Next Article in Journal
Biological Activities of Natural Products III
Previous Article in Journal
Optimization of Supercritical Fluid CO2 Extraction from Yellow Horn Seed and Its Anti-Fatigue and Antioxidant Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N,N-Bis(2,4-Dibenzhydryl-6-cycloalkylphenyl)butane-2,3-diimine–Nickel Complexes as Tunable and Effective Catalysts for High-Molecular-Weight PE Elastomers †

1
School of Pharmaceutical Sciences, Changchun University of Chinese Medicine, Changchun 130117, China
2
Key Laboratory of Engineering Plastics, Beijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
3
Vorozhtsov Novosibirsk Institute of Organic Chemistry, Pr. Lavrentjeva 9, Novosibirsk 630090, Russia
4
Department of Chemistry, University of Leicester, University Road, Leicester LE1 7RH, UK
*
Authors to whom correspondence should be addressed.
The research was undertaken prior to the current Russian/Ukrainian conflict.
Molecules 2023, 28(12), 4852; https://doi.org/10.3390/molecules28124852
Submission received: 27 May 2023 / Revised: 13 June 2023 / Accepted: 15 June 2023 / Published: 19 June 2023
(This article belongs to the Section Organometallic Chemistry)

Abstract

:
Four examples of N,N-bis(aryl)butane-2,3-diimine–nickel(II) bromide complexes, [ArN=C(Me)-C(Me)=NAr]NiBr2 (where Ar = 2-(C5H9)-4,6-(CHPh2)2C6H2 (Ni1), Ar = 2-(C6H11)-4,6-(CHPh2)2C6H2 (Ni2), 2-(C8H15)-4,6-(CHPh2)2C6H2 (Ni3) and 2-(C12H23)-4,6-(CHPh2)2C6H2 (Ni4)), disparate in the ring size of the ortho-cycloalkyl substituents, were prepared using a straightforward one-pot synthetic method. The molecular structures of Ni2 and Ni4 highlight the variation in the steric hindrance of the ortho-cyclohexyl and -cyclododecyl rings exerted on the nickel center, respectively. By employing EtAlCl2, Et2AlCl or MAO as activators, Ni1Ni4 displayed moderate to high activity as catalysts for ethylene polymerization, with levels falling in the order Ni2 (cyclohexyl) > Ni1 (cyclopentyl) > Ni4 (cyclododecyl) > Ni3 (cyclooctyl). Notably, cyclohexyl-containing Ni2/MAO reached a peak level of 13.2 × 106 g(PE) of (mol of Ni)−1 h−1 at 40 °C, yielding high-molecular-weight (ca. 1 million g mol−1) and highly branched polyethylene elastomers with generally narrow dispersity. The analysis of polyethylenes with 13C NMR spectroscopy revealed branching density between 73 and 104 per 1000 carbon atoms, with the run temperature and the nature of the aluminum activator being influential; selectivity for short-chain methyl branches (81.8% (EtAlCl2); 81.1% (Et2AlCl); 82.9% (MAO)) was a notable feature. The mechanical properties of these polyethylene samples measured at either 30 °C or 60 °C were also evaluated and confirmed that crystallinity (Xc) and molecular weight (Mw) were the main factors affecting tensile strength and strain at break (εb = 353–861%). In addition, the stress–strain recovery tests indicated that these polyethylenes possessed good elastic recovery (47.4–71.2%), properties that align with thermoplastic elastomers (TPEs).

1. Introduction

Polyethylene is a general-purpose resin that can adopt many types of structure, anywhere from linear (high-density) to highly branched (low-density) structures, with the result that it can display a myriad of different mechanical properties [1]. In the field of ethylene polymerization, late-transition metal-mediated processes have garnered considerable attention, as such catalysts can be readily tuned by performing steric/electronic changes to the supporting chelating ligand, the type of metal center and their coordination environment [2,3,4]. Of particular note is the α-diimine family of nickel and palladium catalysts pioneered by Brookhart and co-workers, which have the ability to mediate the formation of branched polyethylene by using ethylene as the sole feed in polymerization. Central to the function of these Group 10 catalysts is their unique ability to promote a process known as chain walking leading to branch formation [5,6,7,8,9,10,11,12,13,14].
With particular reference to nickel catalysts, numerous previous studies have focused on modifying the α-diimine ligand backbone or changing the stereo/electronic properties of the N-aryl substituents. Indeed, these studies have culminated in significant improvements in the activity and thermal stability of α-diimine–nickel catalysts [15,16]. Elsewhere, our team and others have been interested in using such catalysts to generate polyethylenes with high branching density along with mechanical properties characteristic of thermoplastic elastomers (TPEs). Amongst the α-diimine ligand frameworks to be employed in this area, N,N-bis(aryl)butane-2,3-diimine and bis(imino)acenaphthene (A in Chart 1) have been studied the most intensively [5,17,18,19,20,21].
As concerns the N,N-bis(aryl)butane-2,3-diimine class, various developments have been disclosed. For example, B-type precatalysts (Chart 1) bearing relatively small ortho-alkyl substituents (R1 and R2) on both N-aryl groups of the chelating ligand display, following suitable activation, good performance characteristics in ethylene polymerization (activity ≤ 3.01 × 106 g(PE) of (mol of Ni)−1 h−1) [22], with an inclination towards generating high-molecular-weight polyethylene (Mw range: 2.22–8.97 × 105 g mol−1) exhibiting low branching density (18 branches per 1000 Cs; mainly Me branches). By contrast, C-type nickel precatalysts (Chart 1) incorporating bulkier benzhydryl (CHPh2) groups at both ortho-positions display slightly higher-molecular-weight polyethylene (Mw range: 4.4–11.8 × 105 g mol−1 [18]; 8.0–13.1 × 105 g mol−1 [23]), but lower catalytic activity (TOF = 2029–2578 h−1 [18], 0.98–1.58 × 106 g(PE) of (mol of Ni)−1 h−1 [23]). More noticeably, the branching density of polymers produced using C is higher (e.g., 63–75 branches per 1000 Cs [18]; 58 branches per 1000 Cs [23]) than that seen with B.
As an extension to our research program looking at AC, we now disclose a series of N,N-bis(aryl)butane-2,3-diimine–nickel precatalysts (Ni1Ni4 in Chart 1) that contain cycloalkyl substituents in combination with benzhydryl groups at the ortho-positions of the N-aryl groups. These carbocyclic groups were targeted to explore how the systematic modification of the ring size and flexibility (i.e., cyclopentyl, cyclohexyl, cyclooctyl, and cyclododecyl) can impact the catalytic performance and structural characteristics (e.g., branching density and composition) of polyethylene. Furthermore, stress–strain and dynamic mechanical analyses were performed on selected samples to evaluate their mechanical and elastic properties and to investigate how the polymerization conditions, such as activator type and operating temperature, can play a role. Full synthetic and characterization data of all new nickel complexes are additionally reported.

2. Results

2.1. Synthesis and Characterization

The N,N-bis(aryl)butane-2,3-diimine–nickel(II) bromide complexes, [ArN=C(Me)-C(Me)=NAr]NiBr2 (where Ar = 2-(C5H9)-4,6-(CHPh2)2C6H2 (Ni1), Ar = 2-(C6H11)-4,6-(CHPh2)2C6H2 (Ni2), 2-(C8H15)-4,6-(CHPh2)2C6H2 (Ni3) and 2-(C12H23)-4,6-(CHPh2)2C6H2 (Ni4)), were prepared using an in situ one-pot method involving the reaction of 2,3-butanedione, the corresponding 2-cycloalkylaniline and (DME)NiBr2 (DME = 1,2-dimethoxyethane) in acetic acid at reflux (Scheme 1) [24,25]. On work-up, Ni1Ni4 were isolated as brown powders in reasonable yields (17–64%) and characterized with FT-IR spectroscopy, elemental analysis and, in the case of Ni2 and Ni4, single-crystal X-ray diffraction.
Single crystals of Ni2 and Ni4 suitable for X-ray determination were grown using slow diffusion as described in the experimental section. Perspective views of each are shown in Figure 1 and Figure 2; selected bond lengths and angles are listed in Table 1.
The structures of both Ni2 and Ni4 comprise a single nickel center that is surrounded by two nitrogen atoms from the bidentate N,N-bis(aryl)butane-2,3-diimine (aryl = 2-cyclohexyl-4,6-dibenzhydrylphenyl (Ni2) and 2-cyclododecyl-4,6-dibenzhydrylphenyl (Ni4)) and two bromide ligands to complete a geometry that can be best described as distorted tetrahedral; similar geometries have been previously reported for this class of N,N-nickel(II) complexes [18,19,26,27]. Some modest variation in the N1-Ni1-N2 bite angles in each is apparent, with values of 81.36(4)° (Ni2) and 80.60(2)° (Ni4), while the disparity in the Br1-Ni1-Br2 angles is slightly more distinct (123.970(11)° (Ni2) vs. 125.57(5)° (Ni4)). The Ni-N distances (2.0221(11), 2.0081(10) Å (Ni2) vs. 2.014(5), 2.004(5) Å (Ni4)) and the Ni-Br distances (2.3331(3), 2.3418(3) Å (Ni2) vs. 2.3304(13), 2.3361(13) Å (Ni4)) show minimal differences and are indeed comparable to those found in previous reports [18,19,26]. Similarly, the imine C=N bond lengths within the chelating ligand are typical for this functional group (range: 1.280(8)–1.2856(17) Å) [18,19,26,27]. The N-aryl rings are inclined towards the perpendicular with respect to the neighboring imine vectors (Ni2 (88.48°, 87.00°) and Ni4 (82.55°, 79.78°)), with the two ortho-cycloalkyl groups (and ortho-benzhydryl groups) being positioned on opposite sides of the diimine coordination plane. The carbocyclic rings themselves are puckered with the cyclohexyl ring in Ni2, adopting a chair conformation, while the larger cyclododecyl ring in Ni4 can be best viewed as assuming a boat–chair–boat conformation. There are no intermolecular contacts of note.

