAFS and FT-EXAFS spectra (in toluene at 25 °C) including fitting curves and the analysis data of [V(N-2,6-Me2C6H3)Cl2(WCA-NHC)] (3) and [V(NAr)Cl2(OAr)] (6a) are shown in the Supporting Information.

Supplementary Material

ao9b02828_si_001.pdf (1.6MB, pdf)
ao9b02828_si_002.cif (3.4MB, cif)
ao9b02828_si_003.xyz (11.4KB, xyz)

References

  1. Brintzinger H. H.; Fischer D.; Mülhaupt R.; Rieger B.; Waymouth R. M. A tailor-made metallocene for the copolymerization of ethene with bulky cycloalkenes. Angew. Chem., Int. Ed. 1995, 34, 1143–1170. Selected early reviews in early transition metal catalysis (refs Figures 1–6) 10.1002/anie.199511431. [DOI] [Google Scholar]
  2. Kaminsky W.; Arndt M. Metallocenes for polymer catalysis. Adv. Polym. Sci. 1997, 127, 144–187. 10.1007/BFb0103631. [DOI] [Google Scholar]
  3. McKnight A. L.; Waymouth R. M. Group 4 ansa-cyclopentadienyl-amido catalysts for olefin polymerization. Chem. Rev. 1998, 98, 2587–2598. 10.1021/cr940442r. [DOI] [PubMed] [Google Scholar]
  4. Britovsek G. J. P.; Gibson V. C.; Wass D. F. The search for new-generation olefin polymerization catalysts: Life beyond metallocenes. Angew. Chem., Int. Ed. 1999, 38, 428–447. . [DOI] [PubMed] [Google Scholar]
  5. Coates G. W.; Hustad P. D.; Reinartz S. Catalysts for the living insertion polymerization of alkenes: access to new polyolefin architectures using Ziegler-Natta chemistry. Angew. Chem., Int. Ed. 2002, 41, 2236–2257. . [DOI] [PubMed] [Google Scholar]
  6. Gibson V. C.; Spitzmesser S. K. Advances in non-metallocene olefin polymerization catalysis. Chem. Rev. 2003, 103, 283–316. 10.1021/cr980461r. [DOI] [PubMed] [Google Scholar]
  7. Selected Recent Reviews/Accounts in Early Transition Metal Catalysis (refs 7–14), see: Organometallic Reactions and Polymerization; The Lecture Notes in Chemistry 85;Osakada K., Ed.; Springer-Verlag: Berlin, 2014. [Google Scholar]
  8. Nomura K.; Zhang S. Design of vanadium complex catalysts for precise olefin polymerization. Chem. Rev. 2011, 111, 2342–2362. 10.1021/cr100207h. [DOI] [PubMed] [Google Scholar]
  9. Makio H.; Terao H.; Iwashita A.; Fujita T. FI Catalysts for olefin polymerization—A comprehensive treatment. Chem. Rev. 2011, 111, 2363–2449. 10.1021/cr100294r. [DOI] [PubMed] [Google Scholar]
  10. Delferro M.; Marks T. J. Multinuclear olefin polymerization catalysts. Chem. Rev. 2011, 111, 2450–2485. 10.1021/cr1003634. [DOI] [PubMed] [Google Scholar]
  11. McInnis J. P.; Delferro M.; Marks T. J. Multinuclear group 4 catalysis: Olefin polymerization pathways modified by strong metal–metal cooperative effects. Acc. Chem. Res. 2014, 47, 2545–2557. 10.1021/ar5001633. [DOI] [PubMed] [Google Scholar]
  12. Valente A.; Mortreux A.; Visseaux M.; Zinck P. Coordinative chain transfer polymerization. Chem. Rev. 2013, 113, 3836–3857. 10.1021/cr300289z. [DOI] [PubMed] [Google Scholar]
  13. Nomura K.; Liu J.; Padmanabhan S.; Kitiyanan B. Nonbridged half-metallocenes containing anionic ancillary donor ligands: New promising candidates as catalysts for precise olefin polymerization. J. Mol. Catal. A: Chem. 2007, 267, 1–28. 10.1016/j.molcata.2006.11.006. [DOI] [Google Scholar]
