Next Article in Journal
Neuroprotective Effects of an Aqueous Extract of Forsythia viridissima and Its Major Constituents on Oxaliplatin-Induced Peripheral Neuropathy
Next Article in Special Issue
Propylene Polymerization Catalyzed by Metallocene/Methylaluminoxane Systems on Rice Husk Ash
Previous Article in Journal
Design and Synthesis of New Quinoxaline Derivatives as Anticancer Agents and Apoptotic Inducers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Linear Polyethylenes Achieved Using Thermo-Stable and Efficient Cobalt Precatalysts Bearing Carbocyclic-Fused NNN-Pincer Ligand

1
Beijing Key Laboratory of Clothing Materials R&D and Assessment, Beijing Engineering Research Center of Textile Nanofiber, School of Materials Science and Engineering, Beijing institute of Fashion Technology, Beijing 100029, China
2
Key Laboratory of Engineering Plastics and Beijing National Laboratory for Molecular Science, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
3
N.N. Vorozhtsov Novosibirsk Institute of Organic Chemistry, Pr. Lavrentjeva 9, Novosibirsk 630090, Russia
4
College of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310014, China
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(6), 1176; https://doi.org/10.3390/molecules24061176
Submission received: 13 March 2019 / Revised: 20 March 2019 / Accepted: 21 March 2019 / Published: 25 March 2019
(This article belongs to the Special Issue Well-Defined Metal Complex Catalysts for Olefin Polymerization)

Abstract

:
Six examples of 2-(1-arylimino)ethyl-9-arylimino-5,6,7,8-tetrahydrocycloheptapyridine-cobalt(II) chloride complexes, [2-(1-ArN)C2H3-9-ArN-5,6,7,8-C5H8C5H3N]CoCl2, (Ar = 2-(C5H9)-6-MeC6H3 Co1, 2-(C6H11)-6-MeC6H3 Co2, 2-(C8H15)-6-MeC6H3 Co3, 2-(C5H9)-4,6-Me2C6H2 Co4, 2-(C6H11)-4,6-Me2C6H2 Co5, and 2-(C8H15)-4,6-Me2C6H2 Co6), were synthesized by the direct reaction of the corresponding ortho-cycloalkyl substituted carbocyclic-fused bis(arylimino)pyridines (L1L6) and cobalt(II) chloride in ethanol with good yields. All the synthesized ligands (L1L6) and their corresponding cobalt complexes (Co1Co6) were fully characterized by FT-IR, 1H/13C-NMR spectroscopy and elemental analysis. The crystal structure of Co2 and Co3 revealed that the ring puckering of both the ortho-cyclohexyl/cyclooctyl substituents and the one pyridine-fused seven-membered ring; a square-based pyramidal geometry is conferred around the metal center. On treatment with either methylaluminoxane (MAO) or modified methylaluminoxane (MMAO), all the six complexes showed high activities (up to 4.09 × 106 g of PE mol−1 (Co) h−1) toward ethylene polymerization at temperatures between 20 °C and 70 °C with the catalytic activities correlating with the type of ortho-cycloalkyl substituent: Cyclopentyl (Co1 and Co4) > cyclohexyl (Co2 and Co5) > cyclooctyl (Co3 and Co6) for either R = H or Me and afforded strictly linear polyethylene (Tm > 130 °C). The narrow unimodal distributions of the resulting polymers are consistent with single-site active species for the precatalyst. Furthermore, compared to the previously reported cobalt analogues, the titled precatalysts exhibited good thermo-stability (up to 70 °C) and possessed longer lifetime along with a higher molecular weight of PE (Mw: 9.2~25.3 kg mol−1).

Graphical Abstract

1. Introduction

Over the past two decades ago, Brookhart [1,2] and Gibson [3,4,5] independently discovered the well-defined bis(imino)pyridine-cobalt and -iron complexes (A) as highly active precatalysts for ethylene oligo-/polymerization. In the intervening years, large numbers of different skeletal modifications, such as changes to the N-aryl groups and the backbone of pyridine, have been made to the bis(imino)pyridine scaffolds (A, Chart 1) [6,7,8,9,10,11,12,13,14,15], culminating in improvements to both catalytic activity and thermal stability [6,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33]. The related progress was collected in several reviews [6,9,10,11,14,15].
To enhance the thermal stability, one efficient method was introduced to the bulky substituent on the ortho-position of N-aryl group [6,7,8,9,10,11,12,13,14,15] (Chart 1). Wu’s group have reported that bis(imino)pyridine iron complexes (A1) bearing a bulky 2-methyl-6-sec-phenethyl group, can produce linear high molecular weight polyethylene with higher activity at elevated temperature [31]. In the recent five years or so, our group have developed series of dibenzhydryl substituted bis(imino)pyridyl iron and cobalt complexes (A2) that displayed good catalytic activity at high temperature and explored a very good thermal stability [34]. There were also reports about incorporating the cycloalkyl substituents in the 2,6-positions of the N-aryl groups of the bis(imino)pyridine (A3) and it has been found to efficiently increase the temperature stability of the iron catalysts over their alkyl analogues [35]. Later Russia’s group documented the 2,6-bis(arylimino)pyridy liron(II) chloride complexes by the introduction of cycloaliphatic substituents into the ortho position of the aryl groups substantially enlarges the temperature interval of efficient use of the related catalysts for ethylene polymerization reactions [36]. Consequently, they have prepared the multifunctional bis(imino)pyridine iron chloride complexes incorporating the cycloalkyl substituent that could display more efficient ethylene polymerization at elevated temperatures (60–90 °C) while retaining high molecular weights of the resulting PE, as distinct from the analogous monofunctional systems (A3) [37].
On the other hand, the modification of bis(imino)pyridine backbone to enhance the catalytic properties at a higher temperature. Our group has been interested in introducing cycloalkyl unit to bis(imino)pyridine ligand sets by adjusting the flexibility of the exterior imine donors to modify the donor properties of the NNN-pincer ligand set and in-turn the performance of the catalyst [6]. With regards to cobalt precatalysts (Chart 1), one [38,39,40,41,42] or two [43,44,45,46,47,48] cycloalkyl unit fused derivatives having ring sizes between five and eight are accessible (see for example B [39], C [40], D [41], E [43,44], F [45], and G [47,48], Chart 1). In terms of catalytic performance and the polymeric products properties, we have observed significant differences between cobalt catalysts (BG). For example, five-membered ring fused bis(imino)pyridine cobalt precatalyst [38], showed lower activities (2.89 × 104 g of PE mol−1(Co) h−1) than their analogues of A and BG [6]; while using B that containing slightly larger six-membered ring (Chart 1) [39], results in much higher activities (up to 1.08 × 107 g of PE mol−1 (Co) h−1) and better thermal stability than A and C [4,6,39,40], and produces polyethylene wax with narrower molecular weight distribution. Expanding the ring size to seven (D and E, Chart 1) [41,43,44], the more flexible structure, leads to increased molecular weight (Mw up to 54.0 kg mol−1) which is most apparent with doubly-fused E containing ortho-cycloalkyl substituent, but E display slightly lower activity their counterparts D. Beyond this, it was found that cyclohexyl fused pyridine cobalt complexes (B and C) produced the polyethylene with very low molecular weight (0.82 kg mol−1) and linear saturated structure. The seven-membered ring fused pyridine(imino)cobalt precatalyst (D and E) produced the higher molecular weight polyethylene (3.2 kg mol−1) that contained the vinyl group, which has a significant potential application in the polywax in the package field [6]. The other interesting is the one cycloalkyl fused pyridine (imino) cobalt (B and D) showed better thermal stability (optimum temperature: 50~60 °C) and higher activity (8.65–10.09 × 106 g of PE mol−1(Co) h−1) than that by double cyclcoalkyl fused bis(imino)pyridine cobalt complexes (C, E, F, and G) (activity: 2.89–5.04 × 106 g of PE mol−1(Co) h−1 and optimum temperature at 30–40 °C). Very recently, we have reported a series of E derivatives that incorporate the cycloalkyl substitution on the ortho position of N-aryl group, which exhibited good activity (2.0 × 106 g of PE mol−1(Co) h−1) with the optimum temperature of 30 °C [43]. Considering the potential positive effect by the cycloalkyl substituent on the ortho position of N-aryl group and the backbone of monocycloheptyl fused pyridine(imino) on the thermal stability and polymerization activity toward ethylene polymerization, therefore, in this work we focused on the mono cycloheptyl-fused bis(imino)pyridine cobalt complexes that incorporated the different ring size of cycloalkyl on the ortho position of N-aryl group and their catalytic behavior toward ethylene polymerization in detail.

2. Results

2.1. Synthesis and Characterization of Ligands and Cobalt Complexes Co1Co6

According to a literature procedure [49], the condensation reactions of 2-acetyl-5,6,7,8-tetrahydrocycloheptapyridin-9-one with 2-cycloalkylaniline hydrochloride derivatives were conducted in the refluxing n-butanol to form the desired NNN-pincer ligands (L1/L1’L6/L6’, Chart 1) in moderate isolated yield (32%–39%, Scheme 1).
The 1H-NMR spectra of the ligands indicated two isomers for each because of the migration of double bond to the imino group linked to the cycloheptane, illustrating major compounds with the remaining imino group (L) instead of double bonds within cycloheptanic derivatives (L′, minor), which is similar to our previous work [49,50]. The isomers with N–H groups are confirmed with the observation range from 3372 to 3379 cm−1 in their IR spectra, while the isomers containing the C=N groups were observed characteristic band around 1644 cm−1. Then using the literature procedure [40,49], the corresponding NNN-pincer ligands (L1/L1’L6/L6) reacted with cobalt(II) dichloride in ethanol to give the corresponding 2-(1-arylimino)ethyl-9-arylimino-5,6,7,8-tetrahydrocyclohepta -pyridine-cobalt(II) chloride complexes, [2-(1-ArN)C2H3-9-ArN-5,6,7,8-C5H8C5H3N]CoCl2, (Ar = 2-(C5H9)-6-MeC6H3 Co1, 2-(C6H11)-6-MeC6H3 Co2, 2-(C8H15)-6-MeC6H3 Co3, 2-(C5H9)-4,6-Me2C6H2 Co4, 2-(C6H11)-4,6-Me2C6H2 Co5, and 2-(C8H15)-4,6-Me2C6H2 Co6), in high yields (91–96%). All the ligands L1L6 have been characterized by FT-IR, 1H/13C-NMR spectroscopy and elemental analysis. The cobalt complexes Co1Co6 have been characterized by IR and elemental analysis and the structure of Co2 and Co3 was confirmed by single-crystal X-ray diffraction. Comparing the IR spectra of free ligands with their corresponding Co complexes, there is no absorption related to the N–H groups; moreover, the adsorption around 1644 cm−1 for νC=N in ligand compounds is shifted to lower wavenumber around 1609 cm−1 within the cobalt complexes indicating effective coordination of Co–Nsp2 [40,41].