2.2. Ethylene Polymerization Studies

2.2.1. Selection of Aluminum Activator

With the aim to determine the most suitable activator that can afford the highest catalytic activity for the nickel precatalyst, cyclohexyl-substituted Ni2 was chosen as the test precatalyst. Five different aluminum-alkyl reagents were initially screened, namely, Et2AlCl (diethylaluminum chloride), EtAlCl2 (ethylaluminum dichloride), EASC (ethyl sesquichloride), MAO (methylaluminoxane) and MMAO (modified methylaluminoxane). A 30 min polymerization run was performed with each Ni2/activator combination, with toluene being employed as reaction solvent and ethylene pressure being maintained at 10 atm; the Al:Ni ratios employed for each aluminum alkyl were based on typical values described elsewhere [22,23,28]. The results of this preliminary screening are listed in Table 2.
The inspection of the results reveals that all five aluminum alkyls were able to successfully activate Ni2 but with varying levels of effectiveness: EtAlCl2 > Et2AlCl > MAO > MMAO > EASC. Based on the superior performance displayed by EtAlCl2, Et2AlCl and MAO, these three activators were selected for more in-depth evaluation (see below).

2.2.2. Ethylene Polymerization Achieved with Activation of Ni1Ni4 with EtAlCl2

In order to optimize the performance of Ni2/EtAlCl2 and to establish an effective set of conditions that can be used to screen the other three nickel precatalysts, the effect of Al:Ni molar ratio, run temperature and reaction time were all systematically studied; the complete set of results is shown in Table 3.
With the Al:Ni molar ratio of Ni2/EtAlCl2 set at 500:1, the polymerization temperature was increased in ten-degree increments from 20 °C to 70 °C (runs 1–6; Table 3). Initially, the activity was seen to rise with the temperature, reaching the maximum activity of 10.3 × 106 g(PE) of (mol of Ni)−1 h−1 at 30 °C (run 2; Table 3). However, at temperatures above 30 °C, some loss in performance was observed (runs 3–6; Table 3), which can be accredited to the partial deactivation of the active species and/or a decrease in the solubility of the ethylene monomer at higher temperatures [18,19,26,27,29]. Nonetheless, an appreciable level of performance was still maintained at 60 °C (6.60 × 106 g(PE) of (mol of Ni)−1 h−1), highlighting the good thermostability of this catalyst (run 5; Table 3). In terms of the molecular weight of the polymer, this decreased gradually over the temperature range, while its dispersity became steadily narrower (Mw/Mn: 2.9 to 2.6). Indeed, the highest molecular weight was reached at 20 °C (12.5 × 105 g mol−1), while at 60 °C and 70 °C, the molecular weight dropped to 4.64 × 105 g mol−1 and 3.89 × 105 g mol−1, respectively. These temperature effects were further borne out by their GPC traces (Figure S1) and can be attributed to thermally induced chain transfer [29,30,31,32]. As a further point, the melting points of the polymers (Tm range of 104.6–44.5 °C) tended to decrease with the increase in polymerization temperature, a finding that is expected to derive from molecular weight and branching variations (see below).
Next, we set about determining the optimum Al:Ni molar ratio of Ni2/EtAlCl2 to promote the highest catalytic activity. Based on the above activity/temperature findings, the polymerization temperature was fixed at 30 °C, and the Al:Ni molar ratio was gradually increased from 300:1 to 700:1 (runs 2 and 7–10; Table 3). At 500:1, the activity attained an uppermost value of 10.3 × 106 g(PE) of (mol of Ni)−1 h−1 (run 2; Table 3) and then, at higher ratios, showed a modest decrease, dropping to 9.12 × 106 g(PE) of (mol of Ni)−1 h−1 at 700:1 (run 10; Table 3). More obviously, the activity increased when the Al:Ni molar ratio was adjusted from 300:1 to 500:1 (runs 2, and 7 and 8; Table 3), an observation that points towards a critical amount of EtAlCl2 being needed to induce good activity. As previously observed, the relative amount of EtAlCl2 used can affect the molecular weight of the polymer [33,34,35]. In this case, when the Al:Ni ratio varied between 300:1 and 700:1, the molecular weights of the resulting polymers were all high, with their levels ranging from 5.71 × 105 g mol−1 at 300:1 to 7.08 × 105 g mol−1 at 600:1 (Figure S2). The dispersity of the polymers was consistently narrow (Mw/Mn range: 2.4–3.1), in line with reasonable control of polymerization.
The time/activity profile of Ni2/EtAlCl2 was then explored with the polymerization temperature being kept at 30 °C, and the amount of EtAlCl2, at 500 equivalents. By employing set run times of 5 min, 15 min, 30 min, 45 min and 60 min, it was observed that the activity reached a maximum after 5 min of 10.9 × 106 g(PE) of (mol of Ni)−1 h−1 and then slowly decreased to 9.24 × 106 g(PE) of (mol of Ni)−1 h−1 after 60 min (runs 2 and 11–14; Table 3). This observation is consistent with a rapid induction period for precatalyst activation, which was then followed by a relatively modest loss in activity over time (~15%), which highlights the sizable lifetime of the active catalyst. By contrast, the molecular weight of the polymer increased with time from 3.1 × 105 g mol−1 after 5 min to 12.1 × 105 g mol−1 after 60 min (Figure S3), while the dispersity remained relatively narrow (Mw/Mn range of 2.3–2.8), underlining the good control over time exhibited by this catalyst.
To examine the effect of ring size variations in the ortho-cycloalkyl group on catalytic performance, the remaining three nickel(II) bromide complexes (Ni1, Ni3 and Ni4) were evaluated for ethylene polymerization using the optimized reaction conditions identified using Ni2/EtAlCl2 (viz., Al:Ni molar ratio = 500:1, run temp. = 30 °C and run time = 30 min); the complete set of data including those on Ni2 are shown in Table 3 (runs 2 and 17–19). By analyzing the data, it is evident that all four nickel complexes showed high to moderate catalytic activity (range: 10.3 × 106 g(PE) of (mol of Ni)−1 h−1 to 0.26 × 106 g(PE) of (mol of Ni)−1 h−1), with the relative order being Ni2 (cyclohexyl) > Ni1 (cyclopentyl) > Ni4 (cyclododecyl) > Ni3 (cyclooctyl). Based on these findings, it is evident that the steric factors exerted by the particular ortho-cycloalkyl group affected the catalytic activity, with the six-membered ring proving optimal. Furthermore, the flexibility of the ortho-cycloalkyl ring also played a role, as evidenced by 12-membered ring-containing Ni4 displaying higher catalytic activity than 8-membered Ni3 [9]. As for the polyethylenes generated, a noticeable variation in the molecular weight of polyethylenes produced with Ni1Ni4 could be seen, with values ranging from 4.37–8.22 × 105 g mol−1 (Figure S4), with Ni2 delivering the top-end value. Good control was again apparent for all four catalysts, with polymer dispersity values (Mw/Mn) between 2.1 and 2.9 being found.

2.2.3. Ethylene Polymerization Achieved with Activation of Ni1Ni4 with Et2AlCl

To investigate the impact of the aluminum activator on both the performance of the nickel precatalyst and on the polyethylene properties, the study was extended to include Et2AlCl. Using an optimization approach similar to that performed using EtAlCl2, Ni2 was again used as the precatalyst; the results of the evaluation using Et2AlCl are shown in Table 4.
As in the previous study employing Ni2/EtAlCl2, the temperature of the polymerization using Ni2/Et2AlCl was adjusted in this case from 30 °C to 70 °C, with the Al:Ni molar ratio being set at 500:1 (runs 1–5; Table 4). The results reveal that the maximum activity of 10.7 × 106 g(PE) of (mol of Ni)−1 h−1 was seen at 40 °C rather than 30 °C for Ni2/EtAlCl2. Indeed, at 30 °C, Ni2/Et2AlCl and Ni2/EtAlCl2 displayed comparable performance (10.3 × 106 g(PE) of (mol of Ni)−1 h−1 EtAlCl2 vs. 10.1 × 106 g(PE) of (mol of Ni)−1 h−1 Et2AlCl). On the other hand, at 50 °C, the catalytic activity of Ni2/Et2AlCl started to show some noticeable loss in activity (7.16 × 106 g(PE) of (mol of Ni)−1 h−1), which is consistent with the onset of partial deactivation of the active species and/or the lower solubility of ethylene at higher temperatures. Nonetheless, at 60 or 70 °C, further loss in activity was minimal, with a level of 6.40 × 106 g(PE) of (mol of Ni)−1 h−1 being noted at 70 °C. By comparison, Ni2/EtAlCl2 exhibited a more significant decline in performance between 50 °C and 70 °C (from 9.24 × 106 g(PE) of (mol of Ni)−1 h−1 to 2.26 × 106 g(PE) of (mol of Ni)−1 h−1), which highlights apparent variations in catalyst thermal stability.
In terms of the influence of run temperature on polymer molecular weight using Ni2/Et2AlCl, it was evident that the molecular weight of polyethylene fell as the temperature increased (from 9.98 to 5.09 × 105 g mol−1). This downward trend was particularly noticeable at run temperatures between 50 °C and 60 °C, where Mw dropped from 7.97 × 105 g mol−1 to 5.50 × 105 g mol−1 (Figure S5). Furthermore, some modest narrowing of the dispersity of the polymer was seen with the increase in temperature (Mw/Mn: from 2.5 to 2.2).
To understand how the Al:Ni molar ratio affected the performance of Ni2/Et2AlCl, this was increased from 300:1 to 700:1, with the run temperature being set at 40 °C, and the run time, at 30 min (runs 2 and 6–9; Table 4). The highest polymerization activity was obtained at 500:1 (10.7 × 106 g(PE) of (mol of Ni)−1 h−1). Moreover, it was apparent that the polymerization activity dramatically increased from 3.24 × 106 g(PE) of (mol of Ni)−1 h−1 to 9.22 × 106 g(PE) of (mol of Ni)−1 h−1 when the ratio was raised from 300:1 to 400:1, emphasizing the critical amount of activator required; similar findings are noted above and in the previous literature [36,37]. It is also noteworthy that the polymerization activity seen at 700:1 (5.54 × 106 g(PE) of (mol of Ni)−1 h−1) was significantly lower than that at 600:1 (9.74 × 106 g(PE) of (mol of Ni)−1 h−1), which indicates that excessive amounts of Et2AlCl can deactivate the catalyst. All the resulting polymers displayed high molecular weights ranging from 5.43 × 105 g mol−1 to 8.85 × 105 g mol−1 (Figure S6), with molar ratios in excess of 500:1 tending to form lower-molecular-weight polyethylene; similar observations have been reported [38,39]. In terms of polymer dispersity (Mw/Mn), the values were found between 2.4 and 3.3 but with no clear, identifiable trends.
Polymerization runs using Ni2/Et2AlCl were then performed over set time periods of 5 min, 15 min, 30 min, 45 min and 60 min (runs 2 and 10–13; Table 4), with the reaction temperature being kept at 40 °C, and the Al:Ni molar ratio, at 500:1. The examination of the data revealed that the catalyst, in this case, exhibited a relatively long induction period before the active species was fully generated, with the highest activity (13.6 × 106 g(PE) of (mol of Ni)−1 h−1) being seen after 15 min (run 11; Table 4). Once reached, the activity progressively declined, reaching the low point of 6.43 × 106 g(PE) of (mol of Ni)−1 h−1 at the one-hour mark. Despite this drop in performance, the level remained considerable even after one hour, indicating that the active species displayed good stability and an appreciable catalytic lifetime. As seen in Ni2/EtAlCl2, the molecular weight of the polymer formed using Ni2/Et2AlCl increased with time (Figure S7) and indeed in a manner akin to that previously reported [38,40].
With an optimum set of conditions for Ni2/Et2AlCl now established (Al:Ni molar ratio = 500:1, run temp. = 40 °C and run time = 30 min), the three remaining nickel precatalysts (Ni1, Ni3 and Ni4) were screened using these conditions; the full set of data is presented in Table 4 (runs 2 and 16–18; Table 4). Collectively, all precatalysts showed moderate to high activity (range: 0.04–10.7 × 106 g(PE) of (mol of Ni)−1 h−1) with the relative order being Ni2 (cyclohexyl) > Ni1 (cyclopentyl) > Ni4 (cyclododecyl) > Ni3 (cyclooctyl). This trend in performance mirrors that seen when using EtAlCl2, with the steric properties and ring flexibility of the ortho-cycloalkyl group being again influential on the catalytic activity. In addition, it can be seen that the molecular weight of polyethylene produced with Ni1Ni4 varied considerably, with values ranging from 3.69 to 8.85 × 105 g mol−1, with cyclohexyl-containing Ni2 once more generating the highest value (Figure S8). As with EtAlCl2 as activator, this class of catalyst again demonstrated good control, as evidenced by the narrow dispersity of the polymer (Mw/Mn range: 2.1–2.5).