  14. Nomura K.; Liu J. Half-titanocenes for precise olefin polymerisation: Effects of ligand substituents and some mechanistic aspects. Dalton Trans. 2011, 40, 7666–7682. 10.1039/c1dt10086f. [DOI] [PubMed] [Google Scholar]
  15. Baier M. C.; Zuideveld M. A.; Mecking S. Post-metallocenes in the industrial production of polyolefins. Angew. Chem., Int. Ed. 2014, 53, 9722–9744. 10.1002/anie.201400799. [DOI] [PubMed] [Google Scholar]
  16. Nomura K.; Zhang W.. Organometallic Reactions and Polymerization. In The Lecture Notes in Chemistry 85; Osakada K., Ed.; Springer-Verlag: Berlin, 2014; pp 89–118. [Google Scholar]
  17. Nomura K.; Hou X.. Handbook of Transition Metal Polymerization Catalysts, 2nd ed.; Hoff R., Ed.; Wiley: Hoboken, NJ, 2018; pp 313–338. [Google Scholar]
  18. Carrick W. L. Mechanism of ethylene polymerization with vanadium catalysts. J. Am. Chem. Soc. 1958, 80, 6455–6456. 10.1021/ja01556a073. [DOI] [Google Scholar]
  19. Carrick W. L.; Kluiber R. W.; Bonner E. F.; Wartman L. H.; Rugg F. M.; Smyth J. J. Transition metal catalysts. I. Ethylene polymerization with a soluble catalyst formed from an aluminum halide, tetraphenyltin and a vanadium Halide. J. Am. Chem. Soc. 1960, 82, 3883–3887. 10.1021/ja01500a025. [DOI] [Google Scholar]
  20. Phillips G. W.; Carrick W. L. Transition metal catalysts VIII. The role of oxygen in ethylene polymerizations with the AlBr3-VXn-Sn(C6H5)4 catalyst. J. Polym. Sci. 1962, 59, 401–412. 10.1002/pol.1962.1205916816. [DOI] [Google Scholar]
  21. Junghanns Von E.; Gumboldt O.; Bier G. Polymerisation von äthylen und propylen zu amorphen copolymerisaten mit katalysatoren aus vanadiumoxychlorid und aluminiumhalogenalkylen. Makromol. Chem. 1962, 58, 18–42. 10.1002/macp.1962.020580102. [DOI] [Google Scholar]
  22. Natta G.; Mazzanti G.; Valvassori A.; Sartori G.; Fiumani D. Ethylene–propylene copolymerization in the presence of catalysts prepared from vanadium triacetylacetonate. J. Polym. Sci. 1961, 51, 411–427. 10.1002/pol.1961.1205115603. [DOI] [Google Scholar]
  23. Christman D. L.; Keim G. I. Reactivities of nonconjugated dienes used in preparation of terpolymers in homogeneous systems. Macromolecules 1968, 1, 358–363. 10.1021/ma60004a017. [DOI] [Google Scholar]
  24. Hagen H.; Boersma J.; van Koten G. Homogeneous vanadium-based catalysts for the Ziegler–Natta polymerization of α-olefins. Chem. Soc. Rev. 2002, 31, 357–364. 10.1039/B205238E. [DOI] [PubMed] [Google Scholar]
  25. Gambarotta S. Vanadium-based Ziegler–Natta: Challenges, promises, problems. Coord. Chem. Rev. 2003, 237, 229–243. 10.1016/S0010-8545(02)00298-9. [DOI] [Google Scholar]
  26. Nomura K.New Developments in Catalysis Research; NOVA Science Publishers: New York, 2005; pp 199–217. [Google Scholar]
  27. Christman D. L. Preparation of polyethylene in solution. J. Polym. Sci., Part A-1: Polym. Chem. 1972, 10, 471. 10.1002/pol.1972.150100213. [DOI] [Google Scholar]
  28. Redshaw C. Vanadium procatalysts bearing chelating aryloxides: structure–activity trends in ethylene polymerization. Dalton Trans. 2010, 39, 5595–5604. 10.1039/b924088h. [DOI] [PubMed] [Google Scholar]
  29. Natta G.; Zambelli A.; Lanzi G.; Pasquon I.; Mognaschi E. R.; Segre A. L.; Centola P. Polymerization of propylene to syndiotactic polymer. Part. I: Valence of active vanadium in the catalytic systems. Makromol. Chem. 1965, 81, 161–172. 10.1002/macp.1965.020810118. [DOI] [Google Scholar]