2.2. Single-Crystal X-ray Diffraction Study

The single-crystals of Co2 and Co3 suitable for the X-ray determinations were obtained by slow diffusion of n-hexane into their solution in dichloromethane at room temperature. Their molecular structures are shown in Figure 1a,b, and selected bond lengths and angles in Table 1.
Co2 and Co3 have a similar structure and each cobalt was surrounded by two chlorides and three nitrogen atoms belonging to the corresponding NNN-pincer ligand. Both of them possessed distorted square-pyramidal geometry, in which three nitrogen atoms N1, N2, N3 and Cl1 formed the base plane and Cl2 occupying the apical position. The cobalt atom lies at a distance of 0.505 Å above the basal plane for Co2 and 0.447 Å for Co3 in a manner similar to that observed in related NNNCoCl2 counterparts [40,41,45,47]. In these two structures, the bond length of Co-Npy [2.063(4) Å (Co2), 2.043(1) Å (Co3)] is significantly shorter than that of Co-Nimine [2.203(4) and 2.183(3) Å (Co2), 2.193(4) and 2.175(4)Å (Co3)], suggesting the stronger donor ability by N-py. At the same time, they are close to that in A (2.051(3) Å) [4], B (2.040(3) Å) [39], D (2.036(4)–2.052(5) Å) [40,41], but are shorter than that of E (2.087(4) Å) [44]. By contrast, the bond length of Co–Nimine [2.175(4)–2.203(4) Å] are longer than that in E (2.128(3) and 2.176(3) Å) [51] and α,α′-bis(arylimino)-2,3:5,6-bis(hexamethylene)pyridine-cobalt(II) complexes (range: 2.1180(18) - 2.1633(18) Å) [47], similar to those A [4], B [39], and D [40,41] (range: 2.193–2.320 Å, Chart 1), highlighting the good chelation properties of D to the cobalt center. The N-aryl rings are almost perpendicular to the coordination plane with dihedral angles of (74.23° and 78.48°) for Co2 and (83.84° and 71.39°) for Co3, respectively. Concurrently, the cyclo-substituents always located in trans-position. The torsion angles for N(1)–C(2)–C(3)–N(2) [2.09° (Co2), −7.08° (Co3)] and N(3)–C(11)–C(12)–N(2) [−10.57° (Co2), 13.35° (Co3)] highlight the deviation from co-planarity between the pyridine ring and the neighboring imine vectors, and the large distinct torsion angles are attributed to the unsymmetrical framework and the different sizes ortho-cycloalkyl substituent [44]. The interesting is the substituents on the ortho position of N-aryl group are always in trans-position.

2.3. Ethylene Polymerization

Based on previous findings for structurally related cobalt (II) chloride complexes (e.g., BE in Chart 1) [39,40,41,42,43,44,45,46,47], the highest catalytic activity for ethylene polymerization tends to be achieved upon activation with either methylaluminoxane (MAO) or modified methylaluminoxane (MMAO) [6,7,8,9,10,11,12,13,14,15]. Therefore, these two co-catalysts were employed to investigate the ethylene polymerization in detail as follows.

2.3.1. Ethylene Polymerization using Co2/MAO

With MAO as co-catalyst, Co2 was employed as precatalyst to investigate the ethylene polymerization systematically at various parameters, such as different Al/Co molar ratios, reaction temperatures and run times and polymerization results are collected in Table 2.
First of all, with the ethylene pressure set at 10 atm and the Al/Co ratio fixed at 1000, ethylene polymerizations were conducted at a different temperature from 30 to 70 °C (runs 1–5, Table 2). The maximum activity of 2.89 × 106 g of PE mol−1 (Co) h−1 was found at 50 °C affording a polymeric product; no trace of oligomers could be detected. Further, higher temperature led to the dramatically decreasing of the polymerization activity (Figure 2a). Meanwhile, the molecular weight (Mw) of the polymer was found to decrease from 22.97 kg mol−1 to 6.56 kg mol−1 when increasing the temperature from 30 °C to 70 °C, which can be ascribed to either the more facile chain transfer or chain termination with respect to the chain propagation yielding lower molecular weight PE [4,40,41,45,51,52,53]. Moreover, their GPC curves showed the unimodal distribution and the resultant polyethylene possessed narrow polydispersity (PDI = 2.0~3.8), which is consistent with the characters of single-site active species [6,39,40,41,42,43,44,45].
Fixing the temperature at 50 °C, ethylene polymerizations by Co2/MAO were conducted at a different Al/Co molar ratio from 500 to 3500 (runs 3 and 6–11, Table 2). The highest activity was observed as 3.57 × 106 g of PE mol−1(Co) h−1 with a molar ratio of 2500 (run 9, Table 2), a lower or higher ratio of co-catalyst led to lower activities. The molecular weight of the polymer (Mw) gradually decreased from 14.2 kg mol−1 to 9.39 kg mol−1, and the molecular weight distribution kept very narrow (PDI = 2.4 − 2.7). This result can be ascribed to increased chain transfer to aluminum occurring as a consequence of the larger amounts of alkyl aluminum reagent employed [4,40,41,44,51,52]. To our surprise with the Al/Co molar ratios changing between 3000 and 3500, a dramatic increase of molecular weight was observed and up to 13.96 kg mol1 with 3500 equivalents of MAO (Figure 2b). It would seem plausible that the formation of new active species has been influenced within this molar ratio window, a similar observation has been noted elsewhere [38,44]. On the other hand, with regards to the molecular weight distribution, the PDI of polymers are falling in the range from 2.4 to 3.0, slightly broader than their counterparts D containing symmetric para-R1 groups (PDI = 2.2) [41], this observation was similar to complex E incorporating α, α′-bis(arylimino)-2,3:5,6-bis(hexamethylene)pyridine containing cycloalkyl ortho-substituents [43,44].
To the end, the lifetime of the Co2/MAO catalyst was probed by conducting the polymerization over 5, 15, 30, 45 and 60 min (runs 9 and 12–15, Table 2). Polymerization activity gradually decreased from 8.12 × 106 g of PE mol−1(Co) h−1 to 2.48 × 106 g of PE mol−1 (Co) h−1 with prolonging the reaction time from, 5 to 60 min, which can be explained by the rapid formation of active species after the addition of MAO and gradual deactivation over extended reaction time [34,40,41,42,43,44,45]. It is indicated that a short time frame was required to generate all active species, and then the onset of partial deactivation of active species occurred over the course of the reaction [34]. Decreasing ethylene pressure to 5 atm, the polymerization activity (2.85 × 106 g of PE mol−1 (Co) h−1) is much lower than that (3.57 × 106 g of PE mol−1 (Co) h−1) at 10 atm (runs 9 and 16, Table 2), but the resultant PE polymer possessed much higher molecular weight (13.04 vs. 9.39 kg mol−1).
In order to explore the effect imparted by the cycloalkyl ortho-substitution pattern on performance and polymer properties, the other complexes were additionally screened using the optimized reaction conditions established independently for Co2/MAO and the results were collected in Table 2 (runs 17–21). On activation with MAO, all the cobalt complexes Co1Co6 displayed good activities in the range of 3.21–4.09 × 106 g of PE mol−1 (Co) h−1, which fall in the order: Co1 [R1 = cyclopentyl] > Co2 [R1 = cyclohexyl] > Co3 [R1 = cyclooctyl], Co4 [R1=cyclopentyl] > Co5 [R1 = cyclohexyl] > Co6 [R1 = cyclooctyl] (Figure 3a). These findings indicate that the size of cycloalkyl group on the ortho position of N-aryl group affects the catalytic activity, in which less bulky cyclopentyl systems gave the higher activities (Co1, Co4) and the bulkiest group of cyclooctyl showed the lowest (Co3, Co6) activities. All the resultant PE by different cobalt complexes Co1Co6 possessed narrow distribution and PDI value ranged from 2.4 to 3.9, a slightly broader than that found in the D/MAO system. The GPC traces clearly showed unimodal distribution (Figure 3). But the polymer by Co3 and Co6 that bear the bulkiest ortho-substituent of cyclooctyl possessed the highest molecular weight among their analogues, the similar trends were also observed for E/MAO system [44]. The para-methyl substituted complexes Co4Co6 showed slightly lower activity than unsubstituted ones Co1Co3, which was demonstrated by the activity order: Co1 (R2 = H) > Co4 (R2 = Me), Co2 (R2 = H) > Co5 (R2 = Me) and Co3 (R2 = H) > Co6 (R2 = Me). However, Figure 3 shows that molecular weights by Co1Co3 are lower than that by Co4Co6 respectively. These trends are similar to their analog complex of B [39].
In comparison with previously reported cobalt precatalysts, such as A, B and D (Chart 2) [4,39,40,41,44], the current systems, Co1Co6, under comparable polymerization conditions (namely MAO as co-catalyst, 10 atm C2H4, 30 min) exhibited relatively lower catalytic activity (3.21–4.09 × 106 g of PE mol−1 (Co) h−1), but are higher than that by precatalyst E (Chart 2). On the other hand, the resulting polymers by Co1Co5/MAO possessed much higher molecular weight than that found in D/MAO. These findings heights cycloalkyl ortho-substituents in the N-aryl group favored the high molecular weight polymer formation.

2.3.2. Ethylene Polymerization Using Co2/MMAO

Under the optimized molar ratio for Co2/MAO (Al/Co = 2500), ethylene polymerization by Co2/MMAO were conducted at various temperatures (from 30 °C to 60 °C) at 10 atm ethylene pressure (runs 1–4, Table 2). The results showed that the highest activity of 2.95 × 106 g of PE mol−1(Co) h−1 was also observed at 50 °C (run 3, Table 3), similar to the results by Co2/MAO (2.89 × 106 g of PE mol−1 (Co) h−1). Further, higher temperature led to the decrease of activity, indicating the partial decomposition of the active species [4,40,41,45,51,52,53]. The molecular weight of PE decreased from 12.98 kg mol−1 to 6.41kg mol−1 with the increasing the temperature from 30 °C to 60 °C, which can be ascribed to either increased chain transfer to aluminum or chain termination by β-H elimination at the higher temperature [47,48,49,50,51,52,53,54]. However, all the resultant PE possessed quite narrow distribution (PDI = 2.5) and GPC traces showed unimodal distribution, which is a typical character of a single site catalyst system. A similar observation was also reported for their counterparts B, D and E (Chart 2).
With the temperature fixed at 50 °C, the influence of the Al/Co molar ratio was investigated in the range of 1500 to 3000 (runs 3, 5–7, Table 3). A topmost of activity 2.95 × 106 g of PE mol−1(Co)h−1 was achieved at the molar ratio of 2500, which is similar to the results with MAO. As with MMAO, the resultant polyethylenes possessed high molecular weights (8.19–10.22 kg mol−1) and narrow polydispersities (Mw/Mn ≈ 2.5), which is consistent with single-site characteristics for the active species [40,41,42,43,44,45].
In order to investigate the lifetime of the active species, the polymerization tests were conducted over different reaction time from 5, 15, 30 and 60 min (runs 3 and 8–10, Table 3). The results showed that polymerization activity decreased from 7.04 to 2.07 × 106 g of PE mol−1 (Co) h−1 when increasing the time from 5 to 60 min. As with MMAO system, the molecular weights of the resultant polymer are falling in the range from 8.19 to 10.49 kg mol−1, and displaying narrow molecular weight distribution (Mw/Mn = 2.4–3.0). With the ethylene pressure reduced to 5 atm, the activity (1.90 × 106 g of PE mol−1 (Co) h−1) was much lower than that at 10 atm (run 8, Table 3).
With MMAO as co-catalyst, all the cobalt complexes Co1Co6 were evaluated for their ethylene polymerization and they also displayed good activity (2.24–3.34 × 106 g of PE mol−1 (Co) h−1). The similar trends in activity and molecular weight as MAO are observed, in which the cyclopentyl substituted Co1 and Co4 displayed the highest activity and cyclooctyl substituted cobalt complex (Co3 and Co6) gave the lowest activity, but produced the highest molecular weight polyethylene (runs 3, 12–16, Table 3). In comparison with D/MMAO, the ortho-cycloalkyl systems in general exhibited more than twice higher molecular weight, although the catalytic activity of D/MMAO is nearly twice as much as that seen for Co1 and Co4, and almost three times that for Co2Co3 and Co5Co6. In addition, the molecular weights obtained using Co1Co6 (range: 8.19–15.94 kg mol−1) are all significantly higher than that seen for D/MMAO system (2.5 kg mol−1) [40], which also illustrated the steric effect of the ortho cycloalkyl substituents suppressed the chain transfer and led to the high molecular weight of the polymer. As shown by their GPC curves in Figure 4, the MWD of polymeric products are falling in the range 2.4 to 4.4.