2.2.4. Ethylene Polymerization Achieved with Activation of Ni1Ni4 with MAO

By employing an approach similar to that described for Ni2/EtAlCl2 and Ni2/Et2AlCl, Ni2/MAO was initially optimized to establish an effective set of reaction parameters that could be applied to Ni1, Ni2 and Ni3; the full set of results is shown in Table 5.
In terms of temperature response (runs 1–5; Table 5), a peak in the activity of Ni2/MAO of 13.2 × 106 g(PE) of (mol of Ni)−1 h−1 was seen at 40 °C, which then steadily decreased as the temperature was further raised. Notably, this optimum run temperature was also seen with Ni2/Et2AlCl (cf. 30 °C with Ni2/EtAlCl2), but with Ni2/MAO displaying a significantly higher level of performance. Nonetheless, Ni2/MAO still maintained high activity at 60 °C (7.82 × 106 g(PE) of (mol of Ni)−1 h−1) and 70 °C (6.54 × 106 g(PE) of (mol of Ni)−1 h−1). With regard to the molecular weight of the polymer, this was found to decline with an increase in temperature (Mw: from 10.1 × 105 g mol−1 at 30 °C to 5.75 × 105 g mol−1 at 70 °C), while the dispersity (Mw/Mn range: 1.9–2.5) remained narrow; the corresponding GPC traces are shown in Figure S9. As noted in the investigations with EtAlCl2 and Et2AlCl, the rate of chain transfer increased as the temperature increased, leading to the partial deactivation of the active species.
With the run temperature being maintained at 40 °C, the effect of the Al:Ni molar ratio on Ni2/MAO was investigated by increasing it from 1500:1 to 2500:1 (runs 2 and 6–9; Table 5). The scrutiny of the data revealed peak performance to have occurred at 2000:1 (13.2 × 106 g(PE) of (mol of Ni)−1 h−1), above which the level of activity gently decreased to 12.0 × 106 g(PE) of (mol of Ni)−1 h−1 at 2500:1. On the other hand, the activity seen at 1500:1 was notably lower (9.50 × 106 g(PE) of (mol of Ni)−1 h−1), highlighting the critical number equivalents of MAO required to fully activate Ni2; similar observations have been noted elsewhere [20,34]. With respect to the molecular weight of polyethylene, the variation in the Al:Ni molar ratio did not have a clear impact, with values ranging from 9.33 × 105 g mol−1 to 10.9 × 105 g mol−1 (Figure S10); once again, the dispersity was generally narrow (Mw/Mn range: 2.0–2.4).
Run time effects were also investigated using Ni2/MAO with separate polymerization reactions conducted over 5 min, 15 min, 30 min, 45 min and 60 min, with the run temperature and Al:Ni molar ratio being kept at 40 °C and 2000:1, respectively (runs 2 and 10–13; Table 5). As noted in Ni2/EtAlCl2, maximum polymerization activity was observed after 5 min (13.8 × 106 g(PE) of (mol of Ni)−1 h−1), followed by a gradual decrease, reaching a minimum value of 10.7 × 106 g(PE) of (mol of Ni)−1 h−1 after 60 min. Clearly, the active species formed quickly following the addition of MAO, which is dissimilar to the 15 min induction period observed with Et2AlCl. Furthermore, the relatively modest loss in activity (~22%) over time indicates that the active catalyst displayed a substantial lifetime but slightly less than the ~15% loss seen using Ni2/EtAlCl2. With respect to the molecular weight of the polymer, it increased from 7.24 × 105 g mol−1 after 5 min to 9.68 × 105 g mol−1 at 30 min; then, between 45 and 60 min, an unexpected reduction in weight was seen (from 8.69 × 105 g mol−1 to 7.78 × 105 g mol−1). The corresponding GPC traces are shown in Figure S11. The origin of this variation is uncertain but may have plausibly been due to the regeneration of active species having occurred during polymerization. Nonetheless, the dispersity of the polymer remained narrow across the different run times (Mw/Mn range: 1.9–2.0).
With the optimal conditions in place for Ni2/MAO (Al:Ni molar ratio = 2000:1, run temp. = 40 °C and run time = 30 min), Ni1, Ni3 and Ni4 were then evaluated similarly. All the complexes showed moderate to high activity (range: 0.52–13.2 × 106 g(PE) of (mol of Ni)−1 h−1), with the relative order being Ni2 (cyclohexyl) > Ni1 (cyclopentyl) > Ni4 (cyclododecyl) > Ni3 (cyclooctyl). As before, the ring size and flexibility of the cycloalkyl group proved crucial, with peak performance being again seen in the six-membered ring. As noted using the other activators, the molecular weight of polyethylenes produced with Ni1Ni4 varied considerably, with values ranging from 8.22 × 105 g mol−1 to 15.0 × 105 g mol−1 (Figure S12). However, in this case, the larger-ring precatalysts, Ni3 (cyclooctyl) and Ni4 (cyclododecyl), formed the highest-molecular-weight polymers. Nevertheless, all nickel catalysts produced polymers with narrow dispersity (Mw/Mn range: 2.0–2.1).
As is evident from the three studies conducted at PC2H4 = 10 atm using EtAlCl2, Et2AlCl or MAO as activator, the most productive catalyst was ortho-cyclohexyl-containing Ni2 in combination with MAO (run 2; Table 5). To explore how ethylene pressure affected this performance, runs were additionally performed at 5 atm and 1 atm. At 5 atm ethylene, the activity of Ni2/MAO decreased dramatically from 13.2 × 106 g(PE) of (mol of Ni)−1 h−1 to 3.12 × 106 g(PE) of (mol of Ni)−1 h−1 (run 14; Table 5), while the molecular weight of the polymer dropped from 9.68 × 105 g mol−1 to 7.87 × 105 g mol−1. Further reducing the pressure to 1 atm saw the activity decrease to 0.62 × 106 g (PE) of (mol of Ni)−1 h−1 (run 15; Table 5), and the molecular weight of the polymer, to 3.53 × 105 g mol−1. Likewise, comparable pressure effects were noted with EtAlCl2 (see runs 2, 15 and 16; Table 3) and Et2AlCl (see runs 2, 14 and 15; Table 4). Evidently, a critical value of ethylene pressure is also required to achieve good performance with this class of catalysts.
To further explore the influence of the ortho-cycloalkyl group, we show together activity and molecular weight data for Ni2 alongside those on four structurally related nickel(II) bromide precatalysts (B1, B4, B5 and C; see Chart 1) in Figure 3 [22,23,29]. The examination of the bar chart revealed that the catalytic activity values of B1, B4, B5 and C were significantly lower than that observed with Ni2. It would seem likely that the reduced steric properties of B1, B4 and B5, as a result of their particular ortho-substituents, were a contributing factor. On the other hand, the more excessive hindrance provided by the 2,6-dibenzhydryl substitution pattern in C also had a detrimental effect on activity. By contrast, this enhanced steric protection present in C had the effect of generating the highest-molecular-weight polyethylene, as these groups most effectively blocked chain transfer. Similarly for Ni2, the combination of benzhydryl and cyclohexyl as the ortho-substituents also served to effectively protect the metal center, leading to polyethylene of molecular weight that exceeds that produced using B1, B4 and B5.

2.3. Properties of Polyethylene

According to the melting point (Tm) data presented in Table 3, Table 4 and Table 5, polyethylenes exhibited values that fell, in most cases, below 100 °C ((Tm range: 62.8–103.4 °C (EtAlCl2); 51.9–87.8 °C (Et2AlCl); 49.4–91.1 °C (MAO)). Such temperatures are typical of branched polyethylene, and the variations influenced by their specific branching composition and polymer molecular weight. In order to gain more detailed information on polyethylene branching, we firstly explored the influence of reaction temperature and activator on the polymers generated using Ni2. Secondly, the mechanical properties of selected polymer samples were investigated.