  30. Zambelli A.; Pasquon I.; Signorini R.; Natta G. Polymerization of propylene to syndiotactic polymer. III. Behaviour of the catalyst system VCl4–Al(C2H5)2Cl in the presence of lewis bases. Makromol. Chem. 1968, 112, 160–182. 10.1002/macp.1968.021120115. [DOI] [Google Scholar]
  31. Lehr M. H. The active oxidation state of vanadium in soluble monoolefin polymerization catalysts. Macromolecules 1968, 1, 178–184. 10.1021/ma60002a014. [DOI] [Google Scholar]
  32. Lehr M. H.; Carman C. J. Electron spin resonance evidence of inactive V(III) precursor to catalytically active V(III) in vanadium tetrachloride Ziegler catalysts. Macromolecules 1969, 2, 217–219. 10.1021/ma60008a025. [DOI] [Google Scholar]; ESR study for monitoring the catalytically active species (VCl4–Et2AlCl catalyst system).
  33. Talsi E.; Bryliakov K.. Application of EPR and NMR Spectroscopy in Homogeneous Catalysis; CRC Press, Taylor & Francis: Boca Raton, FL, 2017. [Google Scholar]
  34. Goswami M.; Chirila A.; Rebreyend C.; de Bruin B. EPR Spectroscopy as a Tool in Homogeneous Catalysis Research. Top. Catal. 2015, 58, 719–750. 10.1007/s11244-015-0414-9. [DOI] [Google Scholar]
  35. Bigmore Helen R.; Zuideveld M. A.; Kowalczyk R. M.; Cowley A. R.; Kranenburg M.; McInnes E. J. L.; Mountford P. Synthesis, structures, and olefin polymerization capability of vanadium(4+) imido compounds with fac-N3 donor ligands. Inorg. Chem. 2006, 45, 6411–6423. 10.1021/ic060454i. [DOI] [PubMed] [Google Scholar]
  36. Soshnikov I. E.; Semikolenova N. V.; Shubin A. A.; Bryliakov K. P.; Zakharov V. A.; Redshaw C.; Talsi E. P. EPR Monitoring of vanadium(IV) species formed upon activation of vanadium(V) polyphenolate precatalysts with AlR2Cl and AlR2Cl/ethyltrichloroacetate (R = Me, Et). Organometallics 2009, 28, 6714–6720. 10.1021/om900515h. [DOI] [Google Scholar]
  37. Soshnikov I. E.; Semikolenova N. V.; Bryliakov K. P.; Zakharov V. A.; Redshaw C.; Talsi E. P. An EPR study of the vanadium species formed upon interaction of vanadyl N and C-capped tris(phenolate) complexes with AlEt3 and AlEt2Cl. J. Mol. Catal. A: Chem. 2009, 303, 23–29. 10.1016/j.molcata.2008.12.013. [DOI] [Google Scholar]
  38. Soshnikov I. E.; Semikolenova N. V.; Bryliakov K. P.; Shubin A. A.; Zakharov V. A.; Redshaw C.; Talsi E. P. An EPR study of the V(IV) species formed upon activation of a vanadyl phenoxyimine polymerization catalyst with AlR3 and AlR2Cl (R = Me, Et). Macromol. Chem. Phys. 2009, 210, 542–548. 10.1002/macp.200800556. [DOI] [Google Scholar]
  39. Zhang S.; Nomura K. Highly efficient dimerization of ethylene by (imido)vanadium complexes containing (2-anilidomethyl)pyridine ligands: Notable ligand effect toward activity and selectivity. J. Am. Chem. Soc. 2010, 132, 4960–4965. 10.1021/ja100573d. [DOI] [PubMed] [Google Scholar]
  40. Igarashi A.; Zhang S.; Nomura K. Ethylene dimerization/polymerization catalyzed by (adamantylimido)vanadium(V) complexes containing (2-anilidomethyl)pyridine ligands: Factors affecting the ethylene reactivity. Organometallics 2012, 31, 3575–3581. 10.1021/om3000532. [DOI] [Google Scholar]