2.4. Characterization of Polyethylene

Regardless of MMAO or MAO activation, narrow unimodal distributions of the polymers are observed (Mw/Mn =2.2–4.5) consistent with single-site active species. All the Tm values of the polymer ranging from 129 °C to 133 °C [6,40,41,42,43,44,45]. To provide further support for the linearity of the polymers, representative PE samples using Co1/MAO at 50 °C (run 9, Table 2) were characterized by 1H/13C-NMR spectroscopy at high temperature (recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C, Figure 5 and Figure 6). The 1H-NMR spectra clearly showed the weak downfield multiplets at δ 5.90 and δ 5.00, the typical vinylic signal (Figure 5). While the integration ratio for H1:H2/H2’:HI is close to 1:2:3, indicating the only the unsaturated chain structure in the PE. The shift 114.38 and 139.54 in 13C-NMR spectra was ascribed to the unsaturated chain ends (–CH=CH2) [16,40,41,43,45], which agreed well with 1H-NMR spectra. In addition, the lower intensity peaks at 32.24, 22.92 and 14.22 corresponding to an n-propyl end-group further supported the linear structure of the polyethylene [16,18,45]. These findings are further evidenced by their high melting temperature (Tm = 130.0 °C, run 9, Table 2).
The resultant PE by Co2/MMAO were also characterized by 1H/13C-NMR spectra, and shown in Figure 7 and Figure 8. The 1H-NMR spectrum also showed the signal of a vinyl group (δ 5.90 and δ 5.00), which was also confirmed by the signal in 13C-NMR (δ 139.55, 114.40). However, the integration ratio for H1:H2/H2’:HI is close to 1:2:5 (Figure 7), dissimilar to that seen with MAO at the same temperature, which suggested both the saturated and unsaturated chain structure in polyethylene. Moreover, the signal of the lower intensity peaks at 32.24, 22.93 and 14.26 in 13C-NMR corresponds to n-propyl end-group, indicating the linear structure of PE. These observations would suggest that, chain termination via β-H elimination in Co2/MMAO system is no longer the sole chain transfer pathway operative, with chain transfer to aluminum now competitive [16,18,44].

3. Materials and Methods

3.1. General Considerations

All the synthetic procedures for air/moisture-sensitive compounds were performed under a nitrogen atmosphere with standard Schlenk techniques. Toluene as a solvent for polymerization was refluxed over sodium (a small amount of benzophenone) and distilled under nitrogen prior to use. Methylaluminoxane (MAO, 1.46 M solution in toluene) and modified methylaluminoxane (MMAO, 1.93 M in n-heptane) were bought from Akzo Nobel Corp (Nanjing, China). High-purity ethylene was bought from Beijing Yanshan Petrochemical company (Beijing, China) and used as received. NMR spectra were recorded using a Bruker DMX 400 MHz instrument (Beijing, China) at ambient temperature using TMS as an internal standard. FT-IR spectra were recorded using a Perkin-Elmer System 2000 FT-IR spectrometer (Shanghai, China). Elemental analysis was carried out using a Flash EA1112 microanalyzer (Beijing, China). Molecular weights and molecular weight distribution of polyethylenes were determined using a PL-GPC220 GPC/SEC High Temperature System (Beijing, China). Data collection and handling were carried out using Cirrus GPC Software (Beijing, China) and Multi Detector Software (Beijing, China). The calibrants for constructing conventional calibration is Polystyrene Calibration KitS-M-10 from PL Company (Beijing, China). The true average molecular weights of PE are transferred by inputting the M–H constants of PE. K of 0.727 and α of 40.6 are provided by PL Company (Beijing, China). Samples were dissolved at a concentration of 1.0 to 2.5 mg mL−1, depending on the molecular weights. DSC trace and melting points of polyethylene were obtained from the second scanning run on DSCQ2000 at a heating rate of 10 °C min−1 from −20 °C to 200 °C. 1H/13C-NMR spectra of the polyethylene were recorded using a Bruker DMX 300 MHz instrument (Beijing, China) at 120 °C in deuterated 1,1,2,2-tetrachloroethane with TMS as an internal standard. The compound of 2-acetyl-5,6,7,8-tetrahydrocycloheptapyridin-9-one was synthesized according to the literature [49] and the 2-cycloalkylaniline hydrochlorides were prepared using literature methods [34].

