Next Article in Journal
Angle-Independent Color Change in Thermoresponsive Gel-Immobilized Colloidal Amorphous Film Attached to PET Substrate
Next Article in Special Issue
Parallel Catalyst Synthesis Protocol for Accelerating Heterogeneous Olefin Polymerization Research
Previous Article in Journal
Modifying Effect and Mechanism of Polymer Powder on the Properties of Asphalt Binder for Engineering Application
Previous Article in Special Issue
Synthesis of Degradable Polyolefins Bearing Disulfide Units via Metathesis Copolymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Behavior of Cobalt Complexes Bearing Pyridine–Oxime Ligands in Isoprene Polymerization

1
School of Chemistry and Chemical Engineering, Linyi University, Linyi 276000, China
2
College of Agriculture and Forestry, Linyi University, Linyi 276000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2023, 15(24), 4660; https://doi.org/10.3390/polym15244660
Submission received: 13 November 2023 / Revised: 29 November 2023 / Accepted: 6 December 2023 / Published: 10 December 2023
(This article belongs to the Special Issue Catalytic Olefin Polymerization and Polyolefin Materials)

Abstract

:
Several cobalt(II) complexes Co1Co3 bearing pyridine–oxime ligands (L1 = pyridine-2-aldoxime for Co1; L2 = 6-methylpyridine-2-aldoxime for Co2; L3 = phenyl-2-pyridylketoxime for Co3) and picolinaldehyde O-methyl oxime (L4)-supported Co4 were synthesized and well characterized by FT-IR, mass spectrum and elemental analysis. The single-crystal X-ray diffraction of complex Co2 reveals that the cobalt center of CoCl2 is coordinated with two 6-methylpyridine-2-aldoxime ligands binding with Npyridine and Noxime atoms, which feature a distorted octahedral structure. These Co complexes Co1Co4 displayed extremely high activity toward isoprene polymerization upon activation with small amount of AlClEt2 in toluene, giving polyisoprene with high activity up to 16.3 × 105 (mol of Co)−1(h)−1. And, the generated polyisoprene displayed high molecular weights and narrow molecular distribution with a cis-1,4-enriched selectivity. The type of cobalt complexes, cocatalyst and reaction temperature all have effects on the polymerization activity but not on the microstructure of polymer.

