Synthesis of Branched Polyethylene via Bulky α-Diimine Nickel(II)-Catalyzed Ethylene Chain-Walking Polymerization

Abstract

The catalysis of olefin polymerization through the chain-walking process is a subject of great interest. In this contribution, the successful synthesis of a Brookhart-type unsymmetrical α-diimine nickel catalyst Ni, which contains both dibenzhydryl and phenyl groups, was determined by X-ray crystallography. The compound has a pseudo-tetrahedral geometry at the Ni center, showing pseudo-C2-symmetry. Upon activation with modified methylaluminoxane (MMAO), Ni1 exhibits high catalytic activity up to 1.02 × 107 g PE (mol Ni h)−1 toward ethylene polymerization, enabling the synthesis of high molecular weight branched polyethylene. The molecular weights and branching densities could be tuned over a very wide range. The polymerization results indicated the possibility of precise microstructure control, depending on the polymerization temperature. The branching densities were decreased with increasing the polymerization temperature.

Share and Cite:

Dong, Z. , Li, P. , Xu, G. and Wang, F. (2024) Synthesis of Branched Polyethylene via Bulky α-Diimine Nickel(II)-Catalyzed Ethylene Chain-Walking Polymerization. Open Journal of Organic Polymer Materials, 13, 1-12. doi: 10.4236/ojopm.2024.131001.

1. Introduction

In 1995, a significant discovery was made by Brookhart and his colleagues [1], demonstrating the high activity of α-diimine palladium and nickel catalysts in ethylene polymerization [2], as well as their capability to generate polymers with high molecular weight. This particular type of α-diimine catalyst has been widely recognized as the “Brookhart catalyst” and has attracted significant attention over the past two decades [3]-[8]. Chain-walking polymerization can give polyolefins with unique structures which cannot be obtained by common vinyl polymerization [9] [10]. The highly branched polyethylene, characterized by a significant presence of methyl and alkyl branches [11] [12] [13], exhibits an amorphous nature, whereas the chain-straightened polyethylene with its distinctive slow chain-walking mechanism demonstrates semi-crystalline behavior. Furthermore, the Brookhart catalyst demonstrated exceptional efficacy in facilitating the copolymerization of ethylene, propylene, and methyl acrylate, thus representing a significant breakthrough in addressing challenges associated with polar monomers [14] [15] [16] [17] [18].

We have previously reported that a series of bulky nickel catalysts with systematically varied ligand sterics for ethylene polymerization [19] [20] [21], the molecular weights and branching densities could be tuned over a very wide range [22] [23] [24]. The present study focuses on the synthesis of an unsymmetrical α-diimine nickel complex featuring dibenzhydryl and phenyl groups, as well as the investigation into how ligand structure and polymerization conditions influence ethylene polymerization.

2. Experimental Section

2.1. General Considerations

All manipulations were performed under nitrogen gas using standard Schlenk techniques. Research grade ethylene and propylene were purified by passing it through a deoxygenation and a dry columns. Methylene chloride and o-dichlorobenzene were pre-dried with 4 Å molecular sieves and distilled from CaH2 under dry nitrogen. Toluene, hexane, diethyl ether and 1,2-dimethoxyethane (DME) were distilled from sodium/benzophenone under nitrogen atmosphere and distilled before use. MMAO were donated by Tosoh-Finechem. Complex Ni' [25] was prepared according to reported procedures. Other chemicals were commercially obtained and purified with common procedures.