2.3.1. Analysis of Polyethylene Branching with 13C NMR Spectroscopy

With the aim to gain some information on branching density and composition, the 13C NMR spectra of three representative samples, PE30E|Ni2 (run 2; Table 3), PE40D|Ni2 (run 2; Table 4) and PE40M|Ni2 (run 2; Table 5), were recorded. To allow suitable solubility, these polymer samples were dissolved in C6D4Cl2 at high temperature, and their spectra were recorded at 110 °C. On the basis of the chemical shift, the signal characteristics of the specific carbon environment within the branched polymer chain can be assigned and semi-quantified using methods documented in the literature [41,42,43,44,45]. The main results of this branching analysis are presented in Table 6, while the corresponding spectra are presented in Figure 4, Figure 5 and Figure 6.
Regarding the 13C NMR spectrum of PE30E|Ni2, the analysis revealed 73 branches per 1000 carbons, which comprised methyl (81.8%), ethyl (2.13%), propyl (4.79%), butyl (4.03%), pentyl (2.11%) and longer branched chains (5.17%) (Figure 4). Conversely, the spectrum of PE40D|Ni2 showed relatively high branching density, containing 104 branches per 1000 carbons, including methyl (81.1%), ethyl (1.18%), propyl (3.93%), butyl (3.54%), pentyl (2.97%) and longer branches (7.23%) (Figure 5). By comparing these two polymer samples, it is evident that the content of the longer chain branches increased from 5.17% to 7.23%. On the other hand, for PE40M|Ni2, the branching density of 99 branches per 1000 carbons was determined based on methyl (82.9%), ethyl (2.53%), propyl (6.01%), butyl (3.31%) and longer branched chains (5.29%) (Figure 6).
The above results clearly show that all three polyethylene samples contained a significant number of branches, which can be expected based on the chain-walking mechanism [46,47,48,49]. Moreover, both PE40D|Ni2 and PE40M|Ni2 showed higher branching density than PE30E|Ni2, but with a propensity for forming short-chain branches (>81%), which is corroborated by the relatively low Tm values. These methyl branches can be formed with a single-chain-walking process comprising β-H elimination, 2,1-insertion and ethylene coordination/insertion [31,34,50,51]. While the variation in aluminum activator may be a contributing factor in this, differences in polymerization run temperatures are likely responsible. Indeed, similar findings have been previously noted [50,51], whereby higher reaction temperatures increase the degree of branching of polyethylene.

2.3.2. Mechanical Performance of Branched Polyethylenes

To study the mechanical properties of the branched polyethylene produced with Ni2 at distinct polymerization run temperatures of 30 °C and 60 °C, six polyethylene samples (PE30E|Ni2, PE30D|Ni2, PE30M|Ni2, PE60E|Ni2, PE60D|Ni2 and PE60M|Ni2) prepared using EtAlCl2, Et2AlCl and MAO were selected, and their properties were examined.
In the first instance, tensile stress–strain data were measured for all six samples (Table 7). In order to maintain consistency among the tests, each sample was divided into five parts to allow us to perform multiple tests; the stress–strain curves are shown in Figure 7, Figure 8 and Figure 9. For the two samples prepared using EtAlCl2 as activator, the stress–strain curves are presented in Figure 7. PE30E|Ni2 was observed to exhibit higher ultimate tensile stress (6.98 MPa) than PE60E|Ni2 (4.29 MPa), with correspondingly higher crystallinity (Xc = 16.2% vs. 4.3%). By contrast, PE30E|Ni2 exhibited lower strain at break (εb = 353%) than PE60E|Ni2 (εb = 462%). Similar results were observed for the two samples prepared using Ni2/Et2AlCl (Figure 8), that is PE60D|Ni2 showed lower tensile stress (8.59 MPa) and crystallinity (Xc = 5.5%), as well as higher strain-at-break (εb = 861%), than PE30D|Ni2. This was also the case for the samples prepared using Ni2/MAO (Figure 8), among which PE30M|Ni2 exhibited a higher tensile stress (15.91 MPa) and crystallinity (22.3%) and lower strain at break (713%) than PE60M|Ni2.
Overall, these results indicate that crystallinity influenced the tensile properties of these polyethylenes [53], with high-crystallinity polyethylene displaying high ultimate tensile strength but low strain at break. In addition, when comparing the data among samples, the difference in the type of activator also influenced the mechanical properties. For example, the ultimate tensile stress of PE30E|Ni2 was 6.98 MPa, which compares to 15.91 MPa for PE30M|Ni2, while the strain at break value of 353% compared to 713%, respectively. Notably, when comparing these test data with those of previously reported polyethylene produced under similar conditions (i.e., Et2AlCl or MAO as activator, PC2H4 = 10 atm and run temp. = 60 °C), PE60E|Ni2 and PE60M|Ni2 exhibited higher values of tensile stress and strain at break [37,38,54].
Secondly, to evaluate the capacity for elastic recovery of these polyethylenes, the six samples were tested for hysteresis at 30 °C using dynamic mechanical analysis; the stress–strain response curves of each sample are shown in Figure 10, Figure 11 and Figure 12. The strain recovery value (SR) was calculated with the standard equation SR = 100 (εaεr)/εa (where εa is the applied strain and εr is the cyclic strain at zero load after 10 cycles). Each sample was tested over 10 cycles, and elastic recovery was set at approximately 80% of the maximum tensile strength. The inspection of Figure 10, Figure 11 and Figure 12 reveals that all samples displayed a constant level of recovery after the first cycle. By comparing the data of the samples produced at different run temperatures using the same activator, it is evident that the materials produced at a higher run temperature, 60 °C, had a higher elastic recovery rate (61.6% for PE60E|Ni2; 60.5% for PE60D|Ni2; 71.2% for PE60M|Ni2) than those produced at the lower run temperature of 30 °C (47.4% for PE30E|Ni2; 48.9% for PE30D|Ni2; 47.9% for PE30M|Ni2). In effect, these data indicate that apart from crystallinity, the operating temperature of the polymerization and in turn the polymer molecular weight had a significant effect on the elastic properties of polyethylene. As a further notable point, the elastic recovery properties of these polymer samples were higher than those reported in previous studies of nickel catalysts [38,55,56].

3. Materials and Methods

3.1. General Considerations

All reactions involving air- and moisture-sensitive compounds were conducted in a strict water- and oxygen-free environment (nitrogen purity > 99.9%). Under a nitrogen atmosphere (>99.9%), toluene was firstly dried over sodium/benzophenone ketyl and then distilled immediately prior to use. Methylaluminoxane (MAO; 1.45 M in toluene) and modified methylaluminoxane (MMAO; 2.50 M in heptane) were purchased from Anhui Botai Electronic Materials Co., Ltd (Anhui, China). Ethylaluminum dichloride (EtAlCl2), ethylaluminum sesquichloride (EASC) and diethylaluminum chloride (Et2AlCl) were bought from Lianli Chemical (Beijing, China), while high-purity ethylene was acquired from Beijing Yanshan Petrochemical Co (Beijing, China). Other reagents and solvents were purchased from Acros, Aldrich (Beijing, China) or other local suppliers and were used upon receipt. The FT-IR spectrum data of the complexes were recorded at room temperature with a Nicolet 6700 FT-IR spectrometer (Thermo fisher, Shanghai, China) in the region 4000–400 cm−1, and the relative intensity of the IR bands was described as w (weak), m (medium) or s (strong). The elemental analysis was performed with a FlashEA 1112 microanalyzer (Thermo Flash Smart, Beijing, China). The melting temperatures of the polyethylene samples were measured under nitrogen during the fourth scan with a PerkinElmer TA-Q2000 (Beijing, China) differential scanning calorimeter. The molecular weight (Mw) and dispersity (Mw/Mn) of the polymers were determined using an Agilent PL-GPC 220 instrument (Beijing, China) by employing 1,2,4-trichlorobenzene as the eluting solvent at an operating temperature of 160 °C. A Bruker AVANCE III 500 MHz instrument (Beijing, China) at 110 °C was used to record the 13C NMR spectra of the polyethylene samples. Sample preparation consisted in dissolving 60–80 mg of polyethylene samples in 1,2-dichlorobenzene-d4 (2 mL), using TMS as an internal standard. 2-Cyclopentyl-4,6-dibenzhydrylaniline, 2-cyclohexyl-4,6-dibenzhydrylaniline, 2-cyclooctyl-4,6-dibenzhydrylaniline and 2-cyclododecyl-4,6-dibenzhydrylaniline were prepared using the corresponding procedure detailed in the literature [57,58,59].