Graphical Abstract

1. Introduction

The development of novel catalysts for diene polymerization with high activity and selectivity has attracted considerable attention in the past decades for improved synthetic rubber [1,2]. Isoprene is one of the most popular monomers, playing an increasingly important role in the synthetic rubber industry ever since being isolated from natural rubber by Williams in 1860. The stereospecific polymerization of isoprene allows the following stereoregular polymers to be given, including cis-1,4-, trans-1,4-, 3,4-, 1,2-unit, which exhibit different mechanical properties and application direction due to the otherness of chain parameters. For example, cis-1,4- polyisoprene has a microstructure similar to natural rubber and can be used as an alternative to natural rubber. The 3,4- polyisoprene is generally applied in the tire industry to improve the wet skid resistance of tread stock. Hence, developing well-performed cis-1,4-alt-3,4- polyisoprene materials with various component contents and sequence structure is one of the main goals of rubber industry [3,4].
Transition metal (Sc, Nd, Ti, V, Cr, Fe, Co, Ni, Pd, etc.)-catalyzed diene polymerization is a successful technique invention, and has achieved a significant breakthrough for the field of synthetic rubber with good performances [5,6,7,8]. Ever since the Ziegler–Natta catalysts were invented, the research has mainly focused on simple metal salts (Ti [9], Nd [10], and Co [11] with halide, acetonate, etc.) in combination with alkyl aluminum. To improve the activity of the transition metal catalyst, a variety of ligands containing pyridine, imine, phosphine and phenolate moieties have been applied to modify the electronic and steric environment at the metal’s center. And, it has been proven that the ancillary ligand plays an important role in the catalytic process and polymer properties [12,13,14,15,16,17,18]. Therefore, much research has been focused on cobalt catalyst, which is known as the most versatile catalyst among the various transition metals, and has characteristics of low cost, high stability and easy preparation. And, these cobalt catalysts show good catalytic effects in regulating polyolefin microstructure, molecular weight and molecular weight distribution. At first, several cobalt salts (e.g., CoCl2 and Co(acac)2) combined with suitable alkylating agents were used to catalyze diene polymerization with disappointing catalytic effects [6,19,20,21]. Subsequently, efforts to develop well-defined cobalt complexes to catalyze the polymerization of conjugated dienes were made. Many kinds of monodentate, bidentate, tridentate and tetradentate ligands with phosphine, nitrogen, oxygen, and sulfur atoms were applied for the synthesis of cobalt catalysts [8,22]. For example, researchers started using phosphorus ligands to synthesize cobalt complexes, which demonstrated the desired management of microstructure and molecular weight in the polymerization of butadiene, isoprene and 1,3-hexadienes in combination with an Al cocatalyst [3,23,24,25,26,27,28]. By skillfully altering the ligand structure and ligating atom, the cobalt catalysts were able to exhibit high activity, and a wide variety of polymers with different properties and structures were available. Further work on polymerization of diene catalyzed by cobalt complexes with tridentate ligands or tetradentate ligand containing a N atom and other O, P or S heteroatoms (N-N-N [29,30,31,32,33,34,35,36,37,38,39], N-N-O [40,41,42,43], N-N-S [44], N-N-P [45,46,47] and N-N-O-O [48], etc.) have been reported. For instance, Gong and his coworkers synthesized a novel type of hemilabile donor ligated cobalt dichloride complexes (N,N,P/O/S), and the nature of the hemilabile donor (P, O and S) determines the coordination chemistry of the complexes, which demonstrated the high activity and enriched cis-1,4- selectivity of polyisoprene. The kinetic studies reveal that the catalyst systems follow a typical living mechanism [46]. Meanwhile, N,N-containing bidentate ligands, such as iminopyridine and α-diimine, have also been tested for cobalt-catalyzed polymerization, and attracted extensive investigations in the past few years [49,50,51,52,53,54,55,56,57,58]. For example, Sun and coworkers reported a series of (8-(arylimino)-5,6,7-trihydroquinolin-2-yl)methyl acetate (N,N) ligated cobalt complexes and applied them in isoprene polymerization. This catalyst system demonstrated a high activity of 1.37 × 105 (mol of Co)−1(h)−1 to give polyisoprene with high molecular weight and the cis-1,4 configuration nearly to 70%.
Due to our ongoing burning desire to develop polymerization catalysts with high performance, we are interested in transition metal catalysts coordinating with N,N-bidentate ligands for diene polymerization [58,59,60,61,62,63,64]. In a recent piece of work, we reported that the pyridine–oxime ligand exhibited high performance in iron catalyzed isoprene polymerization with very low amount of methylaluminoxane [65]. Confirming the fundamental role played by the nature of the pyridine–oxime ligand and the type of cobalt catalyst in determining the catalytic activity and selectivity [66,67], herein we describe a series of pyridine–oxime ligated cobalt catalysts and study their catalytic properties for isoprene polymerization.

2. Materials and Methods

2.1. General Considerations

All experiments were carried out under argon atmosphere by using standard Schlenk techniques or in glovebox. CoCl2 (Macklin Biochemical Co., Ltd., Shanghai, China) and diethylaluminum chloride (AlEt2Cl, 2.0 M solution in hexane, Macklin Biochemical Co., Ltd., Shanghai, China) were purchased and used without further purification. Toluene (Sinopharm Chemical Reagent, Shanghai, China) was refluxed over sodium and distilled and stored over molecular sieves under nitrogen. Hexane and dichloromethane (Sinopharm Chemical Reagent, Shanghai, China) were refluxed over calcium hydride and distilled and stored over molecular sieves in glove box. Isoprene, picolinaldehyde, MeOH, Na2CO3 and Na2SO4 (Sinopharm Chemical Reagent, Shanghai, China) was purchased and freshly distilled over calcium hydride under nitrogen atmosphere.