1H and 13C NMR spectra were recorded with a Bruker Ascend 400 spectrometer at ambient temperature unless otherwise stated. The chemical shifts of the 1H and 13C NMR spectra were referenced to tetramethylsilane (TMS). Elemental analysis was performed by the Analytical Center of the Changzhou University. X-ray diffraction data were collected at 298(2) K on a Bruker Smart CCD area detector with graphite-monochromated MoKα radiation (λ = 0.71073 Å). Gel permeation chromatography (GPC) was carried out at 150˚C by using a PL-GPC 220 high-temperature gel permeation chromatography system. 1,2,4-Trichlorobenzene (TCB) was used as the solvent at a flow rate of 1.0 mL min–1, and the system was calibrated by using a polystyrene standard and are corrected for linear polyethylene by universal calibration by using the Mark–Houwink parameters of Rudin: K = 1.75 × 10–2 cm3 g–1 and R = 0.67 for polystyrene and K = 5.90 × 10–2 cm3 g–1 and R = 0.69 for polyethylene.

2.2. Synthesis and Characterizations

Synthesis of (2,4-dibenzhydryl-6-phenylphenylimino)-butanone L'

A solution of 2,6-bis(diphenylmethyl)-4-phenylaniline (20 mmol) [26], 2,3-butadione (100 mmol) and Formic acid (1mL) in methyl alcohol (80 mL) was stirred at 60˚C for 12 h, until there was one main point on the TLC plate. The solvent was partially evaporated under reduced pressure until the formation of a light yellow solid (9.10 g, 80%). 1H NMR (400 MHz, CDCl3, ppm): δ 7.24 - 6.98 (m, 24H), 6.94 - 6.86 (m, 2H), 6.73 (t, J = 7.6 Hz, 1H), 5.45 (d, 2H, -CHPh2), 2.43 (s, 3H, -CH3), 0.72 (s, 3H, -CH3).

Synthesis of [N-(2,4-dibenzhydryl-6-phenylphenyl)-N'-(2,6-dimethylphenyl)]-2,3 -butadiene L

A mixture of L1' (5.69 g, 10 mmol), aniline (1.86 g, 10 mmol), and a catalytic amount of p-toluenesulfonic acid in 150 mL toluene was refluxed for 24 h. The solution was evaporated at reduced pressure, and the remaining solution was diluted in methanol (300 mL). The yellow solid was isolated by filtration, followed by recrystallization from dichloromethane and methanol (5.50g, 95%). 1H NMR (400 MHz, CDCl3, ppm): δ 7.27 (t, J = 7.6 Hz, 4H), 7.24 - 7.04 (m, 20H), 7.02 - 6.96 (m, 4H), 6.87 (t, J = 7.6 Hz, 1H), 6.74 (t, J = 7.6 Hz, 1H), 5.56 (s, 2H, -CHPh2), 5.48 (s, 2H, -CHPh2), 1.92 (s, 6H, -CH3), 1.70 (s, 3H, -CH3), 0.93 (s, 3H, -CH3). 13C NMR (100 MHz, CDCl3, ppm): δ 168.77 (C=N), 167.34 (C=N), 130.09, 129.71, 129.56, 129.37, 128.38, 128.15, 127.98, 127.75, 126.55, 126.23, 124.53, 123.12, 56.54 (-CHPh2), 52.83 (-CHPh2), 17.91 (-CH3), 17.53 (-CH3), 16.19 (-CH3), 15.75 (-CH3). Anal. Calcd. for C50H44N2 (672.92): C, 89.25; H, 6.59; N, 4.16. Found: C, 89.29; H, 6.63; N, 4.10. FT-IR (KBr): 1649 cm1 (vCN).

Synthesis of {[N-(2,4-dibenzhydryl-6-phenylphenyl)-N'-(2,6-dimethylphenyl)]-2,3-butadiene}dibromonickel Ni

A mixture of the ligand L (1 mmol), (DME)NiBr2 (308 mg, 1 mmol) in CH2Cl2 (20 mL) was stirred for 2 hours at room temperature, Then concentrate the solvent, add ether to precipitate the solid catalyst, filter and clear with ether for three to five times, and vacuum dry the solid to obtain the catalyst Ni (93%, 5.50 g). Anal. Calcd. for C50H44Br2N2Ni (891.42): C, 67.37; H, 4.98; N, 3.14. Found: C, 67.41; H, 4.93; N, 3.17. FT-IR (KBr): 1643 cm1 (vC=N).