3.2. Synthesis of 2-(1-ArN) C2H3-9-ArN-5,6,7,8-C5H8C5H3N (L1L6)

Ar = 2-(C5H9)-6-MeC6H3 (L1/L1′). 2-cyclopentyl-6-methylphenylamine hydrochloride (0.46 g, 2.2 mmol) was added to a solution of (0.20 g, 1.0 mmol,) 2-acetyl-5,6,7,8-tetrahydrocycloheptapyridin-9-one with p-toluenesulfonic acid (20.0 mg) as catalyst in 30 mL of n-butanol. The mixture was stirred at reflux temperature for 6h. The solvent was then evaporated at reduced pressure, and the residue was purified by alumina column chromatography with petroleum ether/dichloromethane/triethylamine [100/1/1, v/v/v] to afford L1 (0.19 g, 38.1%) as a yellow oil with molar ratio of L1/L1′ = 1/0.12 (detected by 1H-NMR). 1H-NMR (CDCl3, 400 MHz, TMS): δ 8.39 (t, J = 7.1 Hz, 1H, L1HPy), 8.25 (d, J = 8.2 Hz, 1H, L1′HPy), 7.68 (d, J = 7.8 Hz, 1H, L1′HPy), 7.62 (d, J = 8.0 Hz, 1H, L1HPy), H), 7.18 (t, J = 8.5 Hz, 2H, 2 × L1HAr), 7.11 (t, J = 5.8 Hz, 2H, 2 × L1′HAr), 7.04 (t, J = 6.8 Hz, 2H, 2 × L1HAr and 2 × L1′HAr), 6.99 (d, J = 7.3 Hz, 2H, 2 × L1HAr), 6.85 (d, J = 7.0 Hz, 2H, 2 × L1′HAr), 6.46 (s, 1H, L1′HNH), 4.66 (t, J = 6.9 Hz, 1H, L1′HCH=), 3.03-2.91 (m, 2H, L1HCH2), 2.90–2.81 (m, 2H, L1HCH2), 2.77 (t, J = 6.3 Hz, 2H, L1′HCH2), 2.37 (s, 3H, L1′HCH3), 2.34 (t, J = 6.1 Hz, 2H, L1HCH2), 2.26 (s, 2H, L1HCH2), 2.21 (s, 3H, L1HCH3), 2.01 (d, J = 6.2 Hz, 6H, 2 × L1HPhCH3), 1.95–1.88 (m, 2H, 2 × L1HCH2), 1.82 (s, 4H, L1HCH2), 1.72 (t, J = 5.7 Hz, 4H, L1HCH2), 1.60 (d, J = 5.3 Hz, 4H, L1HCH2), 1.54 (s, 4H, L1HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 173.7, 167.5, 156.6, 155.0, 148.1, 137.4, 135.9, 134.4, 134.0, 128.1, 127.9, 125.5, 125.3 125.1, 124.1, 124.0,123.9, 123.5, 123.3, 121.6, 40.6, 40.5, 34.2, 34.1, 33.9, 33.7, 32.4, 31.9, 27.2, 26.3, 26.0, 25.9, 25.8, 24.1, 22.9, 18.5, 17.0, 16.9. FTIR (KBr, cm−1): 3372 (νN–H, w), 2946 (s), 2864 (m), 1642(νC=N, m), 1564 (w), 1455 (s), 1359 (m), 1302 (w), 1261 (m), 1192 (m), 1164 (w), 1118 (w), 1088 (m), 1032 (w), 848 (w), 805 (w), 746 (s). Anal. Calcd for C36H43N3: C, 83.51, H, 8.37, N, 8.12; Found: C, 83.18, H, 8.61, N, 8.43%.
Ar = 2-(C6H11)-6-MeC6H3 (L2/L2′). Using a similar procedure and molar ratios to that described for L1/L1′ but with 2-cyclohexyl-6-methylphenylamine hydrochloride as the amine, L2/L2′ was obtained as a yellow oil (0.22 g, 39.7%) with molar ratio of L2/L2′ = 1/0.11 (detected by 1H-NMR).1H-NMR (CDCl3, 400 MHz, TMS): δ 8.38 (t, J = 8.5 Hz, 1H, L2HPy), 8.24 (d, J = 7.8 Hz, 1H, L2′HPy), 7.68 (d, J = 7.7 Hz, 1H, L2′HPy), 7.63 (d, J = 7.6 Hz, 1H, L2HPy), H), 7.14 (t, J = 8.1 Hz, 2H, 2 × L2HAr and 2 × L2′HAr), 7.05 (m, 2H, 2 × L2HAr and 2 × L2′HAr), 6.99 (d, J = 7.3 Hz, 2H, 2 × L2HAr), 6.94 (s, 2H, 2 × L2′HAr), 6.41 (s, 1H, L2′HNH–), 4.66 (t, J = 6.9 Hz, 1H, L2′HCH=), 2.96 (t, J = 6.0 Hz, 2H, L2HCH2), 2.79–2.71 (m, 2H, L2HCH2), 2.64 (s, 2H, L2′HCH2), 2.38 (s, 3H, L2′HCH3), 2.34 (t, J = 6.3 Hz, 2H, L2HCH2), 2.28 (s, 2H, L2HCH2), 2.21 (s, 3H, L2HCH3), 2.01 (d, J = 5.0 Hz, 6H, 2 × L2HCH3-Ph), 1.97 (s, 2H, 2 × L2HCH2), 1.91–1.82 (m, 4H, L2HCH2), 1.78–1.67 (m, 8H, 2 × L2HCH2), 1.53 (s, 8H, 2 × L2HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 173.9, 167.5, 156.7, 155.1, 148.1, 137.5, 146.4, 138.2, 137.5, 135.9, 135.8, 135.4, 128.1, 128.0, 124.4, 124.0, 123.5, 123.4, 39.2, 38.8, 34.1, 33.6, 33.5, 33.2, 32.4, 31.2, 27.5, 27.3, 26.6, 26.3, 24.1, 18.6, 17.1. FT-IR (KBr, cm−1): 3379 (νN–H, w), 2951 (s), 2866 (m), 1644 (νC=N, m), 1564 (w), 1457 (s), 1362 (m), 1303 (w), 1268 (m), 1195 (m), 1164 (w), 1121 (w), 1088 (m), 1033 (w), 850 (w), 805 (w), 747 (s). Anal. Calcd for C38H47N3: C, 83.62, H, 8.68, N, 7.70; Found: C, 83.89, H, 8.31, N, 8.07.
Ar = 2-(C8H15)-6-MeC6H3 (L3/L3′). Using a similar procedure and molar ratios to that described for L1/L1′, but with 2-cyclooctyl-6-methylphenylamine hydrochloride as the amine, L3/L3′ was obtained as a yellow powder (0.21 g, 34.5%) with molar ratio of L3/L3′ = 1/0.28 (detected by 1H-NMR). 1H-NMR (CDCl3, 400 MHz, TMS): δ 8.38 (m, 1H, L3HPy), 8.24 (d, J = 7.6 Hz, 1H, L3′HPy), 7.68 (d, J = 7.3 Hz, 1H, L3′HPy), 7.63 (d, J = 8.0 Hz, 1H, L3HPy), 7.12 (t, J = 5.1 Hz, 2H, 2 × L3HAr and 2 × L3′HAr), 7.05–6.96 (m, 2H, 2 × L3HAr and 2 × L3′HAr), 6.92 (d, J = 7.4 Hz, 2H, 2 × L3HAr and 2 × L3′–HAr), 6.35 (s, 1H, L3′HNH), 4.64 (t, J = 6.5 Hz, 1H, L3′HCH=), 2.99–2.93 (m, 2H, L3HCH2), 2.78–2.74 (m, 2H, L3HCH2), 2.38 (s, 3H, L3′HCH3), 2.34 (t, J = 7.4 Hz, 2H, L3HCH2), 2.29 (s, 2H, L3HCH2), 2.22 (s, 3H, L3–HCH3), 2.01 (d, J = 6.2 Hz, 6H, 2 × L3HCH3-Ph), 1.88 (s, 2H, 2 × L3HCH2), 1.88–1.76 (m, 8H, 2 × L3HCH2), 1.72–1.59 (m, 8H, 4 × L3–HCH2), 1.53 (s, 8H, 2 × L3HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 173.7, 167.5, 156.6, 155.0, 148.1,147.5, 146.4, 138.2, 137.5, 135.9, 134.4, 134.0, 127.9, 127.8,127.7, 125.3, 125.1, 124.9, 124.6, 123.7, 123.4, 121.7, 118.5, 38.4, 35.5, 34.4, 33.9, 33.3, 32.7, 31.9, 27.2, 26.9, 26.8, 26.6, 26.5, 26.2, 24.0, 18.8, 18.7, 18.4, 17.4. FT-IR (KBr, cm−1): 3377 (νN–H, w), 2952 (s), 2865 (m), 1644 (νC=N, m), 1567 (w), 1456 (s), 1363 (m), 1304 (w), 1267 (m), 1198 (m), 1164 (w), 1123 (w), 1090 (m), 1033 (w), 851 (w), 806 (w), 749 (s). Anal. Calcd for C42H55N3: C, 83.81, H, 9.21, N, 6.98; Found: C, 84.12, H, 9.06, N, 6.68%.
Ar = 2-(C5H9)-6-Me2C6H2 (L4/L4′). Using a similar procedure and molar ratios to that described for L1/L1′ but with 2-cyclopentyl-4,6-dimethylphenylamine hydrochloride as the amine, L4/L4′ was obtained as a yellow powder (0.20 g, 36.6%) with molar ratio of L4/L4′ = 1/0.09 (detected by 1H-NMR). 1H-NMR (CDCl3, 400 MHz, TMS): δ 8.39 (t, J = 6.9 Hz, 1H, L4HPy), 8.26 (d, J = 8.2 Hz, 1H, L4′HPy), 7.69 (d, J = 7.6 Hz, 1H, L4′HPy), 7.61 (d, J = 8.3 Hz, 1H, L4HPy), 7.18 (t, J = 8.5 Hz, 2H, 2 × L4HAr), 7.13 (t, J = 5.6 Hz, 2H, 2 × L4′HAr), 7.07 (t, J = 6.8 Hz, 2H, 2 × L4HAr and 2 × L4′HAr), 6.97 (d, J = 7.2 Hz, 2H, 2 × L4HAr), 6.82 (d, J = 6.7 Hz, 2H, 2 × L4′HAr), 6.46 (s, 1H, L4′HNH), 4.66 (t, J = 7.2 Hz, 1H, L4′HCH=), 3.01–2.94 (m, 2H, L4HCH2), 2.89–2.78 (m, 2H, L4HCH2), 2.77 (t, J = 6.3 Hz, 2H, L4′HCH2),2.37 (s, 3H, L4′HCH2), 2.34 (t, J = 6.1 Hz, 2H, L4HCH2), 2.26 (s, 2H, L4HCH2), 2.21 (s, 3H, L4HCH3), 2.01 (d, J = 6.2 Hz, 6H, 2 × L4HCH3-Ph), 1.96 (s, 6H, 2 × L4HCH3-Ph), 1.88–1.83 (m, 2H, 2 × L4HCH2), 1.81 (s, 4H, L4HCH2), 1.74 (t, J = 5.7 Hz, 4H, L4HCH2), 1.61 (d, J = 5.3 Hz, 4H, L4HCH2), 1.54 (s, 4H, L4HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 173.4, 167.3, 156.6, 155.1, 147.9, 137.6, 136.2, 134.5, 134.2, 128.1, 127.9, 125.3, 125.3 125.3, 123.8, 124.1, 123.9, 123.7, 123.3, 121.8, 40.6, 40.2, 34.3, 34.1, 33.8, 33.7, 32.4, 31.6, 27.3, 26.3, 26.0, 25.9, 25.8, 24.0, 23.1, 18.5, 17.4, 16.9. FT-IR (KBr, cm−1): 3376 (νN–H, w), 2949 (s), 2869 (m), 1646 (νC=N, m), 1565 (w), 1458 (s), 1363 (m), 1307 (w), 1264 (m), 1195 (m), 1164 (w), 1119 (w), 1091 (m), 1033 (w), 849 (w), 804 (w), 747 (s). Anal. Calcd for C38H47N3: C, 83.62, H, 8.68, N, 7.70; Found: C, 83.37, H, 8.92, N, 7.59%.
Ar = 2-(C6H11)-6-Me2C6H2 (L5/L5′). Using a similar procedure and molar ratios to that described for L1/L1′ but with 2-cyclohexyl-4,6-dimethylphenylamine hydrochloride as the amine, L5/L5′ was obtained as a yellow powder (0.22 g, 38.5%) with molar ratio of L5/L5′ = 1/0.12 (detected by 1H-NMR). 1H-NMR (CDCl3, 400 MHz, TMS): δ 8.37 (t, J = 8.4 Hz, 1H, L5HPy), 8.24 (d, J = 8.0 Hz, 1H, L5′HPy), 7.66 (d, J = 7.6 Hz, 1H, L5′HPy), 7.62 (d, J = 7.6 Hz, 1H, L5HPy), 7.14 (t, J = 8.1 Hz, 2H, 2 × L5HAr and 2 × L5′HAr), 7.02 (m, 2H, 2 × L5HAr and 2 × L5′HAr), 6.96 (d, J = 7.1 Hz, 2H, 2 × L5HAr), 6.93 (s, 2H, 2 × L5′–HAr), 6.40 (s, 1H, L5′HNH), 4.65 (t, J = 6.9 Hz, 1H, L5′HCH=), 2.94 (t, J = 6.7 Hz, 2H, L5HCH2), 2.83–2.76 (m, 2H, L5HCH2), 2.67 (s, 2H, L5′HCH2), 2.42 (s, 3H, L5′HCH3), 2.38 (t, J = 6.4 Hz, 2H, L5HCH2), 2.31 (s, 2H, L5HCH2), 2.26 (s, 3H, L5HCH2), 2.05 (d, J = 5.0 Hz, 6H, 2 × L5HCH3-Ph), 2.01 (s, 6H, 2 × L5HCH3-Ph), 1.97 (s, 2H, 2 × L5HCH2), 1.88–1.82 (m, 4H, L5HCH2), 1.77–1.65 (m, 8H, 2 × L5HCH2), 1.54 (s, 8H, 2 × L5HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 174.1, 167.8, 156.9, 155.4, 148.2, 137.7, 146.4, 138.3, 137.5, 135.9, 135.7, 135.4, 128.2, 128.0, 124.4, 124.1, 123.5, 123.3, 39.4, 39.0, 34.3, 33.7, 33.4, 33.1, 32.4, 31.2, 27.6, 27.4, 26.5, 26.3, 24.0, 18.6, 17.2. FT-IR (KBr, cm−1): 3379 (νN–H, w), 2951 (s), 2864 (m), 1645 (νC=N, m), 1563 (w), 1457 (s), 1364 (m), 1301 (w), 1272 (m), 1194 (m), 1164 (w), 1120 (w), 1089 (m), 1034 (w), 850 (w), 804 (w), 748 (s). Anal. Calcd for C40H51N3: C, 83.72, H, 8.96, N, 7.32; Found: C, 83.45, H, 8.73, N, 7.58%.
Ar = 2-(C8H15)-6-Me2C6H2 (L6/L6′). Using a similar procedure and molar ratios to that described for L1/L1′ but with 2-cyclooctyl-4,6-dimethylphenylamine hydrochloride as the amine, L6/L6′ was obtained as a yellow powder (0.20 g, 32.1%) with crude molar ratio of L6/L6′ = 1/0.21 (detected by 1H-NMR). 1H-NMR (CDCl3, 400 MHz, TMS): δ 8.41 (m, 1H, L6HPy), 8.27 (d, J = 7.6 Hz, 1H, L6′HPy), 7.70 (d, J = 7.3 Hz, 1H, L6′–HPy), 7.66 (d, J = 8.0 Hz, 1H, L6HPy), 7.15 (t, J = 6.1 Hz, 2H, 2 × L6HAr and 2 × L6′HAr), 7.11–6.99 (m, 2H, 2 × L6HAr and 2 × L6′HAr), 6.94 (d, J = 7.7 Hz, 2H, 2 × L6HAr and 2 × L6′HAr), 6.38 (s, 1H, L6′HNH), 4.66 (t, J = 6.9 Hz, 1H, L6′HCH=), 3.07–3.01 (m, 2H, L6HCH2), 2.81–2.75 (m, 2H, L6HCH2), 2.42 (s, 3H, L6′HCH3), 2.35 (t, J = 7.4 Hz, 2H, L6HCH2), 2.33 (s, 2H, L6HCH2), 2.25 (s, 3H, L6–HCH3), 2.11 (d, J = 6.0 Hz, 6H, 2 × L6HCH3-Ph), 2.03 (d, J = 5.9 Hz, 6H, 2 × L6HCH3-Ph), 1.90 (s, 2H, 2 × L6HCH2), 1.88–1.76 (m, 8H, 2 × L6HCH2), 1.71–1.60 (m, 8H, 4 × L6HCH2), 1.51 (s, 8H, 2 × L6HCH2). 13C-NMR (CDCl3, 100 MHz, TMS): δ 174.6, 168.1, 157.3, 155.5, 148.5,147.9, 146.6, 138.3, 137.8, 136.2, 134.4, 134.1, 128.3, 127.9, 127.7, 125.7, 125.4, 124.9, 124.8, 123.7, 123.6, 121.9, 118.5, 38.4, 36.0, 34.6, 33.9, 33.4, 32.5, 32.0, 27.4, 26.9, 26.7, 26.4, 26.3, 26.0, 24.1, 18.8, 18.6, 18.2, 17.4. FT-IR (KBr, cm−1): 3374 (νN–H, w), 2948 (s), 2865 (m), 1644 (νC=N, m), 1565 (w), 1459 (s), 1361 (m), 1303 (w), 1263 (m), 1194 (m), 1165 (w), 1119 (w), 1090 (m), 1036 (w), 851 (w), 805 (w), 749 (s). Anal. Calcd for C44H59N3: C, 83.89, H, 9.44, N, 6.67; Found: C, 83.63, H, 9.26, N, 6.98%.