2.2. Co(II) Complexes Preparation

Ligands L1 and L3 were purchased and used directly without treatment. The ligands L2 and L4 were synthesized and characterized according to the previous report [65]. A typical procedure for the synthesis of ligand was carried out as follows. To a suspension of picolinaldehyde (1.0 equiv.) in MeOH, the corresponding amine (1.0 equiv.) was added, and followed by Na2CO3 (0.5 equiv.). The mixture was stirred at room temperature for 4 h. Then, the suspension was filtered and the solution was transferred to a separatory funnel, washed with water (3 × 20 mL) and dried over anhydrous Na2SO4. The solvent was removed in the rotary evaporator and the product was dried under vacuum to give ligand. All the 1H NMR and 13C NMR of ligands were collected on a Bruker Avance III 400 MHz instrument (Bruker, Karlsruhe, Germany) at 298K, using tetramethyl silane (TMS; CIL, Andover, MA, USA) as internal standard.
L2: yellow solid, 2.1 g, 90% yield; 1H NMR (400 MHz, CDCl3, 298 K) δ 8.27–8.24 (m, 2H), 7.63–7.54 (m, 2H), 7.14 (m, 1H), 2.58 (s, 3H); 13C NMR (100 MHz, CDCl3, 298 K) δ 158.5, 151.0, 150.9, 136.7, 123.8, 118.2, 24.3.
L4: light yellow oil, 3.1 g, 89% yield; 1H NMR (400 MHz, CDCl3, 298 K) δ 8.61–8.59 (m, 1H), 8.15 (s, 1H), 7.77 (dt, J = 8.0, 1.2 Hz, 1H), 7.68 (dt, J = 7.7, 1.8 Hz, 1H), 7.24 (m, 1H), 4.02 (s, 3H). 13C NMR (100 MHz, CDCl3, 298 K) δ 151.5, 149.5, 149.0, 136.2, 123.9, 121.0, 62.4.
All the cobalt complexes Co1Co4 were synthesized using the following methods. Into a 25 mL flask, CoCl2 (1 equiv.) and 8 mL CH2Cl2 in glove box were added, and the ligand in 2 mL CH2Cl2 was added dropwise. The resultant mixture was stirred at room temperature overnight. The solvent was removed under vacuum to 5 mL and the complex was precipitated by addition of hexane (3 mL). The precipitation was collected by filtration and washed three times with hexane (3 × 5 mL) and dried under vacuum to produce Co(II) complex. Mass spectra for cobalt complexes were detected using ACQUITYTM UPLC & Q-TOF MS Premier (Waters, Milford, MA, USA) at Shanghai Jiao Tong University (Shanghai, China). Elemental analysis was recorded using Vario EL III elemental analyzer (Elementar Corporation, Hanau, Germany) at Shanghai Institute of Organic Chemistry (Shanghai, China). X-ray diffraction data was obtained on Smart 1000 diffractometer with Mo K-alpha X-ray source (λ = 0.71073 Å) at 298 K (Bruker, Karlsruhe, Germany) at Liaocheng university (Shandong, China). Attenuated total reflection–infrared (ATR-IR) spectroscopy was conducted using Thermo Scientific Nicolet iN10 (Thermo Fisher Scientific, Waltham Mass, America) at Shanghai Jiao Tong University (Shanghai, China).
Co1: CoCl2 (1 equiv.) and ligand L1 (2 equiv.) were conducted to give jade-green solid. A total of 359 mg, 83% yield. ATR-IR (cm−1): 3064, 1640, 1602, 1478, 1450, 1303, 1254, 1037, 1014, 949, 886, 774. TOF-MS-ES+ (m/z): calcd. For [C12H12ClCoN4O2·H2O·3CH3CN]+: 481.0853, found: 480.9859. Anal.: calcd. For C12H12Cl2CoN4O2: C, 38.53; H, 3.23; N, 14.98; found: C, 38.34; H, 3.23; N, 14.81.
Co2: CoCl2 (1 equiv.) and ligand L2 (2 equiv.) were conducted to give orange solid. A total of 421 mg, 87% yield. ATR-IR (cm−1): 3055, 1649, 1601, 1493, 1458, 1317, 1255, 1046, 1005, 953, 802, 695. TOF-MS-ES+ (m/z): calcd. For [C14H16ClCoN4O2·3MeOH·3CH3CN]+: 555.1585, found: 555.0580. Anal.: calcd. For C14H16Cl2CoN4O2: C, 41.81; H, 4.01; N, 13.93; found: C, 41.96; H, 4.02; N, 13.92.
Co3: CoCl2 (1 equiv.) and ligand L3 (2 equiv.) were conducted to give green solid. A total of 272 mg, 86% yield. ATR-IR (cm−1): 3064, 1640, 1602, 1478, 1450, 1303, 1254, 1037, 1014, 949, 886, 774, 717. TOF-MS-ES+ (m/z): calcd. for [C12H10ClCoN2O·2MeOH·2CH3CN]+: 438.0863, found: 438.0762. Anal.: calcd. For C24H20Cl2CoN4O2: C, 54.77; H, 3.83; N, 10.65; found: C, 54.21; H, 4.02; N, 10.39.
Co4: CoCl2 (1 equiv.) and ligand L4 (2 equiv.) were conducted to give purple solid. A total of 257 mg, 83% yield. ATR-IR (cm−1): 1655, 1569, 1522, 1356, 1274, 1023, 930, 771. TOF-MS-ES+ (m/z): calcd. for [C7H8ClCoN2O·2MeOH·CH3CN]+: 335.0441, found: 335.0557. Anal.: calcd. For C14H16Cl2CoN4O2: C, 41.81; H, 4.01; N, 13.93; found: C, 41.75; H, 4.00; N, 13.79.