2.3. X-Ray Structure Determination

Single crystal of ligand for X-ray analysis were obtained by dissolving the nickel complex in CH2Cl2, followed by slow layering of the resulting solution with at room temperature. Data collections were performed at 150 K on a Bruker SMART APEX diffractometer with a CCD area detector, using graphite monochromated Moradiation (λ = 0.71073 Å). The determination of crystal class and unit cell parameters was carried out by the SMART program package. The raw frame data were processed using SAINT and SADABS to yield the reflection data file. The structures were solved by using the SHELXTL program. Refinement was performed on F2 anisotropically for all non-hydrogen atoms by the full-matrix least-squares method. The hydrogen atoms were placed at the calculated positions and were included in the structure calculation without further refinement of the parameters.

2.4. Ethylene Polymerization

Ethylene polymerization was performed in a 100-mL glass reactor equipped with a magnetic stirrer. After drying the reactor under N2 atmosphere, toluene was added to the reactor. The solvent was then saturated with a prescribed ethylene pressure. The co-catalyst (MMAO) was added in Al/Ni molar ratios to the reactor via a syringe, the solution was thermostated to the desired temperature and allowed to equilibrate for 10 min. Then, the catalyst solution in toluene was added to the reactor. The polymerization, conducted under 1.2 atm of ethylene pressure, was terminated with 200 mL of a 3% HCl-MeOH solution. The polymers obtained were adequately washed with methanol and dried under vacuum at 50˚C for 6 h.

Analysis of the polyethylene branching by 1H NMR spectroscopy [27]: branching density, branches/1000C = (CH3/3)/[(CH + CH2 + CH3)/2] × 1000. CH3 (alkyl methyl, alk-CH3, m, 0.70 - 0.95 ppm), CH2 and CH (alk-CH and alk-CH2, m, ca. 1.00 - 1.45 ppm) refer to the intensities of the methyl, methylene and methine resonances in 1H NMR spectra.

3. Results and Discussion

3.1. Synthesis and Characterization of the Nickel Complex

The monoimine ligand L' (2,4-dibenzhydryl-6-phenylphenylimino)-butanone was prepared from the reaction of 2,4-bis(diphenylmethyl)-6-methylaniline [26] with 5 equiv of 2,3-butadione at 80% yield on multigram scale (Scheme 1). Subsequently, the reaction with 1 equiv of the corresponding aniline led to the formation of the α-diimine ligand L at 95% yields. The reaction of ligand L with 1.1 equiv of (DME)NiBr2 in CH2Cl2 afforded the desired nickel complex Ni at 93% yield (Scheme 1). Compounds were characterized by 1H and 13C NMR spectroscopies, and elemental analysis. In addition, the classic catalyst Ni' [25] was also used for comparison in this study.

Scheme 1. Synthesis of α-diimine ligand L and its nickel complex Ni.

3.2. X-Ray Crystallographic Studies

Furthermore, to confirm the molecular structure of catalyst the single crystal of Ni1 was also grown by slow diffusion of n-hexane in dichloromethane, and their molecular structures were confirmed by single-crystal X-ray diffraction analysis (Figure 1). X-ray diffraction data of single crystal was collected at 100 K on a Bruker Smart CCD area detector with graphite monochromated MoKα radiation (λ = 0.71073 Å). Selected bond lengths (Å) and angles (˚) for Ni are listed in Table 1. Crystal data, data collection and refinement parameters are listed in Table 2.

Figure 1. Molecular structure of Ni1 at 30% probability ellipsoids (CCDC 2339033). Hydrogen atoms were omitted for clarity.

Table 1. Bond lengths (Å) and angles (˚) for Ni.