3.3. Synthesis of [2-(1-ArN) C2H3-9-ArN-5,6,7,8-C5H8C5H3N] CoCl2 (Co1Co6)

Ar = 2-(C5H9)-6-MeC6H3 (Co1). The ligand L1/L1′ (110.0 mg, 0.23 mmol) and CoCl2·6H2O (47.3 mg, 0.20 mmol) were dissolved in 5 mL of ethanol. The reaction mixture was stirred at room temperature for overnight after which time diethyl ether (30 mL) was added to precipitate the complex. The precipitate was collected by filtration and washed with diethyl ether (3 × 5 mL), followed by drying under vacuum to give complex Co1 as a light brown powder (122.0 mg, 94.0%). FT-IR (KBr, cm−1): 2947 (s), 2863 (m), 2167 (w), 2138 (w), 2012 (w), 1992 (w), 1608 (νC=N, m), 1571 (s), 1454 (s), 1369 (m), 1309 (w), 1257 (s), 1198 (s), 1113 (m), 1086 (w), 934 (w), 888 (w), 845 (m), 744 (vs). Anal. Calcd for C36H43Cl2CoN3: C, 66.77, H, 6.69, N, 6.49; Found: C, 66.42, H, 6.61, N, 6.55%.
Ar = 2-(C6H11)-6-MeC6H3 (Co2). Using a similar procedure and molar ratios to that described for Co1 but with L2/L2′ as the ligand, the complex Co2 was obtained as the light brown powder in 126.0 mg (yield: 93.5%). FT-IR (KBr, cm−1): 2923 (s), 2851 (m), 2168 (w), 2119 (w), 2022 (w), 1989 (w), 1606 (νC=N, m), 1570 (s), 1448 (s), 1369 (m), 1314 (w), 1260 (m), 1230 (s), 1194 (s), 1114 (m), 1088 (w), 936 (w), 886 (w), 846 (m), 774 (vs), 736 (w). Anal. Calcd for C38H47Cl2CoN3: C, 67.55, H, 7.01, N, 6.22; Found: C, 67.77, H, 6.94, N, 6.34%..
Ar = 2-(C8H15)-6-MeC6H3 (Co3). Using a similar procedure and molar ratios to that described for Co1 but with L3/L3′ as the ligand, the complex Co3 was obtained as the light brown powder in 130.0 mg (yield: 91.2%). FT-IR (KBr, cm−1): 2916 (s), 2853 (m), 2167 (w), 2082 (w), 2014 (w), 1989 (w), 1609 (νC=N, m), 1569 (s), 1465 (s), 1368 (m), 1313 (w), 1257 (s), 1193 (s), 1114 (m), 1085 (w), 931 (w), 845 (m), 771 (vs), 735 (w). Anal. Calcd for C42H55Cl2CoN3: C, 68.94, H, 7.58, N, 5.74; Found: C, 69.31, H, 7.47, N, 5.89%.
Ar = 2-(C5H9)-4,6-Me2C6H2 (Co4). Using a similar procedure and molar ratios to that described for Co1 but with L4/L4′ as the ligand, the complex Co4 was obtained as the light brown powder in 126.0 mg (yield: 93.3%). FT-IR (KBr, cm−1): 2916 (s), 2857 (m), 2166 (w), 2078 (w), 2032 (w), 1991 (w), 1609 (νC=N, m), 1569 (s), 1472 (s), 1448 (s), 1368 (m), 1311 (w), 1257 (s), 1205 (s), 1113 (m), 1085 (w), 926 (w), 850 (s), 768 (s). Anal. Calcd for C38H47Cl2CoN3: C, 67.55, H, 7.01, N, 6.22; Found: C, 67.82, H, 7.31, N, 6.06%.
Ar = 2-(C6H11)-4,6-Me2C6H2 (Co5). Using a similar procedure and molar ratios to that described for Co1 but with L5/L5′ as the ligand, the complex Co5 was obtained as the light brown powder in 135.0 mg (yield: 95.6%). FT-IR (KBr, cm−1): 2918 (vs), 2852 (s), 2169 (w), 2111 (w), 2011 (w), 1989 (w), 1611 (νC=N, m), 1569 (s), 1447 (s), 1367 (m), 1313 (w), 1258 (s), 1204 (m), 1116 (w), 1086 (w), 850 (s), 768 (s). Anal. Calcd for C40H51Cl2CoN3: C, 68.27, H, 7.31, N, 5.97; Found: C, 67.89, H, 7.47, N, 6.33%.
Ar = 2-(C8H15)-4,6-Me2C6H2 (Co6). Using a similar procedure and molar ratios to that described for Co1 but with L6/L6′ as the ligand, the complex Co6 was obtained as the light brown powder in 133.0 mg (yield: 87.4%). FT-IR (KBr, cm−1): 2916 (s), 2854 (m), 2166 (w), 2112 (w), 2012 (w), 1990 (w), 1611 (νC=N, m), 1568 (s), 1466 (s), 1441 (s), 1368 (m), 1314 (w), 1257 (m), 1212 (m), 1114 (w), 1084 (w), 926 (w), 852 (s), 768 (s). Anal. Calcd for C44H59Cl2CoN3: C, 69.55, H, 7.83, N, 5.53; Found: C, 69.81, H, 7.58, N, 5.85%.

3.4. X-ray Crystallographic Studies

Single-crystal X-ray diffraction studies of Co1 and Co3 were conducted on a Rigaku Sealed Tube CCD (Saturn 724+) diffractometer with graphite-mono chromated Mo-Kα radiation (λ = 0.71073 Å) at 173(2) K and the cell parameters obtained by global refinement of the positions of all collected reflections. Intensities were corrected for Lorentz and polarization effects and empirical absorption. The structures were solved by direct methods and refined by full-matrix least-squares on F2. All non-hydrogen atoms were refined anisotropically and all hydrogen atoms were placed in calculated positions. Structure solution and refinement were performed by using the SHELXT (Sheldrick, 2015) [55]. The disorder displayed by the carbon atoms in Co3 was processed by the SHELXL-2015 software [56]. Crystal data and processing parameters for Co2 and Co3 are summarized in Table 4. X-ray crystallographic data in CIF for the Cambridge Crystallographic Data Centre (CCDC) 1,897,124 (Co2) and 1,897,125 (Co3) are available free of charge from The Cambridge Crystallographic Data Centre.

3.5. General Procedure for Ethylene Polymerization under 5/10 Atm Pressure

A 250 mL stainless steel autoclave (Dalian Sanling Electronic Manufacture, Dalian, China), equipped with an ethylene pressure control system, a mechanical stirrer and a temperature controller, was employed for the reaction. The autoclave was evacuated and refilled with ethylene three times. When the desired reaction temperature was reached, toluene, the co-catalyst (MAO or MMAO), and a toluene solution of the catalytic precursor (the total volume was 100 mL) were injected into the autoclave by using syringes. Then, the ethylene pressure was increased to 5/10 atm, maintained at this level with constant feeding of ethylene. After the reaction was carried out for their quired period, the reactor was cooled with a water bath and the excess ethylene was vented. The reaction solution was quenched with 10% HCl/ethanol. The precipitated polymer was collected through filtration, washed with ethanol and dried under vacuum at 60 °C until constant weight.

4. Conclusions

A family of six cobalt (II) chloride complexes, Co1Co6, bound by single ring-fused 2-(1-cycloalkylphenylimino)ethyl-9-cycloalkylphenylimino-5,6,7,8-tetrahydrocyclo-heptapyridines, was prepared by the reaction of the corresponding carbocyclic-fused NNN-pincer ligands (L1/L1′L6/L6′) and cobalt(II) chloride and fully characterized. On activation with either MMAO or MAO, all these ortho cycloalkyl substituted cobalt complexes displayed good activity for ethylene polymerization with the optimum temperature of 50 °C, affording the linear polyethylene. The molecular weight (Mw = 9.78–25.6 kg mol−1) is higher than that observed with their analogues D bearing alkyl substituent. The ring size of the ortho substituent greatly affected their molecular weight and polymerization activity, which was demonstrated by the highest activity achieved by the cyclopentyl substituted ones (Co1 and Co4) and the highest molecular weight by cyclooctyl substituted ones (Co3 and Co6). Notably, polyethylene using MAO as activator displayed high selectivity for vinyl end-groups (–CH=CH2), while with MMAO the polyethylene possessed both the saturated and saturated linear structure.

Author Contributions

Design of the study and experiments, W.Z. and W.-H.S.; synthesis and catalysis, J.G., Z.W., I.I.O., I.V.O. and A.V.; manuscript, W.Z. and W.-H.S.; interpretation of the data obtained from the single crystal X-ray diffraction, Y.S.; chemical support, X.H.