2.3. Procedure for Isoprene Polymerization

The isoprene polymerization was carried out in a 25 mL Schlenk reactor by using toluene as solvent. In a universal method, the reactor was pumped and inflated three times. The cobalt complex was weighed in glove box and then added into a Schlenk reactor. The required amount of solvent, isoprene and cocatalyst was sequentially added into the Schlenk reactor under argon atmosphere outside of the glove box. At the specified time, the polymerization was quenched with HCl solution in methanol (methanol/HCl = 50/1). Polymer was recovered in methanol with an antioxidant 2,6-di–tert–butyl–4-methylphenol (BHT, 0.5 wt.%). The polymer was collected by filtration and washed with ethanol for several times and dried under vacuum at 50 °C until no change was observed in weight. The polymer yields were determined by gravimetry.

2.4. Polymer Characterizations

The polymer structure was analyzed by NMR spectra conducting on a Bruker Advance 400 spectrometer at 298 K. 1H NMR (400 Hz, CDCl3) and 13C NMR spectra (100 Hz, CDCl3) of polyisoprene were recorded by using trimethylsilane as internal reference. The polyisoprene microstructure of the 1,4 and 3,4 ratio was determined from 1H NMR of the 1,4 =CH signals at 5.15 ppm and the 3,4 =CH2 signal at 4.7 ppm. The trans/cis-1,4 stereoisomer ratio was distinguished from 13C NMR of the –CH3 signals at 23.8 ppm (for cis-1,4) and at 16.3 ppm (for trans-1,4). The molecular weights (Mn) and molecular weight distributions (Mw/Mn) of polymers were measured using gel permeation chromatography (GPC) using a PL-GPC 50 chromatography and maintained at 25 °C by using THF as eluent and polystyrene as standard.