Br01—Ni03

2.3573 (10)

Cl1—C52

1.730 (15)

Br02—Ni03

2.3317 (11)

N006—C18

1.455 (7)

Ni03—N006

2.009 (5)

N006—C40

1.291 (7)

Ni03—N007

1.996 (5)

N007—C44

1.445 (7)

Cl00—C52

1.664 (15)

N007—C41

1.291 (7)

Br02—Ni03—Br01

117.60 (4)

C44—N007—Ni03

124.2 (4)

N006—Ni03—Br01

122.19 (14)

C41—N007—Ni03

114.5 (4)

N006—Ni03—Br02

107.21 (13)

C41—N007—C44

121.3 (5)

N007—Ni03—Br01

109.04 (13)

C19—C18—N006

117.5 (5)

N007—Ni03—Br02

114.86 (14)

C17—C18—N006

120.7 (5)

N007—Ni03—N006

80.50 (19)

C17—C18—C19

121.7 (5)

C18—N006—Ni03

122.9 (4)

C45—C44—N007

118.5 (5)

C40—N006—Ni03

115.0 (4)

C50—C44—N007

118.7 (6)

C40—N006—C18

121.9 (5)

C44—N007—Ni03

124.2 (4)

Table 2. Crystal data and structure refinement for Ni.

Identification code

Ni1

CCDC

2339033

Empirical formula

C50H44Br2N2Ni·CH2Cl2

Formula weight

976.33

Temperature/K

150

Crystal system

Monoclinic

Space group

P2(1)/n

a/Å

9.1499(19)

b/Å

35.806(7)

c/Å

13.852(3)

α/˚

90

β

99.890(7)

γ

90

Volume/Å3

4470.8(15)

Z

4

ρcalc g/cm3

1.451

μ/mm1

2.378

F(000)

1992.0

Crystal size/mm3

0.07 × 0.05 × 0.04

Radiation

MoKα (λ = 0.71073)

2Θ range for data collection/˚

3.752 to 52.96

Index ranges

−11 ≤ h ≤ 11, −44 ≤ k ≤ 44, −14 ≤ l ≤ 17

Reflections collected

28200

Independent reflections

9008 [Rint = 0.1014, Rsigma = 0.1019]

Data/restraints/parameters

9008/521/527

Goodness-of-fit on F2

1.028

Final R indexes [I ≥ 2σ (I)]

R1 = 0.0702, wR2 = 0.1603

Final R indexes [all data]

R1 = 0.1194, wR2 = 0.1860

Ni-CH2Cl2 has a pseudo-tetrahedral geometry at the Ni center, showing pseudo-C2-symmetry. The crystals exhibit an asymmetrical α-diimine structure and encapsulate a CH2Cl2 molecule. Bond lengths and angles are within the expected range for α-diimine, for example, the bond length of N006-C40 [1.291(7) Å] and N007-C41 [1.291(7) Å] have typical imine double bonds character. Its structure is similar to those reported in the literature for a [NiBr2(α-diimine)] compound characterized by X-ray diffraction, {bis[N,N'-(2,4,6-trimethylphenyl)imino]acenaphthene}dibromonickel Ni' [25]. In fact, the Ni–N bond distances in complex Ni (2.005 and 1.996 Å) are similar to those determined for Ni' (2.021 Å), and the Ni-Br bond distances in complexes Ni and Ni' are almost identical (2.3573 and 2.3317 Å for Ni' vs. 2.323 Å for Ni'). In addition, the N-Ni-Br angles (112.96˚ for complex C1) are also approximate to those for complex C7 (114.4˚). Specifically, the Br01-Ni03-Br02 angle in Ni is more open and measures 117.60(4)˚, while the bite angles of N007-Ni03-N006 are 80.50(19)˚. This asymmetrical dibenzhydryl and phenyl substituted α-diimine ligand may induce a coordination effect in olefin polymerization [9].