3.2. Synthesis of [ArN=C(Me)-C(Me)=NAr]NiBr2 (Ni1Ni4)

(a)
Ar = 2-(C5H9)-4,6-(CHPh2)2C6H2 (Ni1). 2,3-Butanedione (0.009 g, 0.10 mmol), 2-cyclopentyl-4,6-dibenzhydrylaniline (0.198 g, 0.40 mmol) and (DME)NiBr2 (0.031 g, 0.10 mmol) were stirred and heated to reflux in glacial acetic acid (15 mL) for 6 h. After cooling to room temperature, excess anhydrous diethyl ether was added to induce precipitation. The precipitate was then filtered, washed with diethyl ether (4 × 20 mL) and dried under reduced pressure to give Ni1 (0.021 g, 17%) as brown powder. FT-IR (cm−1): 2954(w), 2869(w), 1644(w, vC=N), 1578(m), 1494(m), 1448(m), 1418(w), 1340(w), 1265(w), 1215(w), 1186(w), 1163(w), 1077(w), 1031(w), 745(m), 699(s). Anal. Calcd for C78H72Br2N2Ni (1255.95): C, 74.59; H, 5.78; N, 2.23. Found: C, 74.23; H, 6.01; N, 2.04.
(b)
Ar = 2-(C6H11)-4,6-(CHPh2)2C6H2 (Ni2). Using a synthetic procedure similar to that described for Ni1 but using 2-cyclohexyl-4,6-dibenzhydrylaniline as the arylamine, Ni2 was isolated as brown powder (0.327 g, 64%). FT-IR (cm−1): 2921(w), 2850(w), 1637(w, vC=N), 1600(m), 1495(m), 1446(m), 1368(m), 1260(w), 1239(w), 1215(w), 1162(w), 1077(w), 1032(w), 776(w), 748(m), 699(s). Anal. Calcd for C80H76Br2N2Ni (1284.00): C, 74.83; H, 5.97; N, 2.18. Found: C, 74.67; H, 6.22; N, 1.94.
(c)
Ar = 2-(C8H15)-4,6-(CHPh2)2C6H2 (Ni3). Using a synthetic procedure similar to that described for Ni1 but using 2-cyclooctyl-4,6-dibenzhydrylaniline as the arylamine, Ni3 was isolated as brown powder (0.068 g, 20%). FT-IR (cm−1): 2919(w), 2857(w), 1644(w, vC=N), 1593(m), 1494(m), 1447(m), 1421(w), 1345(w), 1185(w), 1078(w), 1030(w), 914(w), 849(w), 744(m), 698(s). Anal. Calcd for C84H84Br2N2Ni (1340.11): C, 75.29; H, 6.32; N, 2.09. Found: C, 74.98; H, 6.44; N, 1.83.
(d)
Ar = 2-(C12H23)-4,6-(CHPh2)2C6H2 (Ni4). Using a similar synthetic procedure to that described for Ni1 but using 2-cyclododecyl-4,6-dibenzhydrylaniline as the arylamine, Ni4 was isolated as brown powder (0.072 g, 25%). FT-IR (cm−1): 2920(w), 2858(w), 1639(w, vC=N), 1594(m), 1494(m), 1445(m), 1373(w), 1258(w), 1212(w), 1161(w), 1075(w), 1031(w), 914(w), 847(w), 743(m), 697(s). Anal. Calcd for C92H100Br2N2Ni (1452.33): C, 76.09; H, 6.94; N, 1.93. Found: C, 75.83; H, 7.03; N, 1.77.

3.3. X-ray Diffraction Studies

Single crystals of Ni2 and Ni4 were grown by diffusing diethyl ether into a dichloromethane solution of each complex. Selected crystals of each were mounted on an XtaLAB SynergyR single-crystal diffractometer. Cell parameters were obtained by performing the global refinement of the positions of all collected reflections with monochromatic Cu-Kα radiation (λ = 1.54184 Å) at a temperature of 170.01(10) K. The intensity was corrected for Lorentz and polarization effects, as well as empirical absorption. The structure was determined using the SHELXT (Sheldrick, 2015) [60] direct method (and Patterson’s method) and refined with full matrix least squares for F2 based on SHELXL (Sheldrick, 2015) [61]. All hydrogens were placed in the calculated positions, and all non-hydrogen atoms were refined with anisotropic displacement parameters. The Squeeze option in the crystallographic program PLATON was applied to remove disordered solvents from the structure, and possible twin structures were searched with TwinRotMat [62,63]. The graphical representation of the molecular structure of both complexes was generated with the ORTEP application [64]. X-ray structure determination and detailed data are provided in Table S1.

3.4. Polymerization Experiments

Ethylene polymerization runs were carried out in a stainless-steel, 250 mL capacity autoclave equipped with a mechanical stirrer and a temperature controller. The autoclave was evacuated and backfilled with nitrogen three times and then with ethylene once. Freshly distilled toluene (25 mL) was injected into the autoclave under an ethylene atmosphere. After the temperature had stabilized, a solution of the precatalyst (1 μmol) in toluene (50 mL) was added, followed by a pre-determined amount of activator and finally some more toluene (25 mL). At the designated ethylene pressure, the reaction mixture was stirred (400 rpm/min) for the selected time. Once the run was completed, the reactor was allowed to cool to room temperature. The pressure within the vessel was slowly released, and the contents were acidified with ethanol to quench the polymerization. After a period of stirring, the polymer was collected using filtration and washed with ethanol. The product was dried under reduced pressure at 60 °C and then weighed.

4. Conclusions

In summary, four α-diimine–nickel(II) complexes bearing four distinct ortho-cycloalkyl substituents, namely, cyclopentyl (Ni1), cyclohexyl (Ni2), cyclooctyl (Ni3) and cyclododecyl (Ni4), were successfully prepared using a simple one-pot synthetic route. All the complexes were well characterized, and in the cases of Ni2 and Ni4, this was performed additionally using single-crystal X-ray diffraction. When activated with either EtAlCl2, Et2AlCl or MAO, Ni1Ni4 displayed moderate to very high activity as catalysts for the polymerization of ethylene, forming polyethylene with high molecular weight, high branching density and a distinct preference for short-chain methyl branches (>80%). In particular, cyclohexyl-containing Ni2, following activation with MAO, was the most productive of this work, exhibiting a peak activity of 13.2 × 106 g(PE) of (mol of Ni)−1 h−1 at 40 °C. Furthermore, this MAO-activated nickel catalyst formed the highest-molecular-weight polymer (9.68 × 105 g mol−1) when compared with those obtained using Ni2/EtAlCl2 (7.51 × 105 g mol−1) and Ni2/Et2AlCl (8.85 × 105 g mol−1) at the same run temperature. Overall, the ring size of the ortho-cycloalkyl group displayed a distinct effect on catalyst activity, with Ni2 (cyclohexyl) > Ni1 (cyclopentyl) > Ni4 (cyclododecyl) > Ni3 (cyclooctyl). By contrast, its correlation with polymer molecular weight was less clear and dependent on the aluminum activator employed. Nonetheless, all catalysts showed good control, with narrow dispersity being a feature of all polymeric materials produced. In terms of the mechanical properties of these macromolecules, it was evident that the degree of crystallinity (Xc) and molecular weight (Mw, linked to run temperature) were influential factors in tensile strength, strain at break and elastic recovery. Further studies to explore the applications of these thermoplastic elastomers (TPEs) are ongoing.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124852/s1, Table S1: Crystallographic parameters for Ni2 and Ni4. Figures S1–S18: GPC traces of the of polyethylenes and DSC thermograms of PE30E|Ni2, PE30D|Ni2, PE30M|Ni2, PE60E|Ni2, PE60D|Ni2 and PE60M|Ni2. In addition, crystallographic information on Ni2 (CCDC 2262894) and Ni4 (CCDC 2262895) (CIF) are provided in the supporting information.

Author Contributions

Design of the study, W.-H.S.; synthesis of organic compounds and nickel complexes, S.J. and Y.Z.; synthesis of cycloalkyl-substituted anilines, I.V.O. and I.I.O.; characterization of compounds, S.J., Z.Y., M.L., Y.M., G.A.S. and W.-H.S.; X-ray study, M.L., S.J. and T.L.; catalytic study, S.J., M.L. and Y.M.; characterization of polyethylenes, S.J., Y.Z., M.L. and Y.M.; writing and editing, S.J., Z.Y.; G.A.S., I.I.O. and W.-H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was self-funded and received no governmental funds.

Data Availability Statement

Not applicable.