3. Results and Discussions

3.1. Synthesis and Characterization of Co(II) Complexes

The Co(II) complexes were synthesized by the prepared ligands coordinating with cobalt chloride in anhydrous CH2Cl2 under argon atmosphere to provide the targeted complexes Co1Co4 with high yields. All the cobalt complexes performed well in characterization analysis, including ATR-IR, high resolution mass spectroscopies and elemental analysis (Scheme 1). Single crystals of complex Co2 were obtained by diffusing hexane into its saturated dichloromethane solution at −30 °C in glovebox, and have been deposited in CCDC with codes number of 2160527. As seen in the structure shown in Figure 1 and Tables S1–S3, the Co2 complex adopted a distorted octahedron coordination geometry surrounded by two L2 ligands and one CoCl2 molecule, wherein two Npyridine atoms (N1 and N3) and two chloride atoms formed the basal equatorial plane and the angle of the axis N2-Co-N4 is 172.1°. Moreover, the Noxime atoms of each ligand have a stronger coordination than Npyridine at the cobalt center according to the bond lengths [Co–N1 = 2.188(17), Co–N2 = 2.098(16), Co–N3 = 2.219(16), Co–N4 = 2.132(16)]. Consistent with the previous work of pyridine–oxime iron complex [65], the oxime proton is unionized.