3.3. Ethylene Polymerization Studies

Polymerization of ethylene with Ni activated by MMAO were carried out at various polymerization temperatures and the [Al]/[Ni] ratio for 10 min under 6 atm of ethylene, and the results are listed in Table 3. At room temperature, the influence of the [Al]/[Ni] with MMAO was investigated by increasing the [Al]/[Ni] molar ratio from 200 to 800 (entries 1 - 3, Table 1). The highest activity of 6.45 × 106 g PE/(mol Ni h) and the highest molecular weight of the polymers were achieved with the [Al]/[Ni] ratio of 500 (entry 3, Table 1).

Table 3. Effect of catalyst and temperature on ethylene polymerizationa.

Entry

Precat.

[Al]/[Ni]

Temp (˚C)

Yield (g)

Activityb

Mnc

(× 104)

Mw/Mnc

Bd

1

Ni

200

RT

2.41

5.16

35.1

2.13

-

2

Ni

500

RT

3.01

6.45

38.6

2.27

-

3

Ni

800

RT

2.83

6.06

37.5

2.64

-

4

Ni

500

0

2.77

5.94

20.3

1.97

39

5

Ni

500

25

3.85

8.25

51.7

2.27

46

6

Ni

500

50

4.74

10.16

48.9

2.39

59

7

Ni

500

75

3.58

7.67

37.1

2.55

68

8

Ni

500

100

2.15

4.61

23.0

2.69

76

9

Ni'

500

50

0.78

1.67

9.4

2.47

81

aPolymerization conditions: Ni = 2.8 μmol in CH2Cl2 (2 mL); cocatalyst MMAO; solvent = toluene (30 mL); 6 atm of ethylene; time = 10 min. b106 g of PE (mol of Ni)−1 h−1. cMn and Mw/Mn determined by GPC, 104 g mol−1. dBranching numbers per 1000C were determined by 1H NMR.

The influence of the polymerization temperature was studied by varying the temperature from 0˚C to 100˚C (entries 4 - 8, Table 1). The maximum catalytic activities of complex Ni1 was observed at 50˚C (entry 6, Table 1), and the polymerization at 100˚C still gave high catalytic activities on the level of 106 g of PE/(mol Ni h) (entry 8, Table 1).

The molecular weights of the polymers were also investigated at various polymerization temperatures from 0˚C to 100˚C (entries 4 - 8, Table 1). The polymer obtained by Ni exhibited its highest molecular weight at 25˚C (entry 5, Table 1). Subsequent temperature increases resulted in a decrease in the molecular weight accompanied by broadening of the molecular weight distribution. The observation implies that rapid chain transfer occurs at elevated temperatures [9].

Steric effect of ortho-position in anilinic moiety can be evaluated by comparing Ni and Ni' (entries 6 and 9, Table 1). Ni1-MMAO exhibited higher activity and much higher thermal stability than the corresponding methyl substituted Ni' (entries 6 vs 9, Table 1), which can be attributed to the steric effect caused by the presence of dibenzhydryl and phenyl groups in the ortho-position [9]. The significantly higher molecular weight of 4.89 × 105 g mol1 obtained by Ni (Figure 2), as compared to that from Ni', suggests that the presence of bulky dibenzhydryl and phenyl ortho-substituents on the ligand's aryl rings greatly enhances the rate of chain propagation, effectively suppressing chain-transfer reactions [28].

Figure 2. GPC traces for the PEs obtained with Ni1-MMAO and Ni'-MMAO at 25˚C (entries 6 and 9, Table 1).

The branching densities of the polyethylenes obtained were determined using 1H NMR spectroscopy [29] [30] [31]. The branching densities of 39 - 76 branches per 1000C were increased with polymerization temperature from 0˚C to 100˚C (entries 4 - 8, Table 1). In addition, the branching densities were reduced by increasing the steric bulk of the ortho-substituents on the α-diimine ligand, as evidenced by a comparison between Ni and Ni' at 50˚C (entries 6 vs 9, Table 1). The observed disparity in microstructure and lower degree of branching density for the sterically bulkier catalysts may be attributed to a relatively higher propensity for ethylene insertion into primary metal alkyl species compared to secondary metal alkyl species [9].