  41. Tang X.-Y.; Igarashi A.; Sun W.-H.; Inagaki A.; Liu J.; Zhang W.; Li Y.-S.; Nomura K. Synthesis of (imido)vanadium(V) complexes containing 8-(2,6-dimethylanilide)-5,6,7-trihydroquinoline ligands: Highly active catalyst precursors for ethylene dimerization. Organometallics 2014, 33, 1053–1060. 10.1021/om401119y. [DOI] [Google Scholar]
  42. Nomura K.; Oshima M.; Mitsudome T.; Harakawa H.; Hao P.; Tsutsumi K.; Nagai G.; Ina T.; Takaya H.; Sun W.-H.; Yamazoe S. Synthesis, structural analysis of (imido)vanadium dichloride complexes containing 2-(2′-benz-imidazolyl)pyridine ligands: Effect of Al cocatalyst for efficient ethylene (co)polymerization. ACS Omega 2017, 2, 8660–8673. 10.1021/acsomega.7b01225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Nomura K.; Sagara A.; Imanishi Y. Olefin polymerization and ring-opening metathesis polymerization of norbornene by (arylimido)(aryloxo)vanadium(V) complexes of the type VX2(NAr)(OAr‘). Remarkable effect of aluminum cocatalyst for the coordination and insertion and ring-opening metathesis polymerization. Macromolecules 2002, 35, 1583–1590. 10.1021/ma0117413. [DOI] [Google Scholar]
  44. Wang W.; Nomura K. Remarkable effects of aluminum cocatalyst and comonomer in ethylene copolymerizations catalyzed by (arylimido)(aryloxo)vanadium complexes: Efficient synthesis of high molecular weight ethylene/norbornene copolymer. Macromolecules 2005, 38, 5905–5913. 10.1021/ma050629s. [DOI] [Google Scholar]
  45. Wang W.; Nomura K. Notable effects of aluminum alkyls and solvents for highly efficient ethylene (co)polymerizations catalyzed by (arylimido)(aryloxo)vanadium complexes. Adv. Synth. Catal. 2006, 348, 743–750. 10.1002/adsc.200505446. [DOI] [Google Scholar]
  46. Nomura K.; Bahuleyan B. K.; Zhang S.; Sharma P. M. V.; Katao S.; Igarashi A.; Inagaki A.; Tamm M. Synthesis and structural analysis of (imido)vanadium(V) dichloride complexes containing iminoimidazolide and iminoimidazolidide ligands, and their use as catalyst precursors for ethylene (co)polymerization. Inorg. Chem. 2014, 53, 607–623. 10.1021/ic402747d. [DOI] [PubMed] [Google Scholar]
  47. Igarashi A.; Kolychev E. L.; Tamm M.; Nomura K. Synthesis of (imido)vanadium(V) dichloride complexes containing anionic N-heterocyclic carbenes that contain a weakly coordinating borate moiety: new MAO-free ethylene polymerization catalysts. Organometallics 2016, 35, 1778–1784. 10.1021/acs.organomet.6b00200. [DOI] [Google Scholar]
  48. Nomura K.; Mitsudome T.; Igarashi A.; Nagai G.; Tsutsumi K.; Ina T.; Omiya T.; Takaya H.; Yamazoe S. Synthesis of (adamantylimido)vanadium(V) dimethyl complex containing (2-anilidomethyl)pyridine ligand and selected reactions: Exploring the oxidation state of the catalytically active species in ethylene dimerization. Organometallics 2017, 36, 530–542. 10.1021/acs.organomet.6b00727. [DOI] [Google Scholar]
  49. For examples (refs (4953)); Thomas J. M.; Sankar G. The role of XAFS in the in situ and ex situ elucidation of active sites in designed solid catalysts. J. Synchrotron Radiat. 2001, 8, 55–60. 10.1107/S090904950001935X. [DOI] [PubMed] [Google Scholar]
  50. Dent A. J. Development of time-resolved XAFS instrumentation for quick EXAFS and energy-dispersive EXAFS measurements on catalyst systems. Top. Catal. 2002, 18, 27–35. 10.1023/A:1013826015970. [DOI] [Google Scholar]