Funding

This work was supported by the National Natural Science Foundation of China (Nos. 51473170, 21871275 and 51861145303) and the Russian Foundation for Basic Research (No. 18-53-80031_BRICS/5171102116). A.V thanks the Chinese Academy of Sciences President’s International Fellowship Initiative for the award of post-doctoral fellowship (Grant No. 2018PM0012).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PEPolyethylene
ORTEPOak Ridge Thermal Ellipsoid Plot
CIFCalibration Index File
GPCGel Permeation Chromatography
MAOmethylaluminoxane
MMAOModified methylaluminoxane
PDIPolydispersity index
TmMelting temperature

References

  1. Small, B.L.; Brookhart, M.; Bennett, A.M.A. Highly active iron and cobalt catalysts for the polymerization of ethylene. J. Am. Chem. Soc. 1998, 120, 4049–4050. [Google Scholar] [CrossRef]
  2. Small, B.L.; Brookhart, M. Iron-based catalysts with exceptionally high activities and selectivities for oligomerization of ethylene to linear α-olefins. J. Am. Chem. Soc. 1998, 120, 7143–7144. [Google Scholar] [CrossRef]
  3. Britovsek, G.J.P.; Gibson, V.C.; Kimberley, B.S.; Maddox, P.J.; McTavish, S.J.; Solan, G.A.; White, A.J.P.; Williams, D.J. Novel olefin polymerization catalysts based on iron and cobalt. Chem. Commun. 1998, 7, 849–850. [Google Scholar] [CrossRef]
  4. Britovsek, G.J.P.; Bruce, M.; Gibson, V.C.; Kimberley, B.S.; Maddox, P.J.; Mastroianni, S.; McTavish, S.J.; Redshaw, C.; Solan, G.A.; Stromberg, S.; et al. Iron and cobalt ethylene polymerization catalysts bearing 2,6-bis(Imino)pyridyl ligands: Synthesis, structures, and polymerization studies. J. Am. Chem. Soc. 1999, 121, 8728–8740. [Google Scholar] [CrossRef]
  5. Britovsek, G.J.P.; Mastroianni, S.; Solan, G.A.; Baugh, S.P.D.; Redshaw, C.; Gibson, V.C.; White, A.J.P.; Williams, D.J.; Elsegood, M.R. Oligomerisation of ethylene by bis(imino)pyridyliron and -cobalt complexes. Chem. Eur. J. 2000, 6, 2221–2231. [Google Scholar] [CrossRef]
  6. Wang, Z.; Solan, G.A.; Zhang, W.; Sun, W.-H. Carbocyclic-fused N,N,N-pincer ligands as ring-strain adjustable supports for iron and cobalt catalysts in ethylene oligo-/polymerization. Coord. Chem. Rev. 2018, 363, 92–108. [Google Scholar] [CrossRef]
  7. Small, B.L. Discovery and development of pyridine-bis(imine) and related catalysts for olefin polymerization and oligomerization. Acc. Chem. Res. 2015, 48, 2599–2611. [Google Scholar] [CrossRef]
  8. Burcher, B.; Breuil, P.-A.R.; Magna, L.; Olivier-Bourbigou, H. Iron-catalyzed oligomerization and polymerization Reactions. Top. Organomet. Chem. 2015, 50, 217–258. [Google Scholar]
  9. Flisak, Z.; Sun, W.-H. Progression of diiminopyridines: From single application to catalytic versatility. ACS Catal. 2015, 5, 4713–4724. [Google Scholar] [CrossRef]
  10. Ma, J.; Feng, C.; Wang, S.; Zhao, K.-Q.; Sun, W.-H.; Redshaw, C.; Solan, G.A. Bi- and tri-dentate imino-based iron and cobalt pre-catalysts for ethylene oligo-/polymerization. Inorg. Chem. Front. 2014, 1, 14–34. [Google Scholar] [CrossRef] [Green Version]
  11. Zhang, W.; Sun, W.-H.; Redshaw, C. Tailoring iron complexes for ethylene oligomerization and/or polymerization. Dalton Trans. 2013, 42, 8988–8997. [Google Scholar] [CrossRef]
  12. Gibson, V.C.; Solan, G.A. Olefin Oligomerizations and Polymerizations Catalyzed by Iron and Cobalt Complexes Bearing Bis(imino)pyridine Ligands. In Catalysis without Precious Metals; Bullock, M., Ed.; Wiley-VCH: Weinheim, Germany, 2010; pp. 111–141. [Google Scholar]
  13. Gibson, V.C.; Solan, G.A. Iron-based and cobalt-based olefin polymerisation catalysts. Top. Organomet. Chem. 2009, 26, 107–158. [Google Scholar]
  14. Gibson, V.C.; Redshaw, C.; Solan, G.A. Bis(imino)pyridines: Surprisingly reactive ligands and a gateway to new families of catalysts. Chem. Rev. 2007, 107, 1745–1776. [Google Scholar] [CrossRef] [PubMed]
  15. Bianchini, C.; Giambastiani, G.; Rios, I.G.; Mantovani, G.; Meli, A.; Segarra, A.M. Ethylene oligomerization, homopolymerization and copolymerization by iron and cobalt catalysts with 2,6-(bis-organylimino)pyridyl ligands. Coord. Chem. Rev. 2006, 250, 1391–1418. [Google Scholar] [CrossRef]
  16. Mahmood, Q.; Guo, J.; Zhang, W.; Ma, Y.; Liang, T.; Sun, W.-H. Concurrently improving the thermal stability and activity of ferrous precatalysts for the production of saturated/unsaturated polyethylene. Organometallics 2018, 37, 957–970. [Google Scholar] [CrossRef]
  17. Mahmood, Q.; Yue, E.; Guo, J.; Zhang, W.; Ma, Y.; Hao, X.; Sun, W.-H. Nitro-functionalized bis(imino)pyridylferrous chlorides as thermo-stable precatalysts for linear polyethylenes with high molecular weights. Polymer 2018, 159, 124–137. [Google Scholar] [CrossRef]
  18. Semikolenova, N.V.; Sun, W.-H.; Soshnikov, I.E.; Matsko, M.A.; Kolesova, O.V.; Zakharov, V.A.; Bryliakov, K.P. Origin of “multisite-like” ethylene polymerization behavior of the single-site nonsymmetrical bis(imino)pyridine iron(II) complex in the presence of modified methylaluminoxane. ACS Catal. 2017, 7, 2868–2877. [Google Scholar] [CrossRef]
  19. Zhao, W.; Yue, E.; Wang, X.; Yang, W.; Chen, Y.; Hao, X.; Cao, X.; Sun, W.-H. Activity and stability spontaneously enhanced toward ethylene polymerization by employing 2-(1-(2,4-Dibenzhydrylnaphthyl- imino)ethyl)-6-(1-(arylimino)ethyl)pyridyliron(II) dichlorides. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 988–996. [Google Scholar] [CrossRef]
  20. Mitchell, N.E.; Anderson, W.C.; Long, B.K. Mitigating chain-transfer and enhancing the thermal stability of co-based olefin polymerization catalysts through sterically demanding ligands. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 3990–3996. [Google Scholar] [CrossRef]
  21. Yue, E.; Zeng, Y.; Zhang, W.; Sun, Y.; Cao, X.-P.; Sun, W.-H. Highly linear polyethylenes using the 2-(1-(2,4-dibenzhydrylnaphthylimino)ethyl)-6-(1-(arylimino)ethyl)pyridylcobalt chlorides: Synthesis, characterization and ethylene polymerization. Sci. China Chem. 2016, 59, 1291–1300. [Google Scholar] [CrossRef]
  22. Zhang, W.; Wang, S.; Du, S.; Guo, C.-Y.; Hao, X.; Sun, W.-H. 2-(1-(2,4-Bis((di(4-fluorophenyl)ethyl)- methylphenylimino)ethyl)-6-(1-(arylimino)ethyl)pyridylmetal (iron or cobalt) complexes: Synthesis, characterization, and ethylene polymerization behavior. Macromol. Chem. Phys. 2014, 215, 1797–1809. [Google Scholar] [CrossRef]
  23. Wang, S.; Li, B.; Liang, T.; Redshaw, C.; Li, Y.; Sun, W.-H. Synthesis, characterization and catalytic behavior toward ethylene of 2-[1-(4,6-dimethyl-2-benzhydrylphenylimino)ethyl]-6-[1-(arylimino)ethyl] pyridylmetal (iron or cobalt) chlorides. Dalton Trans. 2013, 42, 9188–9197. [Google Scholar] [CrossRef]
  24. Smit, T.M.; Tomov, A.K.; Britovsek, G.J.P.; Gibson, V.C.; White, A.J.P.; Williams, D.J. The effect of imine-carbon substituents in bis(imino)pyridine-based ethylene polymerisation catalysts across the transition series. Catal. Sci. Technol. 2012, 2, 643–655. [Google Scholar] [CrossRef]
  25. Zhao, W.; Yu, J.; Song, S.; Yang, W.; Liu, H.; Hao, X.; Redshaw, C.; Sun, W.-H. Controlling the ethylene polymerization parameters in iron pre-catalysts of thetype 2-[1-(2,4-dibenzhydryl-6-methylphenylimino) ethyl]-6-[1-(arylimino)ethyl] pyridyliron dichloride. Polymer 2012, 53, 130–137. [Google Scholar] [CrossRef]
  26. Cao, X.; He, F.; Zhao, W.; Cai, Z.; Hao, X.; Shiono, T.; Redshaw, C.; Sun, W.-H. 2-[1-(2,6-Dibenzhydryl-4-chlorophenylimino)ethyl]-6-[1-(arylimino)ethyl]pyridyliron(II) dichlorides: Synthesis, characterization and ethylene polymerization behavior. Polymer 2012, 53, 1870–1880. [Google Scholar] [CrossRef]
  27. Lai, J.; Zhao, W.; Yang, W.; Redshaw, C.; Liang, T.; Liu, Y.; Sun, W.-H. 2-[1-(2,4-Dibenzhydryl-6-methylphenylimino)ethyl]-6-[1-(arylimino)ethyl]pyridylcobalt(II) dichlorides: Synthesis, characterization and ethylene polymerization behavior. Polym. Chem. 2012, 3, 787–793. [Google Scholar] [CrossRef]
  28. He, F.; Zhao, W.; Cao, X.-P.; Liang, T.; Redshaw, C.; Sun, W.-H. 2-[1-(2,6-Dibenzhydryl-4-chloro phenylimino)ethyl]-6-[1-aryliminoethyl]pyridylcobalt dichlorides: Synthesis, characterization and ethylene polymerization behavior. J. Organomet. Chem. 2012, 713, 209–216. [Google Scholar] [CrossRef]
  29. Yu, J.; Huang, W.; Wang, L.; Redshaw, C.; Sun, W.-H. 2-[1-(2,6-Dibenzhydryl-4-methylphenylimino) ethyl]-6-[1-(arylimino)ethyl]-pyridylcobalt(II) dichlorides: Synthesis, characterization and ethylene polymerization behavior. Dalton Trans. 2011, 40, 10209–10214. [Google Scholar] [CrossRef]
  30. Yu, J.; Liu, H.; Zhang, W.; Hao, X.; Sun, W.-H. Access to highly active and thermally stable iron procatalysts using bulky 2-[1-(2,6-dibenzhydryl-4-methylphenylimino)ethyl]-6-[1-(arylimino)ethyl] pyridine Ligands. Chem. Commun. 2011, 47, 3257–3259. [Google Scholar] [CrossRef]
  31. Guo, L.; Gao, H.; Zhang, L.; Zhu, F.; Wu, Q. An unsymmetrical iron(II) bis(imino)pyridyl catalyst for ethylene polymerization: Effect of a bulky ortho substituent on the thermostability and molecular weight of polyethylene. Organometallics 2010, 29, 2118–2125. [Google Scholar] [CrossRef]
  32. Wang, L.; Sun, W.-H.; Han, L.; Yang, H.; Hu, Y.; Jin, X. Late transition metal complexes bearing 2,9-bis(imino)-1,10-phenanthrolinyl ligands: Synthesis, characterization and their ethylene activity. J. Organomet. Chem. 2002, 658, 62–70. [Google Scholar] [CrossRef]
  33. Guo, L.; Zada, M.; Zhang, W.; Vignesh, A.; Zhu, D.; Ma, Y.; Liang, T.; Sun, W.-H. Highly linear polyethylenes tailored by 2,6-bis[1-(p-dibenzo-cycloheptylarylimino)ethyl]pyridylcobalt dichlorides. Dalton Trans. 2019. [Google Scholar] [CrossRef]
  34. Oleinik, I.I.; Oleinik, I.V.; Abdrakhmanov, I.B.; Ivanchev, S.S.; Tolstikov, G.A. Design of arylimine postmetallocene catalytic systems for olefin polymerization: I. Synthesis of substituted 2-cycloalkyl- and 2,6-dicycloalkylanilines. Russ. J. Gen. Chem. 2004, 74, 1423–1427. [Google Scholar] [CrossRef]
  35. Ivancheva, S.S.; Yakimansky, A.V.; Rogozin, G.D. Quantum-chemical calculations of the effect of cycloaliphatic groups in a-diimine and bis(imino)pyridine ethylene polymerization precatalysts on their stabilities with respect to deactivation reactions. Polymer 2004, 45, 6453–6459. [Google Scholar] [CrossRef]
  36. Ivanchev, S.S.; Tolstikov, G.A.; Badaev, V.K.; Oleinik, I.I.; Ivancheva, N.I.; Rogozin, D.G.; Oleinik, I.V.; Myakin, S.V. New bis(arylimino)pyridyl complexes as components of catalysts for ethylene polymerization. Kinet. Catal. 2004, 45, 176–182. [Google Scholar] [CrossRef]
  37. Tolstikov, G.A.; Ivanchev, S.S.; Oleinik, I.I.; Ivancheva, N.I.; Oleinik, I.V. New multifunctional bis(imino)pyridine–iron chloride complexes and ethylene polymerization catalysts on their basis. Dokl. Phys. Chem. 2005, 404, 208–211. [Google Scholar] [CrossRef]
  38. Ba, J.; Du, S.; Yue, E.; Hu, X.; Flisak, Z.; Sun, W.-H. Constrained formation of 2-(1-(arylimino)ethyl)-7-arylimino-6,6-dimethylcyclopentapyridines and their cobalt(II) chloride complexes: Synthesis, characterization and ethylene polymerization. RSC Adv. 2015, 5, 32720–32729. [Google Scholar] [CrossRef]
  39. Sun, W.-H.; Kong, S.; Chai, W.; Shiono, T.; Redshaw, C.; Hu, X.; Guo, C.; Hao, X. 2-(1-(Arylimino)ethyl)-8- arylimino-5,6,7-trihydroquinolylcobalt dichloride: Synthesis and polyethylene wax formation. Appl. Catal. A 2012, 447–448, 67–73. [Google Scholar] [CrossRef]
  40. Huang, F.; Zhang, W.; Yue, E.; Liang, T.; Hu, X.; Sun, W.-H. Controlling the molecular weights of polyethylene waxes using the highly active precatalysts of 2-(1-aryliminoethyl)-9-arylimino- 5,6,7,8-tetrahydrocycloheptapyridylcobalt chlorides: Synthesis, characterization, and catalytic behavior. Dalton Trans. 2016, 45, 657–666. [Google Scholar] [CrossRef] [PubMed]
  41. Huang, F.; Zhang, W.; Sun, Y.; Hu, X.; Solan, G.A.; Sun, W.H. Thermally stable and highly active cobalt precatalysts for vinyl-polyethylenes with narrow polydispersities: Integrating fused-ring and imino-carbon protection into ligand design. New J. Chem. 2016, 40, 8012–8023. [Google Scholar] [CrossRef]