The branching structures analysis based on 13C NMR [20] showed that 56 methyl, 5 ethyl, 3 n-propyl, 2 n-butyl, 1 sec-butyl and 10 longer chains (>C4 branches) exist for the polyethylene produced by complex Ni at 100˚C (Figure 3(i), entry 8, Table 1). In contrast, only 39 methyl branches were observed for Ni at 0˚C (Figure 3(ii), entry 4, Table 1).

Figure 3. 13C NMR spectrum of polyethylenes obtained by complex Ni at 0˚C (ii) and 100˚C (i) (A and B refer to methyl carbon of sec-butyl branches, entries 4 and 8, Table 1).

4. Conclusion

In conclusion, an unsymmetrical α-diimine nickel catalyst Ni bearing dibenzhydryl and phenyl groups was prepared and investigated in ethylene polymerization. The catalytic activity of Ni activated by MMAO is significantly enhanced compared to the classic nickel catalyst Ni' with methyl substitution, thereby facilitating the production of high molecular weight branched polyethylene. Moreover, the branching densities and molecular weights can be finely adjusted over a wide range. The branching densities decrease as the polymerization temperature increases. The chain-walking polymerization mechanism is responsible for the formation of major types of branches (methyl, ethyl, propyl, butyl, and longer chains) in the resulting polyethylenes.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant Nos 21801002), the Overseas Students Innovation and Entrepreneurship Support Program Project of Anhui Province (2021LCX022), the Key R&D Projects in Anhui Province (2022i01020012), the Natural Science Foundation of Hefei (2022039), Excellent Research and Innovation Team Project of Anhui Province (2022AH010001).