  51. Thomas J. M.; Catlow C. R. A.; Sankar G. Determining the structure of active sites, transition states and intermediates in heterogeneously catalysed reactions. Chem. Commun. 2002, 24, 2921–2925. 10.1039/b210679p. [DOI] [PubMed] [Google Scholar]
  52. Bare S. R.; Ressler T. Characterization of catalysts in reactive atmospheres by X–ray absorption spectroscopy. Adv. Catal. 2009, 52, 339–465. 10.1016/S0360-0564(08)00006-0. [DOI] [Google Scholar]
  53. Iwasawa Y.; Asakura K.; Tada M.. XAFS Techniques for Catalysts, Nanomaterials, and Surfaces; Springer: Switzerland, 2017. [Google Scholar]
  54. For feature articles (refs Figures 4–6); Linehan J. C.; Balasubramanian M.; Fulton J. L.. XAFS Techniques for Catalysts, Nanomaterials, and Surfaces; Springer: Switzerland, 2017; pp 431. [Google Scholar]
  55. Nomura K.; Mitsudome T.; Tsutsumi K.; Yamazoe S. Solution XAS analysis for exploring the active species in homogeneous vanadium complex catalysis. J. Phys. Soc. Jpn. 2018, 87, 061014 10.7566/JPSJ.87.061014. [DOI] [Google Scholar]
  56. Nomura K.; Mitsudome T.; Yamazoe S. Direct observation of catalytically active species in reaction solution by X-ray absorption spectroscopy (XAS). Jpn. J. Appl. Phys. 2019, 58, 100502 10.7567/1347-4065/ab3e5c. [DOI] [Google Scholar]
  57. For reported examples in homogeneous early transition metal complex catalysis (refs (42), (48), (57)−60); Bartlett S. A.; Moulin J.; Tromp M.; Reid G.; Dent A. J.; Cibin G.; McGuinness D. S.; Evans J. Activation of [CrCl3{R-SN(H)S-R}] catalysts for selective trimerization of ethene: A freeze-quench Cr K-edge XAFS study. ACS Catal. 2014, 4, 4201–4204. 10.1021/cs501017g. [DOI] [Google Scholar]
  58. Nagai G.; Mitsudome T.; Tsutsumi K.; Sueki S.; Ina T.; Tamm M.; Nomura K. Effect of Al cocatalyst in ethylene and ethylene/norbornene (co)polymerization by (imido)vanadium bichloride complexes containing anionic N-heterocyclic carbenes having weakly coordinating borate moiety. J. Jpn. Pet. Inst. 2017, 60, 256–262. 10.1627/jpi.60.256. [DOI] [Google Scholar]
  59. Nomura K.; Tsutsumi K.; Nagai G.; Omiya T.; Ina T.; Yamazoe S.; Mitsudome T. Solution XAS analysis of various (imido)vanadium(V) dichloride complexes containing monodentate anionic ancillary donor ligands: effect of aluminium cocatalyst in ethylene/norbornene (co)polymerization. J. Jpn. Pet. Inst. 2018, 61, 282–287. 10.1627/jpi.61.282. [DOI] [Google Scholar]
  60. Kuboki M.; Nomura K. (Arylimido)niobium(V) complexes containing 2-pyridylmethylanilido ligand as catalyst precursors for ethylene dimerization that proceeds via cationic Nb(V) species. Organometallics 2019, 38, 154–1559. 10.1021/acs.organomet.9b00017. [DOI] [Google Scholar]
  61. General recent reviews in frustrated Lewis pairs (refs Figures 13):; Stephan D. W. Frustrated Lewis pairs: From concept to catalysis. Acc. Chem. Res. 2015, 48, 306–316. 10.1021/ar500375j. [DOI] [PubMed] [Google Scholar]
  62. Stephan D. W.; Erker G. Frustrated Lewis pair chemistry: Development and perspectives. Angew. Chem., Int. Ed. 2015, 54, 6400–6441. 10.1002/anie.201409800. [DOI] [PubMed] [Google Scholar]