  42. Appukuttan, V.K.; Liu, Y.; Son, B.C.; Ha, C.-S.; Suh, H.; Kim, I. Iron and cobalt complexes of 2,3,7,8-tetrahydroacridine-4,5(1H,6H)-diimine sterically modulated by substituted aryl rings for the selective oligomerization to polymerization of ethylene. Organometallics 2011, 30, 2285–2294. [Google Scholar] [CrossRef]
  43. Du, S.; Zhang, W.; Yue, E.; Huang, F.; Liang, T.; Sun, W.-H. α,α′-Bis(arylimino)-2,3:5,6-bis(penta methylene)pyridylcobalt chlorides: Synthesis, characterization, and ethylene polymerization behavior. Eur. J. Inorg. Chem. 2016, 1748–1755. [Google Scholar] [CrossRef]
  44. Suo, H.; Oleynik, I.I.; Bariashir, C.; Oleynik, I.V.; Wang, Z.; Solan, G.A.; Ma, Y.; Liang, T.; Sun, W.-H. Strictly linear polyethylene using Co-catalysts chelated by fused bis(arylimino)pyridines: Probing ortho-cycloalkyl ring-size effects on molecular weight. Polymer 2018, 149, 45–54. [Google Scholar] [CrossRef]
  45. Wang, Z.; Ma, Y.; Guo, J.; Liu, Q.; Solan, G.A.; Liang, T.; Sun, W.-H. Bis(imino)pyridines fused with 6- and 7-membered carbocylic rings as N,N,N-scaffolds for cobalt ethylene polymerization catalysts. Dalton Trans. 2019, 48, 2582–2591. [Google Scholar] [CrossRef]
  46. Bariashir, C.; Wang, Z.; Suo, H.; Muhammad, Z.; Solan, G.A.; Ma, Y.; Liang, T.; Sun, W.-H. Narrow dispersed linear polyethylene using cobalt catalysts bearing cycloheptyl-fusedbis(imino)pyridines; probing the effects of ortho-benzhydryl substitution. Eur. Polym. J. 2019, 110, 240–251. [Google Scholar] [CrossRef]
  47. Wang, Z.; Solan, G.A.; Mahmood, Q.; Liu, Q.; Ma, Y.; Hao, X.; Sun, W.-H. Bis(imino)pyridines incorporating doubly fused eight-membered rings as conformationally flexible supports for cobalt ethylene polymerization catalysts. Organometallics 2018, 37, 380–389. [Google Scholar] [CrossRef]
  48. Wang, Z.; Zhang, R.; Zhang, W.; Solan, G.A.; Liu, Q.; Liang, T.; Sun, W.-H. Enhancing thermostability of iron ethylene polymerization catalysts through N,N,N-chelation of doubly fused α,α′-bis(arylimino)-2,3:5,6-bis(hexamethylene)pyridines. Catal. Sci. Techn. 2019. [Google Scholar] [CrossRef]
  49. Huang, F.; Xing, Q.; Liang, T.; Flisak, Z.; Ye, B.; Hu, X.; Yang, W.; Sun, W.-H. 2-(1-Aryliminoethyl)-9-arylimino-5,6,7,8-tetrahydrocycloheptapyridyl iron(II) dichloride: Synthesis, characterization, and the highly active and tunable active species in ethylene polymerization. Dalton Trans. 2014, 43, 16818–16829. [Google Scholar] [CrossRef] [PubMed]
  50. Sun, Z.; Yue, E.; Qu, M.; Oleynik, I.V.; Oleynik, I.I.; Li, K.; Liang, T.; Zhang, W.; Sun, W.-H. 8-(2-Cycloalkylphenylimino)-5,6,7-trihydro-quinolylnickel halides: Polymerizing ethylene to highly branched and lower molecular weight polyethylenes. Inorg. Chem. Front. 2015, 2, 223–227. [Google Scholar] [CrossRef]
  51. Tomov, A.K.; Gibson, V.C.; Britovsek, G.J.P.; Long, R.J.; van Meurs, M.; Jones, D.J.; Tellmann, K.P.; Chirinos, J.J. Distinguishing chain growth mechanisms in metal-catalyzed olefin oligomerization and polymerization systems: C2H4/C2D4 co-oligomerization/polymerization experiments using chromium, iron, and cobalt catalysts. Organometallics 2009, 28, 7033–7040. [Google Scholar] [CrossRef]
  52. Jones, D.J.; Gibson, V.C.; Green, S.M.; Maddox, P.J.; White, A.J.P.; Williams, D.J. Discovery and optimization of new chromium catalysts for ethylene oligomerization and polymerization aided by high-throughput screening. J. Am. Chem. Soc. 2005, 127, 11037–11046. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, M.; Hao, P.; Zuo, W.; Jie, S.; Sun, W.-H. 2-(Benzimidazol-2-yl)-1,10-phenanthrolyl metal (Fe and Co) complexes and their catalytic behaviors toward ethylene oligomerization. J. Organomet. Chem. 2008, 693, 483–491. [Google Scholar] [CrossRef]
  54. Xiao, T.; Hao, P.; Kehr, G.; Hao, X.; Erker, G.; Sun, W.-H. Dichlorocobalt(II) complexes ligated by bidentate 8-(benzoimidazol-2-yl)quinolines: Synthesis, characterization, and catalytic behavior toward ethylene. Organometallics 2011, 30, 4847–4853. [Google Scholar] [CrossRef]
  55. Sheldrick, G.M. SHELXT-integrated space-group and crystalstructure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar]
  56. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
Sample Availability: Samples of the organic compounds and cobalt complexes are available from the authors.
Chart 1. Bis(imino)pyridine cobalt complexes and cycloalkyl ring fused derivatives.
Chart 1. Bis(imino)pyridine cobalt complexes and cycloalkyl ring fused derivatives.
Molecules 24 01176 ch001
Scheme 1. Synthetic route of ligands L1L6 and cobalt complexes Co1–Co6.
Scheme 1. Synthetic route of ligands L1L6 and cobalt complexes Co1–Co6.
Molecules 24 01176 sch001
Figure 1. ORTEP drawing of Co2 (a) and Co3 (b) with thermal ellipsoids at the 30% probability level. Hydrogen atoms are omitted for clarity.
Figure 1. ORTEP drawing of Co2 (a) and Co3 (b) with thermal ellipsoids at the 30% probability level. Hydrogen atoms are omitted for clarity.
Molecules 24 01176 g001
Figure 2. (a) GPC curves of polyethylene obtained by Co2/MAO at various temperature (runs 1–5, Table 2); (b) GPC curves of polyethylene obtained using Co2/MAO with different molar ratio of Al/Co (runs 3 and 6–11, Table 2).
Figure 2. (a) GPC curves of polyethylene obtained by Co2/MAO at various temperature (runs 1–5, Table 2); (b) GPC curves of polyethylene obtained using Co2/MAO with different molar ratio of Al/Co (runs 3 and 6–11, Table 2).
Molecules 24 01176 g002
Figure 3. (a) Polymerization activity and Mw of PE by Co1-Co6/MAO; (b) GPC curves of PE obtained by Co1-Co6/MAO (runs 9 and 17–21, Table 2).
Figure 3. (a) Polymerization activity and Mw of PE by Co1-Co6/MAO; (b) GPC curves of PE obtained by Co1-Co6/MAO (runs 9 and 17–21, Table 2).
Molecules 24 01176 g003
Chart 2. Comparison of the Mw, polydispersity (PDI) and polymerization activity of previously reported cobalt precatalysts (A [4], B [39], D [40], E [44] and H [38]) with MAO as activator under related condition.
Chart 2. Comparison of the Mw, polydispersity (PDI) and polymerization activity of previously reported cobalt precatalysts (A [4], B [39], D [40], E [44] and H [38]) with MAO as activator under related condition.
Molecules 24 01176 ch002
Figure 4. (a) Polymerization activity and Mw of PE by Co1-Co6/MMAO; (b) GPC curves of PE obtained by Co1Co6/MMAO (runs 3 and 12–16, Table 3).
Figure 4. (a) Polymerization activity and Mw of PE by Co1-Co6/MMAO; (b) GPC curves of PE obtained by Co1Co6/MMAO (runs 3 and 12–16, Table 3).
Molecules 24 01176 g004
Figure 5. 1H-NMR spectrum of the polyethylene obtained using Co2/MAO at 50 °C (run 9, Table 2); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Figure 5. 1H-NMR spectrum of the polyethylene obtained using Co2/MAO at 50 °C (run 9, Table 2); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Molecules 24 01176 g005
Figure 6. 13C-NMR spectrum of the polyethylene obtained using Co2/MAO at 50 °C (run 9, Table 2); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Figure 6. 13C-NMR spectrum of the polyethylene obtained using Co2/MAO at 50 °C (run 9, Table 2); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Molecules 24 01176 g006
Figure 7. 1H-NMR spectrum of the polyethylene obtained using Co2/MMAO at 50 °C (run 3, Table 3); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Figure 7. 1H-NMR spectrum of the polyethylene obtained using Co2/MMAO at 50 °C (run 3, Table 3); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Molecules 24 01176 g007
Figure 8. 13C-NMR spectrum of the polyethylene obtained using Co2/MMAO at 50 °C (run 3, Table 3); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Figure 8. 13C-NMR spectrum of the polyethylene obtained using Co2/MMAO at 50 °C (run 3, Table 3); recorded in 1,1,2,2-tetrachloroethane-d2 at 120 °C.
Molecules 24 01176 g008
Table 1. Selected bond lengths (Å) and angles (°) for Co2 and Co3.
Table 1. Selected bond lengths (Å) and angles (°) for Co2 and Co3.
Co2Co3 Co2Co3
Bond Lengths (Å)
Co(1)–N(1)2.203(4)2.193(4)Co(1)–N(2)2.063(4)2.043(4)
Co(1)–N(3)2.182(3)2.175(4)Co(1)–Cl(1)2.2922(13)2.3062(15)
Co(1)–Cl(2)2.2646(12)2.2478(14)N(1)–C(2)1.286(6)1.296(6)
N(3)–C(11)1.286(5)1.296(6)
Bond Angles (deg)
N(1)–Co(1)–N(2)73.80(14)74.91(16)N(1)–Co(1)–N(3)142.70(13)147.01(15)
N(2)–Co(1)–N(3)74.65(14)75.80(16)N(1)–Co(1)–Cl(2)97.55(11)98.74(11)
N(2)–Co(1)–Cl(2)152.75(11)151.03(13)N(3)–Co(1)–Cl(2)101.45(10)99.53(11)
N(1)–Co(1)–Cl(1)100.12(11)98.37(11)N(2)–Co(1)–Cl(1)95.66(11)89.97(12)
N(3)–Co(1)–Cl(1)102.30(10)96.45(12)Cl(2)–Co(1)–Cl(1)111.42(5)119.00(6)
C(11)–N(3)–Co(1)116.0(3)113.9(3)C(2)–N(1)–Co(1)114.9(3)114.8(4)
Table 2. Ethylene polymerization results using Co2/ MAO a.
Table 2. Ethylene polymerization results using Co2/ MAO a.
RunPrecat.Al/CoT (°C)t (min)PE (g)Act. bMwcMw/MncTm (oC) d
1Co2100030303.662.4422.973.7133.7
2Co2100040303.742.4917.943.8131.4
3Co2100050304.332.8914.213.0131.1
4Co2100060302.701.809.372.0130.2
5Co2100070301.450.966.562.5130.1
6Co25005030Trace----
7Co2150050304.372.919.762.4130.3
8Co2200050304.432.959.752.4130.7
9Co2250050305.363.579.392.7130.0
10Co2300050305.103.4010.672.6130.5
11Co2350050301.941.2913.962.5130.6
12Co225005052.038.1210.652.6130.3
13Co2250050154.706.2610.342.5130.7
14Co2250050456.282.7811.212.6131.5
15Co2250050607.442.4811.262.7130.1
16 eCo2250050304.282.8513.042.7130.8
17Co1250050306.134.0910.342.4130.4
18Co3250050304.953.3020.203.9132.4
19Co4250050305.253.5013.402.4131.6
20Co5250050304.873.259.782.4130.2
21Co6250050304.823.2125.964.4132.8
a Reaction conditions: 3 μmol cobalt complexes, 100 mL toluene, 10 atm C2H4; b Values in units of 106 g of PE mol−1 (Co) h−1; c Determined using GPC, Mw: kg mol−1; d Determined using DSC; e 5 atm C2H4.
Table 3. Ethylene polymerization results using Co2/MMAO a.
Table 3. Ethylene polymerization results using Co2/MMAO a.
RunPrecat.Al/CoT (°C)t (min)PE (g)Act. bMwcMw/MncTm (oC) d
1Co2250030303.342.2312.982.8132.5
2Co2250040303.822.5511.422.6130.3
3Co2250050304.432.958.192.5129.1
4Co2250060303.202.136.412.5128.1
5Co2150050303.622.4110.222.4130.1
6Co2200050303.752.509.272.7129.4
7Co2300050302.271.519.862.3129.9
8Co225005051.767.0410.492.5130.0
9Co2250050153.985.318.292.5129.1
10Co2250050606.212.0710.062.4131.2
11 eCo2250050302.851.908.902.5129.9
12Co1250050305.013.3410.012.4130.0
13Co3250050303.742.4914.722.2130.9
14Co4250050304.873.2510.792.4129.8
15Co5250050303.982.659.222.6129.6
16Co6250050303.362.2415.944.5133.3
a Reaction conditions: 3 μmol cobalt complexes, 100 mL toluene, 10 atm C2H4; b Values in units of 106 g of PE mol−1 (Co) h−1; c Determined using GPC, Mw: kg mol−1; d Determined using DSC; e 5 atm C2H4.
Table 4. Crystal data and structure refinement for Co2 and Co3.
Table 4. Crystal data and structure refinement for Co2 and Co3.
Co2Co3
CCDC No.18971241897125
Crystal colorBrownYellow
Empirical formulaC38H47Cl2CoN3C42H55Cl2CoN3
Formula weight675.61731.72
Temperature (K)173.15173.15(2)
Wavelength (Å)0.710730.71073
Crystal systemorthorhombictetragonal
Space groupFdd2I41/a
a (Å)33.4157(6)34.3337(8)
b (Å)31.4531(6)34.3337(8)
c (Å)16.9829(4)14.5297(7)
α (°)9090
β (°)9090
γ (°)9090
Volume (Å3)17849.5(6)17127.7(11)
Z1616
Dcalcd (g cm−3)1.0061.135
μ (mm−1)0.5280.555
F(000)5712.06224.0
Crystal size (mm3)0.288 × 0.193 × 0.1020.274 × 0.144 × 0.118
θ rang (°)4.56–63.0543.044–66.158
Limiting indices−46 ≤ h ≤ 48, −43 ≤ k ≤ 46, −24 ≤ l ≤ 23−50 ≤ h ≤ 48, −47 ≤ k ≤ 51, −21 ≤ l ≤ 10
No. of rflns collected7008849080
No. unique rflns [R(int)]0.0625(10266)0.1028(4627)
Completeness to θ (%)93.0 (θ = 25.00)90.6 (θ = 25.00)
Goodness of fit on F21.0400.971
Final R indices [I > 2σ(I)]R1 = 0.0625, wR2 = 0.1560R1 = 0.1028, wR2 = 0.2154
R indices (all data)R1 = 0.0868, wR2 = 0.1705R1 = 0.2894, wR2 = 0.3034
Largest diff. peak and hole (e Å−3)0.41 and −0.210.42 and −0.42