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Johnson, L.K., Killian, C.M. and Brookhart, M. (1995) New Pd(ii)-and Ni(ii)-Based Catalysts for Polymerization of Ethylene and .α.-Olefins. Journal of the American Chemical Society, 117, 6414-6415.
https://doi.org/10.1021/ja00128a054
[2] Johnson, L.K., Mecking, S. and Brookhart, M. (1996) Copolymerization of Ethylene and Propylene with Functionalized Vinyl Monomers by Palladium(ii) Catalysts. Journal of the American Chemical Society, 118, 267-268.
https://doi.org/10.1021/ja953247i
[3] Tan, C. and Chen, C. (2019) Emerging Palladium and Nickel Catalysts for Copolymerization of Olefins with Polar Monomers. Angewandte Chemie, 131, 7268-7276.
https://doi.org/10.1002/ange.201814634
[4] Mu, H., Zhou, G., Hu, X. and Jian, Z. (2021) Recent Advances in Nickel Mediated Copolymerization of Olefin with Polar Monomers. Coordination Chemistry Reviews, 435, Article ID: 213802.
https://doi.org/10.1016/j.ccr.2021.213802
[5] Chen, C. (2018) Designing Catalysts for Olefin Polymerization and Copolymerization: Beyond Electronic and Steric Tuning. Nature Reviews Chemistry, 2, 6-14.
https://doi.org/10.1038/s41570-018-0003-0
[6] Zhang, Y., Zhang, Y., Hu, X., Wang, C. and Jian, Z. (2022) Advances on Controlled Chain Walking and Suppression of Chain Transfer in Catalytic Olefin Polymerization. ACS Catalysis, 12, 14304-14320.
https://doi.org/10.1021/acscatal.2c04272
[7] Chen, Z. and Brookhart, M. (2018) Exploring Ethylene/Polar Vinyl Monomer Copolymerizations Using Ni and Pd α-Diimine Catalysts. Accounts of Chemical Research, 51, 1831-1839.
https://doi.org/10.1021/acs.accounts.8b00225
[8] Guo, L., Dai, S., Sui, X. and Chen, C. (2015) Palladium and Nickel Catalyzed Chain Walking Olefin Polymerization and Copolymerization. ACS Catalysis, 6, 428-441.
https://doi.org/10.1021/acscatal.5b02426
[9] Wang, F. and Chen, C. (2019) A Continuing Legend: The Brookhart-Type α-Diimine Nickel and Palladium Catalysts. Polymer Chemistry, 10, 2354-2369.
https://doi.org/10.1039/c9py00226j
[10] Vignesh, A., Zhang, Q., Ma, Y., Liang, T. and Sun, W. (2020) 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, 187, Article ID: 122089.
https://doi.org/10.1016/j.polymer.2019.122089
[11] Fang, J., Sui, X., Li, Y. and Chen, C. (2018) Synthesis of Polyolefin Elastomers from Unsymmetrical α-Diimine Nickel Catalyzed Olefin Polymerization. Polymer Chemistry, 9, 4143-4149.
https://doi.org/10.1039/c8py00725j
[12] Tian, S., Zhang, Y., Li, R., Wang, F. and Li, W. (2019) Cationic Para-Benzhydryl Substituted α-Diimine Nickel Catalyzed Ethylene and 1-Decene Polymerizations via Controllable Chain-walking. Inorganica Chimica Acta, 486, 492-498.
https://doi.org/10.1016/j.ica.2018.11.004
[13] Wu, J., Li, P., Zhang, J., Dong, Z., Li, W., Bashir, M.S., Li, Y. and Wang, F. (2023) Chain-Walking Polymerization of Ethylene and 1-Octene with Ortho-Phenyl-Based α-Diimine Ni (II) Catalysts. Applied Organometallic Chemistry, 37, e7246.
https://doi.org/10.1002/aoc.7246
[14] Tan, C., Chen, M. and Chen, C. (2023) ‘Catalyst + X’ Strategies for Transition Metal-Catalyzed Olefin-Polar Monomer Copolymerization. Trends in Chemistry, 5, 147-159.
https://doi.org/10.1016/j.trechm.2022.12.007
[15] Zou, C., Tan, C. and Chen, C. (2023) Heterogenization Strategies for Nickel Catalyzed Synthesis of Polyolefins and Composites. Accounts of Materials Research, 4, 496-506.
https://doi.org/10.1021/accountsmr.2c00263
[16] Zou, C., Si, G. and Chen, C. (2022) A General Strategy for Heterogenizing Olefin Polymerization Catalysts and the Synthesis of Polyolefins and Composites. Nature Communications, 13, Article No. 1954.
https://doi.org/10.1038/s41467-022-29533-9
[17] Zhang, H., Zou, C., Zhao, H., Cai, Z. and Chen, C. (2021) Hydrogen-bonding-induced Heterogenization of Nickel and Palladium Catalysts for Copolymerization of Ethylene with Polar Monomers. Angewandte Chemie, 133, 17586-17591.