3.2. Polymerization of Isoprene Catalyzed by Co(II) Complexes

Initial screening for isoprene polymerization was concentrated on the type of catalyst to identify whether there were some differences between N-OH group and N-OMe group. Complexes Co1 and Co4 were used as precatalysts activated by AlEt2Cl in toluene solution (5 mL) under various temperatures, as summarized in Table 1. The results revealed that the complex Co1 converted the monomer to polyisoprene with high activity under the condition of [Co]/[isoprene]/[AlEt2Cl] = 1/2000/500, giving a polymerization yield of 95% at 25 °C within 10 min. The produced polyisoprene has a structure with 69% of cis-1,4 and 31% of 3,4 units and a molecular weight of 1.2 × 105 g/mol with low molecular weight distribution of 1.6 (entry 1, Table 1). In addition, complex Co4 showed a similar manner of polymerization with a lower activity, with 6.8 × 105 g (mol of Co)−1(h)−1, than Co1, with 7.7 × 105 g (mol of Co)−1(h)−1 (entry 2, Table 1). Changing the reaction temperature did have effects on reactivity and selectivity. At 70 °C, the complex Co1 could lead to a polyisoprene yield higher than 83%, with high thermostability when the reaction was stopped at 10 min (entry 3, Table 1). But for complex Co4, the reaction activity decreased from 6.8 × 105 g (mol of Co)−1(h)−1 at 25 °C to 4.6 × 105 g (mol of Co)−1(h)−1 at 70 °C (entry 4, Table 1). And, the above results may be connected to the fact that the N-OH moiety displayed a higher catalytic activity and higher heat stability than the N-OMe group under the same current conditions. However, when the polymerization reactions were conducted at −30 °C, no polyisoprene could be produced for both complexes Co1 and Co4 (entries 5 and 6, Table 1). In addition, significant composition drift is not observed for the resultant isoprene for complexes Co1 and Co4 with enriched cis-1,4 motif. And, the cis-1,4 units content slightly decrease at high temperature.
Subsequently, complexes Co1 catalyzed isoprene polymerization with various cocatalyst feeding amounts and reaction time were summarized, as seen in Table 2. For complex Co1, changing the amount of [Co]/[AlEt2Cl] from 1/500 to 1/50 (entry 1, Table 2) did not vary reactivity significantly when the reaction went on for 10 min, and also gave >99% yield of polyisoprene. And, complete conversion of isoprene could be achieved even though the [Co]/[AlEt2Cl] is changed to 1/10 (entry 2, Table 2), which is virtually irrelevant to cocatalyst feeding, suggestive of the conformation stability of active species. It should be pointed out that this catalytic system showed extremely high activity (8.16 × 105 g (mol of Co)−1(h)−1), which is comparable with the recent work on rigid Co(II) chloride catalysts (1.37 × 105 g (mol of Co)−1(h)−1) [68] and α-diimine cobalt catalysts (0.54 × 105 g (mol of Co)−1(h)−1) [58]. All resultant polyisoprenes are predominantly cis-1,4 enchained with about 30% of 3,4 units. Interestingly, the content of the 3,4-unit polymer slightly increased from 31% to 36% with the change in [Co]/[AlEt2Cl] ratio from 1/500 to 1/10 (Figure 2 and Figures S11 and S12). The molecular weight of polyisoprene was found to be dependent on the amount of AlEt2Cl. The molecular weight of polyisoprene increased gradually with widened molecular weight distribution as the cocatalyst feeding reduced, which could be explained by the occurrence of transfer reaction. In order to give the best cocatalyst feeding, the polymerization under the [Co]/[AlEt2Cl] = 1/50 and [Co]/[AlEt2Cl] = 1/10 were stopped within 5 min. This proved that a low amount of AlEt2Cl has higher activity (entry 4, Table 2). These may be related to the dialkylated species with a significant amount of AlEt2Cl, which makes the Co center difficult to cationize to coordinate isoprene. Excitingly, even if the [Co]/[AlEt2Cl] ratio was altered to 1/5, the polymerization could still be carried out, and produced a 38% yield of polymer after 1 h, suggestive of the high catalytic ability of the pyridine–oxime ligated Co catalyst. It is worth mentioning that AlEt2Cl dosage is much less than in the previous catalyst systems.
The titled complexes Co2 and Co3 with methyl and phenyl group were individually tested under the same polymerization conditions for comparison. The isoprene polymerization catalyzed by Co2, which had a methyl group on the 6-position of pyridine ring, also showed full conversion at the [Co]/[AlEt2Cl] = 1/50 within 10 min (entry 6, Table 2). However, it needs to be emphasized that the pyridine–oxime ligated iron complex bearing one 6-methyl substituent at pyridine ring showed lower activity because of steric hindrance in the previous work [65]. Zhang’s group also reported that the substituent at the 6-position of the pyridine ring significantly influenced the catalytic performances of the iminopyridine ligated Co complexes. For the aldimine- and ketimine-based cobalt complexes, the introduction of a halogen atom (Br) at 6-position of the pyridine ring afforded polymers in higher yields but with a relatively lower molecular weight than the complexes without substituent. The complexes with CH3 group produced polymers in lower yields but with higher molecular weights and narrower molecular weight distributions [54]. This may be because that the N-OH group in this work is small enough to tolerates steric hindrance at 6-position of the pyridine ring. When the [Co]/[AlEt2Cl] ratio was changed to 1/10, the polymer yield and cis-1,4 motif of polyisoprene decreased slightly (entry 7, Table 2). The Ph-substituted complex Co3 produced polyisoprene with a lower yield at the [Co]/[AlEt2Cl] = 1/50. But, similar to complex Co1, it exhibited higher reaction activity following the reduction in the amount of AlEt2Cl, with [Co]/[AlEt2Cl] = 1/10. Though with high hindrance, complexes Co2 and Co3 provided polyisoprene paralleling that of Co1 in terms of microstructure, molecular weight (1.7–2.2 × 105 g/mol) and molecular weight distribution (1.7–2.0). These results indicated that these cobalt catalysts all possess high activity without the effect of steric hindrance.
Furthermore, the catalytic ability of catalyst Co4 and AlEt2Cl is shown in Table 3. Less cocatalyst feeding leads to increased polymer yield, and it displayed the highest activity of 16.3 × 105 g (mol of Co)−1(h)−1 with [Co]/[AlEt2Cl] = 1/50 (entry 2, Table 3). It was interesting that the complex Co4 exhibited higher activity than Co1 in terms of the reduction of cocatalyst feeding. However, the polymerization did not conduct at [Co]/[AlEt2Cl] = 1/5, even though the reaction time was prolonged to 120 min, which was inferior compared to catalyst Co1. The produced polyisoprene possessed a rather balanced molecular weight of 10.2–14.5 × 104 g/mol with unimodal molecular weight distributions. In addition, changing the cocatalyst feeding does not vary the polyisoprene structures with 65% of cis-1,4- unit and 35% of 3,4- unit.