Acknowledgments

G.A.S. thanks the Chinese Academy of Sciences for a President’s International Fellowship for Visiting Scientists.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Knowles, W.S.; Sabacky, M.J. Catalytic asymmetric hydrogenation employing a soluble, optically active, rhodium complex. Chem. Commun. 1968, 22, 1445–1446. [Google Scholar] [CrossRef]
  2. Ostoja Starzewski, K.A.; Witte, J. Control of the molecular weight of polyethene in syntheses with bis(ylide)nickel catalysts. Angew. Chem. Int. Ed. 1987, 26, 63–64. [Google Scholar] [CrossRef]
  3. Mu, H.; Pan, L.; Song, D.; Li, Y. Neutral Nickel Catalysts for Olefin Homo- and Copolymerization: Relationships between Catalyst Structures and Catalytic Properties. Chem. Rev. 2015, 115, 12091–12137. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, C. Designing catalysts for olefin polymerization and copolymerization: Beyond electronic and steric tuning. Nat. Rev. Chem. 2018, 2, 6–14. [Google Scholar] [CrossRef]
  5. Johnson, L.K.; Killian, C.M.; Brookhart, M. New Pd(II)- and Ni(II)-Based Catalysts for Polymerization of Ethylene and alpha-Olefins. J. Am. Chem. Soc. 1995, 117, 6414–6415. [Google Scholar] [CrossRef]
  6. Walsh, D.J.; Hyatt, M.G.; Miller, S.A.; Guironnet, D. Recent Trends in Catalytic Polymerizations. ACS Catal. 2019, 9, 11153–11188. [Google Scholar] [CrossRef]
  7. Mecking, S. Olefin Polymerization by Late Transition Metal Complexes-A Root of Ziegler Catalysts Gains New Ground. Angew. Chem. Int. Ed. 2001, 40, 534–540. [Google Scholar] [CrossRef]
  8. Gansukh, B.; Zhang, R.; Guo, J.; Zhang, W.; Sun, W.-H. Paradox of Late Transition-Metal Catalysts in Ethylene Polymerization. Gen. Chem. 2020, 6, 190031. [Google Scholar] [CrossRef]
  9. Dai, S.; Li, S.; Xu, G.; Wu, C.; Liao, Y.; Guo, L. Flexible cycloalkyl substituents in insertion polymerization with α-diimine nickel and palladium species. Polym. Chem. 2020, 11, 1393–1400. [Google Scholar] [CrossRef]
  10. Ittel, S.D.; Johnson, L.K.; Brookhart, M. Late-Metal Catalysts for Ethylene Homo- and Copolymerization. Chem. Rev. 2000, 100, 1169–1203. [Google Scholar] [CrossRef]
  11. Guan, Z.; Cotts, P.M.; McCord, E.F.; McLain, S.J. Chain Walking: A New Strategy to Control Polymer Topology. Science 1999, 283, 2059–2062. [Google Scholar] [CrossRef] [PubMed]
  12. Anderson, W.C., Jr.; Rhinehart, L.; Tennyson, A.G.; Long, B.K. Redox-Active Ligands: An Advanced Tool to Modulate Polyethylene Microstructure. J. Am. Chem. Soc. 2016, 138, 774–777. [Google Scholar] [CrossRef] [PubMed]
  13. Gao, W.; Xin, L.; Hao, Z.Q.; Li, G.D.; Su, J.-H.; Zhou, L.J.; Mu, Y. The ligand redox behavior and role in 1,2-bis[(2,6-diisopropylphenyl)imino]-acenaphthene nickel-TMA(MAO) systems for ethylene Polymerization. Chem. Commun. 2015, 51, 7004–7007. [Google Scholar] [CrossRef]
  14. Wang, F.; Chen, C. A continuing legend: The Brookhart-type α-diimine nickel and palladium catalysts. Polym. Chem. 2019, 10, 2354–2369. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, Z.; Liu, Q.; Solan, G.A.; Sun, W.-H. Recent advances in Ni-mediated ethylene chain growth: Nimine-donor ligand effects on catalytic activity, thermal stability and oligo-/polymer structure. Coord. Chem. Rev. 2017, 350, 68–83. [Google Scholar] [CrossRef]
  16. Mahmood, Q.; Li, X.; Qin, L.; Wang, L.; Sun, W.-H. Structural evolution of iminopyridine support for nickel/palladium catalysts in ethylene (oligo)polymerization. Dalton Trans. 2022, 51, 14375–14407. [Google Scholar] [CrossRef]
  17. Gates, D.P.; Svejda, S.A.; Oňate, E.; Killian, C.M.; Johnson, L.K.; White, P.S.; Brookhart, M. Synthesis of Branched Polyethylene Using (α-Diimine)nickel(II) Catalysts:  Influence of Temperature, Ethylene Pressure, and Ligand Structure on Polymer Properties. Macromolecules 2000, 33, 2320–2334. [Google Scholar] [CrossRef]
  18. Rhinehart, J.L.; Brown, L.A.; Long, B.K. A Robust Ni (II) α-Diimine catalyst for high temperature ethylene polymerization. J. Am. Chem. Soc. 2013, 135, 16316–16319. [Google Scholar] [CrossRef]
  19. Jia, D.; Zhang, W.; Liu, W.; Wang, L.; Redshaw, C.; Sun, W.-H. Unsymmetrical α-diiminonickel bromide complexes: Synthesis, characterization and their catalytic behavior toward ethylene. Catal. Sci. Technol. 2013, 3, 2737–2745. [Google Scholar] [CrossRef]
  20. Fan, L.; Du, S.; Guo, C.-Y.; Hao, X.; Sun, W.-H. 1-(2,6-dibenzhydryl-4-fluorophenylimino)-2-aryliminoacenaphthylylnickel halides highly polymerizing ethylene for the polyethylenes with high branches and molecular weights. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 1369–1378. [Google Scholar] [CrossRef]
  21. Wang, L.; Liu, M.; Mahmood, Q.; Yuan, S.; Li, X.; Qin, L.; Zou, S.; Liang, T.; Sun, W.-H. Enhancing the Thermal Stability of α-Diimine Ni(II) Catalysts for ethylene polymerization via Symmetrical axial shielding strategy. Eur. Polym. J. 2023, 194, 112112. [Google Scholar] [CrossRef]
  22. Gao, M.; Du, S.; Ban, Q.; Xing, Q.; Sun, W.-H. Ethylene polymerization by 2,3-diiminobutylnickel bromide pre-catalysts bearing remote benzhydryl substituents. J. Organomet. Chem. 2015, 798, 401–407. [Google Scholar] [CrossRef]
  23. Sun, J.; Wang, F.; Li, W.; Chen, M. Ligand steric effects on α-diimine nickel catalyzed ethylene and 1-hexene polymerization. RSC Adv. 2017, 7, 55051–55059. [Google Scholar] [CrossRef] [Green Version]
  24. Guo, L.; Zhu, D.; Zhang, W.; Zada, M.; Solan. G. A. Remote dibenzocycloheptyl-substitution of an iminotrihydroquinoline-nickel catalyst as a route to narrowly dispersed branched polyethylene waxes with alkene functionality. Eur. Polym. J. 2018, 107, 315–328. [Google Scholar] [CrossRef] [Green Version]
  25. Lin, W.; Liu, M.; Xu, L.; Ma, Y.; Zhang, L.; Flisak, Z.; Hu, X.; Liang, T.; Sun. W.-H. Nickel (II) complexes with sterically hindered 5,6,7-trihydroquinoline derivatives selectively dimerizing ethylene to 1-butene. Appl. Organomet. Chem. 2022, 36, 6596. [Google Scholar] [CrossRef]
  26. Liu, Q.; Zhang, W.; Jia, D.; Hao, X.; Redshaw, C.; Sun. W.-H. 2-{2,6-Bis[bis(4-fluorophenyl)methyl]-4-chlorophenylimino}-3-aryliminobutylnickel(II) bromide complexes: Synthesis, characterization, and investigation of their catalytic behavior. Appl. Catal. A Gen. 2014, 475, 195–202. [Google Scholar] [CrossRef]
  27. Tempel, D.J.; Johnson, L.K.; Huff, R.L.; White, P.S.; Brookhart, M. Mechanistic studies of Pd(II)-α-diimine-catalyzed olefin polymerizations. J. Am. Chem. Soc. 2000, 122, 6686–6700. [Google Scholar] [CrossRef]
  28. Mahmood, Q.; Zeng, Y.; Wang, X.; Sun, Y.; Sun, W.-H. Advancing polyethylene properties by incorporating NO2 moiety in 1,2-bis(arylimino)acenaphthylnickel precatalysts: Synthesis, characterization and ethylene polymerization. Dalton Trans. 2017, 46, 6934–6947. [Google Scholar] [CrossRef]
  29. Wang, F.; Yuan, J.; Song, F.; Li, J.; Jia, Z.; Yuan. B. New chiral α-diimine nickel(II) complexes bearing ortho-sec-phenethyl groups for ethylene polymerization. Appl. Organomet. Chem. 2013, 27, 319–327. [Google Scholar] [CrossRef]
  30. Chen, Y.; Du, S.; Huang, C.; Solan, G.A.; Hao, X.; Sun, W.-H. Balancing high thermal stability with high activity in diaryliminoacenaphthene-nickel(II) catalysts for ethylene polymerization. J. Polym. Sci. Part A Polym. Chem. 2017, 55, 1971–1983. [Google Scholar] [CrossRef]
  31. Liu, H.; Zhao, W.; Yu, J.; Yang, W.; Hao, X.; Redshaw, C.; Chen, L.; Sun, L. Synthesis, characterization and ethylene polymerization behavior of nickel dihalide complexes bearing bulky unsymmetrical α-diimine ligands. Catal. Sci. Technol. 2012, 2, 415–422. [Google Scholar] [CrossRef]
  32. Xia, J.; Kou, S.; Zhang, Y.; Jian, Z. Strategies cooperation on designing nickel catalysts to access ultrahigh molecular weight polyethylenes. Polymer 2022, 240, 124478. [Google Scholar] [CrossRef]
  33. Yuan, S.; Fan, Z.; Zhang, Q.; Flisak, Z.; Ma, Y.; Sun, Y.; Sun, W.-H. Enhancing performance of α-diiminonickel precatalyst for ethylene polymerization by substitution with the 2,4-bis (4,4’-dimethoxybenzhydryl)-6-methylphenyl group. Appl. Organomet. Chem. 2020, 34, e5638. [Google Scholar] [CrossRef]
  34. Wang, Y.; Vignesh, A.; Qua, M.; Wang, Z.; Sun, Y.; Sun, W.-H. Access to polyethylene elastomers via ethylene homo-polymerization using N, N′-nickel (II) catalysts appended with electron withdrawing difluorobenzhydryl group. Eur. Polym. J. 2019, 117, 254–271. [Google Scholar] [CrossRef]
  35. Du, S.; Kong, S.; Shi, Q.; Mao, J.; Guo, C.; Yi, J.; Liang, T.; Sun, W.-H. Enhancing the Activity and Thermal Stability of Nickel complex precatalysts Using 1-[2,6-bis(bis(4-fluorophenyl)methyl)-4-methylphenylimino]-2-aryliminoacenaphthylene Derivatives. Organometallics 2015, 34, 582–590. [Google Scholar] [CrossRef]