https://doi.org/10.1002/ange.202106682
[18] Liang, T., Goudari, S.B. and Chen, C. (2020) A Simple and Versatile Nickel Platform for the Generation of Branched High Molecular Weight Polyolefins. Nature Communications, 11, Article No. 372.
https://doi.org/10.1038/s41467-019-14211-0
[19] Yuan, J., Song, F., Li, J., Jia, Z., Wang, F. and Yuan, B. (2013) Chiral {bis[N, N’-(2-Ethyl-6-Methylphenyl)Imino]-1, 2-Dimethylethane} Dibromonickel Bearing Sec-Phenethyl Groups: Synthesis, Characterization and Application in Synthesis of Nanosized Dendritic Polyethylene. Inorganica Chimica Acta, 400, 99-106.
https://doi.org/10.1016/j.ica.2013.02.009
[20] Wang, F., Tanaka, R., Cai, Z., Nakayama, Y. and Shiono, T. (2016) Synthesis of Highly Branched Polyolefins Using Phenyl Substituted α-Diimine Ni(ii) Catalysts. Polymers, 8, Article 160.
https://doi.org/10.3390/polym8040160
[21] Wang, F., Tanaka, R., Li, Q., Nakayama, Y. and Shiono, T. (2018) Chain-walking Polymerization of Linear Internal Octenes Catalyzed by α-Diimine Nickel Complexes. Organometallics, 37, 1358-1367.
https://doi.org/10.1021/acs.organomet.8b00042
[22] Wang, F., Tian, S., Li, R., Li, W. and Chen, C. (2017) Ligand Steric Effects on Naphthyl-α-Diimine Nickel Catalyzed α-Olefin Polymerization. Chinese Journal of Polymer Science, 36, 157-162.
https://doi.org/10.1007/s10118-018-2038-6
[23] Yu, F., Li, P., Xu, M., Xu, G., Na, Y., Zhang, S., et al. (2021) Iminopyridyl Ligands Bearing Polyethylene Glycol Unit for Nickel Catalyzed Ethylene Polymerization. Polymer, 229, Article ID: 124023.
https://doi.org/10.1016/j.polymer.2021.124023
[24] Li, P., Liu, H., Tian, W., Ma, Z., Wang, X., Xu, G., et al. (2023) Steric and Electronic Effects in Cationic Pyridine Carboxamidate Nickel Mediated Ethylene Polymerization and Copolymerization with Methyl 10-undecenoate. Polymer, 280, Article ID: 126060.
https://doi.org/10.1016/j.polymer.2023.126060
[25] Liu, J., Li, Y., Li, Y. and Hu, N. (2008) Ethylene Polymerization by (α-Diimine)Nickel(ii) Complexes Bearing Different Substituents on para-Position of Imines Activated with Mmao. Journal of Applied Polymer Science, 109, 700-707.
https://doi.org/10.1002/app.28134
[26] Dai, S. and Li, S. (2020) Effect of Aryl Orientation on Olefin Polymerization in Iminopyridyl Catalytic System. Polymer, 200, Article ID: 122607.
https://doi.org/10.1016/j.polymer.2020.122607
[27] Jenkins, J.C. and Brookhart, M. (2002) A Highly Active Anilinoperinaphthenone-Based Neutral Nickel(ii) Catalyst for Ethylene Polymerization. Organometallics, 22, 250-256.
https://doi.org/10.1021/om020648f
[28] Wang, Z., Liu, Q., Solan, G.A. and Sun, W. (2017) Recent Advances in Ni-Mediated Ethylene Chain Growth: Nimine-Donor Ligand Effects on Catalytic Activity, Thermal Stability and Oligo-/Polymer Structure. Coordination Chemistry Reviews, 350, 68-83.
https://doi.org/10.1016/j.ccr.2017.06.003
[29] Leone, G., Canetti, M., Pierro, I., Zanchin, G., De Rosa, C., Ricci, G., et al. (2019) (micro)structure, Thermal Behavior and Mechanical Properties of Ethylene-Propylene-1-Octadecene Terpolymers from Chain-Walking Polymerization of 1-Octadecene. Polymer, 166, 27-37.
https://doi.org/10.1016/j.polymer.2019.01.029
[30] Rhinehart, J.L., Brown, L.A. and Long, B.K. (2013) A Robust Ni(ii) α-Diimine Catalyst for High Temperature Ethylene Polymerization. Journal of the American Chemical Society, 135, 16316-16319.
https://doi.org/10.1021/ja408905t
[31] Wang, Y., Nan, C., Zhuo, W., Zou, C., Jiang, H., Hao, X., et al. (2023) Amine-Imine Nickel Catalysts with Pendant O-Donor Groups for Ethylene (Co)Polymerization. Inorganic Chemistry, 62, 5105-5113.
https://doi.org/10.1021/acs.inorgchem.2c04240.

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.