4. Conclusions

In conclusion, cobalt catalysts bearing pyridine–oxime and picolinaldehyde O-methyl oxime ligands were synthesized and characterized. In combination with AlEt2Cl, the Co complexes exhibited a remarkable activity of 1.6 × 106 g (mol of Co)−1(h)−1 and required only a small amount of cocatalyst feeding ([Co]/[AlEt2Cl] = 1/10), which is very promising for the synthetic rubber industry. The complex Co1Co3 exhibited comparative activity and selectivity to the catalyst Co4, and the reaction catalyzed by Co1 could conduct even the Al feeding as low as to 5. Conversely, the polymerization displayed low conversion at the elevating and reducing reaction temperature, indicating the thermal instability of active species. And, all the catalyst systems provided polyisoprene with parallel molecular weights, molecular weight distribution and cis-14- enriched structure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/2073-4360/15/24/4660/s1, Figures S1–S32: NMR data for polyisoprene. Figures S33–S48: GPC data for polyisoprene. Table S1: Crystal data and structure refinement for complex Co2. Table S2: Bond lengths for complex Co2. Table S3: Selected bond angles for complex Co2.

Author Contributions

Y.D.: Methodology, Formal Analysis, Investigation, Writing—Original Draft Preparation; S.G.: Formal Analysis, Investigation, Writing—Review and Editing; H.M.: Investigation, Visualization; S.L.: Investigation, Visualization; Z.Z.: Supervision, Project Administration; M.Z.: Supervision, Project Administration, Funding Acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Natural Science Foundation of Shandong Province (No. ZR2022QB038 of M.Z.), Linyi University Doctoral Research Foundation grant funded by Linyi University (No. LYDX2020BS012 of M.Z.) and the College Students’ Innovation and Entrepreneurship Training Program by Linyi University (No. X202310452208 of M.Z.).