  63. Nasr A.; Winkler A.; Tamm M. Anionic N-heterocyclic carbenes: Synthesis, coordination chemistry and applications in homogeneous catalysis. Coord. Rev. 2016, 316, 68–124. 10.1016/j.ccr.2016.02.011. [DOI] [Google Scholar]
  64. Kolychev E. L.; Kronig S.; Brandhorst K.; Freytag M.; Jones P. G.; Tamm M. Iridium(I) complexes with anionic N-heterocyclic carbene ligands as catalysts for the hydrogenation of alkenes in nonpolar media. J. Am. Chem. Soc. 2013, 135, 12448–12459. 10.1021/ja406529c. [DOI] [PubMed] [Google Scholar]
  65. Winkler A.; Brandhorst K.; Freytag M.; Jones P. G.; Tamm M. Palladium(II) complexes with anionic N-heterocyclic carbene–borate ligands as catalysts for the amination of aryl halides. Organometallics 2016, 35, 1160–1169. 10.1021/acs.organomet.6b00217. [DOI] [Google Scholar]
  66. Nomura K.; Nagai G.; Nasr A.; Tsutsumi K.; Kawamoto Y.; Koide K.; Tamm M. Synthesis of half-titanocenes containing anionic N-heterocyclic carbenes that contain a weakly coordinating borate moiety (WCA-NHC), Cp’TiX2(WCA-NHC), and their use as catalysts for ethylene (co)polymerization. Organometallics 2019, 38, 3233–3244. 10.1021/acs.organomet.8b00841. [DOI] [Google Scholar]
  67. Abernethy C. D.; Codd G. M.; Spicer M. D.; Taylor M. K. A highly stable N-heterocyclic carbene complex of trichloro-oxo-vanadium(V) displaying novel Cl–Ccarbene bonding interactions. J. Am. Chem. Soc. 2003, 125, 1128–1129. 10.1021/ja0276321. [DOI] [PubMed] [Google Scholar]
  68. Zhang W.; Nomura K. Facile synthesis of (imido)vanadium(V)–alkyl, alkylidene complexes containing an N-heterocyclic carbene ligand from their trialkyl analogues. Organometallics 2008, 27, 6400–6402. 10.1021/om8007822. [DOI] [Google Scholar]
  69. Yamada J.; Fujiki M.; Nomura K. A vanadium(V) alkylidene complex exhibiting remarkable catalytic activity for ring-opening metathesis polymerization (ROMP). Organometallics 2005, 24, 2248–2250. 10.1021/om0501834. [DOI] [Google Scholar]
  70. Zhang W.; Nomura K. Synthesis of (1-adamantylimido)vanadium(V) complexes containing aryloxo, ketimide ligands: Effect of ligand substituents in olefin insertion/metathesis polymerization. Inorg. Chem. 2008, 47, 6482–6492. 10.1021/ic800347n. [DOI] [PubMed] [Google Scholar]
  71. Zhang S.; Tamm M.; Nomura K. 1,2-C–H Activation of benzene promoted by (arylimido)vanadium(V)-alkylidene complexes: Isolation of the alkylidene, benzyne complexes. Organometallics 2011, 30, 2712–2720. 10.1021/om200086n. [DOI] [Google Scholar]
  72. For example (Figures 24); Nomura K.; Naga N.; Miki M.; Yanagi K. Olefin polymerization by (cyclopentadienyl)(aryloxy)titanium(IV) complexes–cocatalyst systems. Macromolecules 1998, 31, 7588–7597. 10.1021/ma980690f. [DOI] [Google Scholar]
  73. Nomura K.; Fujita K.; Fujiki M. Olefin polymerization by (cyclopentadienyl)(ketimide)-titanium(IV) complexes of the type, Cp’TiCl2(N=CtBu2), – methylaluminoxane (MAO) catalyst systems. J. Mol. Catal. A: Chem. 2004, 220, 133–144. 10.1016/j.molcata.2004.06.010. [DOI] [Google Scholar]
  74. Byun D.-J.; Fudo A.; Tanaka A.; Fujiki M.; Nomura K. Effect of cyclopentadienyl and anionic ancillary ligand in syndiospecific styrene polymerization catalyzed by nonbridged half-titanocenes containing aryloxo, amide and anilide ligands–cocatalyst systems. Macromolecules 2004, 37, 5520–5530. 10.1021/ma049549z. [DOI] [Google Scholar]