  36. Zada, M.; Vignesh, A.; Guo, L.; Zhang, R.; Zhang, W.; Ma, Y.; Sun, Y.; Sun, W.-H. Sterically and Electronically Modified Aryliminopyridyl-Nickel Bromide Precatalysts for an Access to Branched Polyethylene with Vinyl/Vinylene End Groups. ACS Omega 2020, 5, 10610–10625. [Google Scholar] [CrossRef]
  37. Yuan, S.; Duan, T.; Zhang, R.; Solan, G.A.; Ma, Y.; Liang, T.; Sun, W.-H. Alkylaluminum activator effects on polyethylene branching using a N,N′-nickel precatalyst appended with bulky 4,4′-dimethoxybenzhydryl groups. Appl. Organomet. Chem. 2019, 33, e4785. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, R.; Wang, Z.; Ma, Y.; Solan, G.A.; Sun, Y.; Sun, W.-H. Plastomeric-like polyethylenes achievable using thermally robust N,N’-nickel catalysts appended with electron withdrawing difluorobenzhydryl and nitro groups. Dalton Trans. 2019, 48, 1878–1891. [Google Scholar] [CrossRef] [Green Version]
  39. Zada, M.; Guo, L.; Zhang, R.; Zhang, W.; Ma, Y.; Solan, G.A.; Sun, Y.; Sun, W.-H. Moderately branched ultra-high molecular weight polyethylene by using N,N′-nickel catalysts adorned with sterically hindered dibenzocycloheptyl groups. Appl. Organomet. Chem. 2019, 33, e4749. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, M.; Zhang, R.; Ma, Y.; Han, M.; Solan, G.A.; Yang, W.; Liang, T.; Sun, W.-H. Trifluoromethoxy-substituted nickel catalysts for producing highly branched polyethylenes: Impact of solvent, activator and N,N’-ligand on polymer properties. Polym. Chem. 2022, 13, 1040–1058. [Google Scholar] [CrossRef]
  41. Cotts, P.M.; Guan, Z.; McCord, E.; McLain, S. Novel Branching Topology in Polyethylenes As Revealed by Light Scattering and 13C NMR. Macromolecules 2000, 33, 6945–6952. [Google Scholar] [CrossRef]
  42. Jurkiewicz, A.; Eilerts, N.W.; Hsieh, E.T. 13C NMR Characterization of Short Chain Branches of Nickel Catalyzed Polyethylene. Macromolecules 1999, 32, 5471–5476. [Google Scholar] [CrossRef]
  43. Kunrath, F.A.; Mota, F.F.; Casagrande, O.L., Jr.; Mauler, R.S.; de Souza, R.F. Synthesis and Characterization of Hyperbranched Polyethylenes Made with Nickel-α-Diimine Catalysts. Macromol. Chem. Phys. 2002, 203, 2407–2411. [Google Scholar] [CrossRef]
  44. Giraudeau, P.; Baguet, E. Improvement of the inverse-gated decoupling sequence for a faster quantitative analysis of various samples by 13C NMR spectroscopy. J. Magn. Reson. 2006, 180, 110–117. [Google Scholar] [CrossRef]
  45. Galland, G.B.; Souza, R.F.; Mauler, R.S.; Nunes, F.F. 13C NMR Determination of the Composition of Linear Low-Density Polyethylene Obtained with [η3-Methallyl-nickel-diimine]PF6 Complex. Macromolecules 1999, 32, 1620–1625. [Google Scholar] [CrossRef]
  46. Leatherman, M.D.; Svejda, S.A.; Johnson, L.K.; Brookhart, M. Mechanistic Studies of Nickel(II) Alkyl Agostic Cations and Alkyl Ethylene Complexes:  Investigations of Chain Propagation and Isomerization in (α-diimine)Ni(II)-Catalyzed Ethylene Polymerization. J. Am. Chem. Soc. 2003, 125, 3068–3081. [Google Scholar] [CrossRef]
  47. Michalak, A.; Ziegler, T. Polymerization of Ethylene Catalyzed by a Nickel(+2) Anilinotropone-Based Catalyst: DFT and Stochastic Studies on the Elementary Reactions and the Mechanism of Polyethylene Branching. Organometallics 2003, 22, 2069–2079. [Google Scholar] [CrossRef]
  48. Falivene, L.; Wiedemann, T.; Göttker-Schnetmann, I.; Caporaso, L.; Cavallo, L.; Mecking, S. Control of Chain Walking by Weak Neighboring Group Interactions in Unsymmetrical Catalysts. J. Am. Chem. Soc. 2018, 140, 1305–1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Pei, L.; Liu, F.; Liao, H.; Gao, J.; Zhong, L.; Gao, H.; Wu, Q. Synthesis of Polyethylenes with Controlled Branching with α-Diimine Nickel Catalysts and Revisiting Formation of Long-Chain Branching. ACS Catal. 2018, 8, 1104–1113. [Google Scholar] [CrossRef]
  50. Camacho, D.H.; Salo, E.V.; Ziller, J.W.; Guan, Z. Cyclophane-Based Highly Active Late-Transition-Metal Catalysts for Ethylene Polymerization. Angew. Chem. Int. Ed. 2004, 43, 1821–1825. [Google Scholar] [CrossRef] [PubMed]
  51. Meinhard, D.; Wegner, M.; Kipiani, G.; Hearley, A.; Reuter, P.; Fischer, S.; Marti, O.; Rieger, B. New Nickel(II) Diimine Complexes and the Control of Polyethylene Microstructure by Catalyst Design. J. Am. Chem. Soc. 2007, 129, 9182–9191. [Google Scholar] [CrossRef]
  52. Meiries, S.; Speck, K.; Cordes, D.B.; Slawin, A.M.Z.; Nolan, S.P. [Pd(IPr*OMe)(acac)Cl]: Tuning the N-Heterocyclic Carbene in Catalytic C-N Bond Formation. Organometallics 2013, 32, 330–339. [Google Scholar] [CrossRef]
  53. He, Z.; Liang, Y.; Yang, W.; Uchino, H.; Yu, J.; Sun, W.-H.; Han, C.C. Random hyperbranched linear polyethylene: One step production of thermoplastic elastomer. Polymer 2015, 56, 119–122. [Google Scholar] [CrossRef]
  54. Wu, R.; Wang, Y.; Zhang, R.; Guo, C.-Y.; Flisak, Z.; Sun, Y.; Sun, W.-H. Finely tuned nickel complexes as highly active catalysts affording branched polyethylene of high molecular weight: 1-(2,6-Dibenzhydryl-4-methoxyphenylimino)-2-(arylimino)acenaphthylenenickel halides. Polymer 2018, 153, 574–586. [Google Scholar] [CrossRef]
  55. Vignesh, A.; Zhang, Q.; Ma, Y.; Liang, T.; Sun, W.-H. Attaining highly branched polyethylene elastomers by employing modified α-diiminonickel(II) catalysts: Probing the effects of enhancing fluorine atom on the ligand framework towards mechanical properties of polyethylene. Polymer 2020, 187, 122089. [Google Scholar] [CrossRef]
  56. Suo, H.; Oleynik, I.V.; Huang, C.; Oleynik, I.I.; Solan, G.A.; Ma, Y.; Liang, T.; Sun, W.-H. ortho-Cycloalkyl substituted N,N′-diaryliminoacenaphthene-Ni(ii) catalysts for polyethylene elastomers; exploring ring size and temperature effects. Dalton Trans. 2017, 46, 15684–15697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. 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-Cycloalkyland 2,6-Dicycloalkylanilines. Rus. J. Gen. Chem. 2004, 74, 1423–1427. [Google Scholar] [CrossRef]
  58. Ivanchev, S.S.; Yakimansky, A.V.; Rogozin, D.G. Quantum-chemical calculations of the effect of cycloaliphatic groups in α-diimine and bis(imino)pyridine ethylene polymerization precatalysts on their stabilities with respect to deactivation reactions. Polymer 2004, 45, 6453–6459. [Google Scholar] [CrossRef]
  59. Chartoire, A.; Claver, C.; Corpet, M.; Krinsky, J.; Mayen, J.; Nelson, D.; Nolan, S.P.; Penafiel, I.; Woodward, R.; Meadows, R.E. Recyclable NHC Catalyst for the Development of a Generalized Approach to Continuous Buchwald-Hartwig Reaction and Workup. Org. Process Res. Dev. 2016, 20, 551–557. [Google Scholar] [CrossRef] [Green Version]
  60. Sheldrick, G.M. SHELXT-Integrated space-group and crystal structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  61. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, 65, 148–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Van der Sluis, P.; Spek, A.L. BYPASS: An effective method for the refinement of crystal structures containing disordered solvent regions. Acta Crystallogr. Sect. A Found. Adv. 1990, 46, 194–201. [Google Scholar] [CrossRef]
  64. Farrugia, L.J. ORTEP-3 for Windows-a version of ORTEPIII with a Graphical User Interface (GUI). J. Appl. Crystallogr. 1997, 30, 565. [Google Scholar] [CrossRef]
Chart 1. α-Diarylimine–nickel(II) (A) and its N,N-bis(aryl)butane-2,3-diimine derivatives (B and C) along with the target of this work, ortho-cycloalkyl-containing Ni1Ni4.
Chart 1. α-Diarylimine–nickel(II) (A) and its N,N-bis(aryl)butane-2,3-diimine derivatives (B and C) along with the target of this work, ortho-cycloalkyl-containing Ni1Ni4.
Molecules 28 04852 ch001
Scheme 1. One-pot synthesis of complexes Ni1Ni4.
Scheme 1. One-pot synthesis of complexes Ni1Ni4.
Molecules 28 04852 sch001
Figure 1. ORTEP plot of Ni2 with thermal ellipsoids shown at the 30% probability level. All hydrogen atoms are omitted for clarity.
Figure 1. ORTEP plot of Ni2 with thermal ellipsoids shown at the 30% probability level. All hydrogen atoms are omitted for clarity.
Molecules 28 04852 g001
Figure 2. ORTEP plot of Ni4 with thermal ellipsoids shown at the 30% probability level. All hydrogen atoms are omitted for clarity.
Figure 2. ORTEP plot of Ni4 with thermal ellipsoids shown at the 30% probability level. All hydrogen atoms are omitted for clarity.
Molecules 28 04852 g002
Figure 3. Comparison of the activity and molecular weight of polyethylene produced with Ni2 with three examples of B, as well as C, under comparable polymerization conditions (Et2AlCl as activator, run temp. = 50 °C and PC2H4 = 10 atm).
Figure 3. Comparison of the activity and molecular weight of polyethylene produced with Ni2 with three examples of B, as well as C, under comparable polymerization conditions (Et2AlCl as activator, run temp. = 50 °C and PC2H4 = 10 atm).
Molecules 28 04852 g003