Share and Cite

MDPI and ACS Style

Guo, J.; Wang, Z.; Zhang, W.; Oleynik, I.I.; Vignesh, A.; Oleynik, I.V.; Hu, X.; Sun, Y.; Sun, W.-H. Highly Linear Polyethylenes Achieved Using Thermo-Stable and Efficient Cobalt Precatalysts Bearing Carbocyclic-Fused NNN-Pincer Ligand. Molecules 2019, 24, 1176. https://doi.org/10.3390/molecules24061176

AMA Style

Guo J, Wang Z, Zhang W, Oleynik II, Vignesh A, Oleynik IV, Hu X, Sun Y, Sun W-H. Highly Linear Polyethylenes Achieved Using Thermo-Stable and Efficient Cobalt Precatalysts Bearing Carbocyclic-Fused NNN-Pincer Ligand. Molecules. 2019; 24(6):1176. https://doi.org/10.3390/molecules24061176

Chicago/Turabian Style

Guo, Jingjing, Zheng Wang, Wenjuan Zhang, Ivan I. Oleynik, Arumugam Vignesh, Irina V. Oleynik, Xinquan Hu, Yang Sun, and Wen-Hua Sun. 2019. "Highly Linear Polyethylenes Achieved Using Thermo-Stable and Efficient Cobalt Precatalysts Bearing Carbocyclic-Fused NNN-Pincer Ligand" Molecules 24, no. 6: 1176. https://doi.org/10.3390/molecules24061176

Article Metrics

Back to TopTop