  75. For example (refs (10), (11) and (75), (76)); Macchioni A. Ion pairing in transition-metal organometallic chemistry. Chem. Rev. 2005, 105, 2039–2073. 10.1021/cr0300439. [DOI] [PubMed] [Google Scholar]
  76. Bochmann M. The Chemistry of catalyst activation: The case of group 4 polymerization catalyst. Organometallics 2010, 29, 4711–4740. 10.1021/om1004447. [DOI] [Google Scholar]
  77. Yamamoto T. Assignment of pre-edge peaks in K-edge x-ray absorption spectra of 3d transition metal compounds: electric dipole or quadrupole?. X-Ray Spectrom. 2008, 37, 572–584. 10.1002/xrs.1103. [DOI] [Google Scholar]
  78. Srivastava U. C.; Higam H. L. X-ray absorption edge spectrometry (xaes) as applied to coordination chemistry. Coord. Chem. Rev. 1973, 9, 275–310. 10.1016/S0010-8545(00)82080-9. [DOI] [Google Scholar]; References cited therein.
  79. Wong J.; Lytle F. W.; Messmer R. P.; Maylotte D. H. K-edge absorption spectra of selected vanadium compounds. Phys. Rev. B 1984, 30, 5596–5610. 10.1103/PhysRevB.30.5596. [DOI] [Google Scholar]; Reports for solid structural analysis of vanadium compounds by XAS spectra.
  80. Yamada J.; Fujiki M.; Nomura K. Synthesisof various (arylimido)vanadium(V)-methyl complexes containing ketimide ligands and reactions with alcohols, thiols, borates: Implications for unique reactivity toward alcohols. Organometallics 2007, 26, 2579–2588. 10.1021/om061121w. [DOI] [Google Scholar]
  81. Estimated value (2.33 Å) for vanadium-chloride single bond according to Pauling’s covalent radii,Pauling L.The Nature of the Chemical Bond, 3rd ed.; Cornell University Press: Ithaca, NY, 1960. [Google Scholar]
  82. Buijink J.-K. F.; Teubin J. H.; Kooijman H.; Spek A. L. Synthesis, molecular structure, and reactivity of a half-sandwich vanadium(III) imido complex: The first vanadium(V) alkylidene. Organometallics 1994, 13, 2922–2924. 10.1021/om00020a003. [DOI] [Google Scholar]
  83. Kronig S.; Theuergarten E.; Daniliuc C. G.; Jones P. G.; Tamm M. Anionic N-heterocyclic carbenes that contain a weakly coordinating borate moiety. Angew. Chem., Int. Ed. 2012, 51, 3240–3244. 10.1002/anie.201108813. [DOI] [PubMed] [Google Scholar]
  84. CrystalClear: Data Collection and Processing Software, Rigaku Corporation; (1998–2015). Tokyo 196-8666, Japan, 2015. [Google Scholar]
  85. CrysAlisPro: Data Collection and Processing Software, Rigaku Corporation; (2015). Tokyo 196-8666, Japan: 2015. [Google Scholar]
  86. Sheldrick G. M. SHELXT: Integrating space group determination and structure solution. Acta Crystallogr., Sect. A: Found. Adv. 2014, 70, C1437. 10.1107/S2053273314085623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. CrystalStructure 4.2: Crystal Structure Analysis Package; Rigaku Corporation: Tokyo 196-8666, Japan, 2000-2015. [Google Scholar]
  88. Sheldrick G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112–122. 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
  89. 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]

Associated Data

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

Supplementary Materials

ao9b02828_si_001.pdf (1.6MB, pdf)
ao9b02828_si_002.cif (3.4MB, cif)
ao9b02828_si_003.xyz (11.4KB, xyz)

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

RESOURCES