Figure 4. 13C NMR spectrum of PE30E|Ni2 (run 2; Table 3) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Figure 4. 13C NMR spectrum of PE30E|Ni2 (run 2; Table 3) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Molecules 28 04852 g004
Figure 5. 13C NMR spectrum of PE40D|Ni2 (run 2; Table 4) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Figure 5. 13C NMR spectrum of PE40D|Ni2 (run 2; Table 4) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Molecules 28 04852 g005
Figure 6. 13C NMR spectrum of PE40M|Ni2 (run 2; Table 5) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Figure 6. 13C NMR spectrum of PE40M|Ni2 (run 2; Table 5) along with a segment of the assigned polyethylene backbone; the spectrum was recorded in C6D4Cl2 at 110 °C.
Molecules 28 04852 g006
Figure 7. Stress–strain curves of PE30E|Ni2 and PE60E|Ni2; the vertical lines indicate sample fracture points.
Figure 7. Stress–strain curves of PE30E|Ni2 and PE60E|Ni2; the vertical lines indicate sample fracture points.
Molecules 28 04852 g007
Figure 8. Stress–strain curves of PE30D|Ni2 and PE60D|Ni2; the vertical lines indicate the sample fracture points.
Figure 8. Stress–strain curves of PE30D|Ni2 and PE60D|Ni2; the vertical lines indicate the sample fracture points.
Molecules 28 04852 g008
Figure 9. Stress–strain curves of PE30M|Ni2 and PE60M|Ni2; the vertical lines indicate the sample fracture points.
Figure 9. Stress–strain curves of PE30M|Ni2 and PE60M|Ni2; the vertical lines indicate the sample fracture points.
Molecules 28 04852 g009
Figure 10. Hysteresis loops for PE30E|Ni2 and PE60E|Ni2 obtained by performing 10 cycles at 30 °C.
Figure 10. Hysteresis loops for PE30E|Ni2 and PE60E|Ni2 obtained by performing 10 cycles at 30 °C.
Molecules 28 04852 g010
Figure 11. Hysteresis loops for PE30D|Ni2 and PE60D|Ni2 obtained by performing 10 cycles at 30 °C.
Figure 11. Hysteresis loops for PE30D|Ni2 and PE60D|Ni2 obtained by performing 10 cycles at 30 °C.
Molecules 28 04852 g011
Figure 12. Hysteresis loops for PE30M|Ni2 and PE60M|Ni2 obtained by performing 10 cycles at 30 °C.
Figure 12. Hysteresis loops for PE30M|Ni2 and PE60M|Ni2 obtained by performing 10 cycles at 30 °C.
Molecules 28 04852 g012
Table 1. Selected bond lengths (Å) and angles (°) of Ni2 and Ni4.
Table 1. Selected bond lengths (Å) and angles (°) of Ni2 and Ni4.
Ni2Ni4
Bond lengths (Å)
Ni1-N12.0221(11)2.014(5)
Ni1-N22.0081(10)2.004(5)
Ni1-Br12.3331(3)2.3304(13)
Ni1-Br22.3418(3)2.3361(13)
C3-N21.2840(17)1.285(8)
C2-N11.2856(17)1.280(8)
Bond angles (°)
N1-Ni1-N281.36(4)80.6(2)
N1-Ni1-Br1113.49(3)100.68(15)
N1-Ni1-Br2109.52(3)121.18(15)
N2-Ni1-Br1112.24(3)121.51(15)
N2-Ni1-Br2108.12(3)99.74(15)
Br1-Ni1-Br2123.970(11)125.57(5)
Table 2. Preliminary evaluation of Ni2 as a precatalyst for ethylene polymerization using five different types of aluminum-alkyl activator a.
Table 2. Preliminary evaluation of Ni2 as a precatalyst for ethylene polymerization using five different types of aluminum-alkyl activator a.
RunActivatorAl:NiT (°C)PE (g)Activity bMwcMw/MncTm (°C) d
1MAO2000303.577.1410.12.592.96
2MMAO2000303.376.749.182.989.75
3Et2AlCl500305.0310.19.982.575.28
4EtAlCl2500305.1510.38.222.980.59
5EASC500300.951.907.612.979.59
a Reaction conditions: 1 μmol Ni2, PC2H4 = 10 atm and 100 mL toluene. b ×106 g(PE) mol−1 (Ni) h −1. c Measured using GPC, Mw in units of ×105 g mol−1. d Measured using DSC.
Table 3. Ethylene polymerization results obtained using Ni1Ni4; EtAlCl2 was used as activator a.
Table 3. Ethylene polymerization results obtained using Ni1Ni4; EtAlCl2 was used as activator a.
RunPrecat.Al:NiT (°C)t (min)PE (g)Activity bMwcMw/MncTm (°C) d
1Ni250020303.877.7412.52.9104.57
2Ni250030305.1510.38.222.980.59
3Ni250040304.849.687.512.870.10
4Ni250050304.629.245.662.755.44
5Ni250060303.306.604.642.746.27
6Ni250070301.132.263.892.644.54
7Ni230030301.953.905.712.476.78
8Ni240030303.286.565.743.168.45
9Ni260030304.869.727.082.990.43
10Ni270030304.569.126.682.973.68
11Ni25003050.9110.93.102.277.26
12Ni250030152.6110.46.482.182.04
13Ni250030457.119.4810.12.882.24
14Ni250030609.249.2412.12.272.26
15 eNi250030301.212.426.101.965.07
16 fNi250030300.210.423.521.933.75
17Ni150030304.308.605.352.6103.41
18Ni350030300.130.264.372.285.62
19Ni450030301.623.245.902.162.82
a Reaction conditions: 1 μmol nickel precatalyst, PC2H4 = 10 atm and 100 mL toluene. b ×106 g(PE) mol−1 (Ni) h −1. c Measured using GPC, Mw in units of ×105 g mol−1. d Measured using DSC. e PC2H4 = 5 atm. f PC2H4 = 1 atm.
Table 4. Ethylene polymerization results obtained using Ni1Ni4; Et2AlCl was used as activator a.
Table 4. Ethylene polymerization results obtained using Ni1Ni4; Et2AlCl was used as activator a.
RunPrecat.Al:NiT (°C)t (min)PE (g)Activity bMwcMw/MncTm (°C) d
1Ni250030305.0310.19.982.575.28
2Ni250040305.3310.78.852.563.72
3Ni250050303.587.167.972.455.77
4Ni250060303.336.665.502.243.99
5Ni250070303.206.405.092.238.77
6Ni230040301.623.246.302.467.54
7Ni240040304.619.227.003.363.67
8Ni260040304.879.745.562.662.21
9Ni270040302.775.545.432.665.86
10Ni25004050.647.683.062.273.52
11Ni250040153.4113.64.802.257.56
12Ni250040455.947.9210.12.366.89
13Ni250040606.436.4310.92.566.44
14 eNi250040300.400.803.081.855.00
15 fNi250040300.180.362.791.524.27
16Ni150040302.064.123.692.187.80
17Ni350040300.020.046.262.472.35
18Ni450040300.150.305.292.451.86
a Reaction conditions: 1 μmol nickel precatalyst, PC2H4 = 10 atm and 100 mL toluene. b ×106 g(PE) mol−1 (Ni) h −1. c Measured using GPC, Mw in units of ×105 g mol−1. d Measured using DSC. e PC2H4 = 5 atm. f PC2H4 = 1 atm.
Table 5. Ethylene polymerization results obtained using Ni1Ni4; MAO was used as activator a.
Table 5. Ethylene polymerization results obtained using Ni1Ni4; MAO was used as activator a.
RunPrecat.Al:NiT (°C)t (min)PE (g)Activity bMwcMw/MncTm (°C) d
1Ni2200030303.577.1410.12.592.99
2Ni2200040306.6013.29.682.066.47
3Ni2200050305.7811.67.952.154.58
4Ni2200060303.917.827.242.143.40
5Ni2200070303.276.545.751.936.67
6Ni2150040304.759.509.332.070.24
7Ni2175040305.8511.710.42.269.31
8Ni2225040306.2112.410.92.467.72
9Ni2250040306.0012.09.582.164.40
10Ni220004051.1513.87.241.966.03
11Ni2200040153.3513.47.281.964.97
12Ni2200040458.5711.48.692.063.71
13Ni22000406010.710.77.782.062.05
14 eNi2200040301.563.127.871.954.73
15 fNi2200040300.310.623.531.517.35
16Ni1200040302.895.788.222.091.10
17Ni3200040300.260.5215.02.078.19
18Ni4200040302.094.1812.22.149.43
a Reaction conditions: 1 μmol nickel precatalyst, PC2H4 = 10 atm and 100 mL toluene. b ×106 g(PE) mol (Ni) −1 h −1. c Measured using GPC, Mw in units of ×105 g mol−1. d Measured using DSC. e PC2H4 = 5 atm. f PC2H4 = 1 atm.
Table 6. Branching analysis of the three polyethylene samples a.
Table 6. Branching analysis of the three polyethylene samples a.
SampleBranches/1000 CsMe/%Et/%Pr/%Bu/%Am/%Longer Branches/%Methyl
(1,4-Paired)
Methyl
(1,6-Paired)
PE30E|Ni27381.82.134.794.032.115.177.556.69
PE40D|Ni210481.11.173.933.542.977.239.749.55
PE40M|Ni29982.92.536.013.3105.298.9212.9
a Data obtained were determined using 13C NMR spectroscopy and the assignment made by the method described in ref. [41,42,43,44,45].
Table 7. Stress–strain properties of PE samples obtained using Ni2.
Table 7. Stress–strain properties of PE samples obtained using Ni2.
SampleTm (°C) aMwbXc (%) cStress (MPa) dStrain (%) d
PE30E|Ni280.598.2216.26.98353
PE30D|Ni275.289.9815.610.72612
PE30M|Ni292.9910.122.315.91713
PE60E|Ni246.274.646.34.29462
PE60D|Ni243.995.505.58.59861
PE60M|Ni243.407.245.19.02764
a Measured using DSC. b Mw (×105 g mol−1), determined using GPC. c Xc = ΔHf (Tm)/ [ΔHf° (Tm°)] [52]; ΔHf° (Tm°) = 248.3 J g−1. d Determined using a universal testing instrument.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiang, S.; Zheng, Y.; Oleynik, I.V.; Yu, Z.; Solan, G.A.; Oleynik, I.I.; Liu, M.; Ma, Y.; Liang, T.; Sun, W.-H. N,N-Bis(2,4-Dibenzhydryl-6-cycloalkylphenyl)butane-2,3-diimine–Nickel Complexes as Tunable and Effective Catalysts for High-Molecular-Weight PE Elastomers. Molecules 2023, 28, 4852. https://doi.org/10.3390/molecules28124852

AMA Style

Jiang S, Zheng Y, Oleynik IV, Yu Z, Solan GA, Oleynik II, Liu M, Ma Y, Liang T, Sun W-H. N,N-Bis(2,4-Dibenzhydryl-6-cycloalkylphenyl)butane-2,3-diimine–Nickel Complexes as Tunable and Effective Catalysts for High-Molecular-Weight PE Elastomers. Molecules. 2023; 28(12):4852. https://doi.org/10.3390/molecules28124852

Chicago/Turabian Style

Jiang, Shu, Yuting Zheng, Irina V. Oleynik, Zhixin Yu, Gregory A. Solan, Ivan I. Oleynik, Ming Liu, Yanping Ma, Tongling Liang, and Wen-Hua Sun. 2023. "N,N-Bis(2,4-Dibenzhydryl-6-cycloalkylphenyl)butane-2,3-diimine–Nickel Complexes as Tunable and Effective Catalysts for High-Molecular-Weight PE Elastomers" Molecules 28, no. 12: 4852. https://doi.org/10.3390/molecules28124852

Article Metrics

Back to TopTop