Data Availability Statement

Data are contained within the article and supplementary materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Visseaux, M. Catalysts for the controlled polymerization of conjugated dienes. Catalysts 2018, 8, 442–445. [Google Scholar] [CrossRef]
  2. Thiele, S.K.H.; Wilson, D.R. Alternate transition metal complex based diene polymerization. J. Macromol. Sci. C Polym. Rev. 2003, 43, 581–628. [Google Scholar] [CrossRef]
  3. Ricci, G.; Leone, G.; Boglia, A.; Boccia, A.C.; Zetta, L. Cis-1,4-alt-3,4 polyisoprene: Synthesis and characterization. Macromolecules 2009, 42, 9263–9267. [Google Scholar] [CrossRef]
  4. Zhao, J.; Chen, H.; Li, W.; Jia, X.; Zhang, X.; Gong, D.R. Polymerization of isoprene promoted by aminophosphine(ory)fused bipyridine cobalt complexes: Precise control of molecular weight and cis-1,4-alt-3,4 sequence. Inorg. Chem. 2018, 57, 4088–4097. [Google Scholar] [CrossRef] [PubMed]
  5. Nekhaeva, L.A.; Bondarenko, G.N.; Frolov, V.M. Transition metal compounds in combination with alkylaluminoxanes as catalysts for olefin and diene polymerization. Kinet. Catal. 2003, 44, 631–638. [Google Scholar] [CrossRef]
  6. Ricci, G.; Sommazzi, A.; Masi, F.; Ricci, M.; Boglia, A.; Leone, G. Well-defined transition metal complexes with phosphorus and nitrogen ligands for 1,3-dienes polymerization. Coordin. Chem. Rev. 2010, 254, 661–676. [Google Scholar] [CrossRef]
  7. Huang, J.M.; Liu, Z.H.; Cui, D.M.; Liu, X.L. Precisely controlled polymerization of styrene and conjugated dienes by group 3 single-site catalysts. Chemcatchem 2018, 10, 42–61. [Google Scholar] [CrossRef]
  8. Ricci, G.; Pampaloni, G.; Sommazzi, A.; Masi, F. Dienes polymerization: Where we are and what lies ahead. Macromolecules 2021, 54, 5879–5914. [Google Scholar] [CrossRef]
  9. Cuomo, C.; Serra, M.C.; Maupoey, M.G.; Grassi, A. Copolymerization of styrene with butadiene and isoprene catalyzed by the monocyclopentadienyl titanium complex Ti(η5-C5H5)(η2-MBMP)Cl. Macromolecules 2007, 40, 7089–7097. [Google Scholar] [CrossRef]
  10. Porri, L.; Ricci, G.; Giarrusso, A.; Shubin, N.; Lu, Z. Recent developments in lanthanide catalysts for 1,3-diene polymerization. Am. Chem. Soc. 1999, 749, 15–30. [Google Scholar]
  11. Racanelli, P.; Porri, L. Cis-1,4-polybutadiene by cobalt catalysts. Some features of the catalysts prepared from alkyl aluminium compounds containing Al-O-Al bonds. Eur. Polym. J. 1970, 6, 751–761. [Google Scholar] [CrossRef]
  12. Liu, B.; Li, L.; Sun, G.; Liu, J.; Wang, M.; Li, S.; Cui, D. 3,4-Polymerization of isoprene by using NSN- and NPN-ligated rare earth metal precursors: Switching of stereo selectivity and mechanism. Macromolecules 2014, 47, 4971–4978. [Google Scholar] [CrossRef]
  13. Diether, D.; Meermann-Zimmermann, M.; Törnroos, K.W.; Maichle-Mössmer, C.; Anwander, R. Rare-earth metal-promoted (double) C–H bond activation of a lutidinyl-functionalized alkoxy ligand: Formation of [ONC] pincer-type ligands and implications for isoprene polymerization. Dalton Trans. 2020, 49, 2004–2013. [Google Scholar] [CrossRef] [PubMed]
  14. Phuphuak, Y.; Bonnet, F.; Vendier, L.; Lorber, C.; Zinck, P. Isoprene polymerization mediated by vanadium-[ONNO] complexes. Dalton Trans. 2016, 45, 12069–12077. [Google Scholar] [CrossRef] [PubMed]
  15. Jing, C.Y.; Wang, L.; Mahmood, Q.; Zhao, M.M.; Zhu, G.Q.; Zhang, X.H.; Wang, X.W.; Wang, Q.G. Synthesis and characterization of aminopyridine iron(ii) chloride catalysts for isoprene polymerization: Sterically controlled monomer enchainment. Dalton Trans. 2019, 48, 7862–7874. [Google Scholar] [CrossRef] [PubMed]
  16. O’Connor, A.R.; Brookhart, M. Polymerization of 1,3-dienes and styrene catalyzed by cationic allyl Ni(II) complexes. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 1901–1912. [Google Scholar] [CrossRef]
  17. Gao, B.; Luo, X.Y.; Gao, W.; Huang, L.; Gao, S.-m.; Liu, X.M.; Wu, Q.L.; Mu, Y. Chromium complexes supported by phenanthrene-imine derivative ligands: Synthesis, characterization and catalysis on isoprene cis-1,4 polymerization. Dalton Trans. 2012, 41, 2755–2763. [Google Scholar] [CrossRef]
  18. Capacchione, C.; Saviello, D.; Avagliano, A.; Proto, A. Copolymerization of ethylene with isoprene promoted by titanium complexes containing a tetradentate [OSSO]-type bis(phenolato) ligand. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 4200–4206. [Google Scholar] [CrossRef]
  19. Cai, Z.; Shinzawa, M.; Nakayama, Y.; Shiono, T. Synthesis of regioblock polybutadiene with CoCl2-based catalyst via reversible coordination of lewis base. Macromolecules 2009, 42, 7642–7643. [Google Scholar] [CrossRef]
  20. Nath, D.C.D.; Shiono, T.; Ikeda, T. Copolymerization of 1,3-butadiene and isoprene with cobalt dichloride/methylaluminoxane in the presence of triphenylphosphine. J. Polym. Sci. Part A Polym. Chem. 2002, 40, 3086–3092. [Google Scholar] [CrossRef]
  21. Endo, K.; Hatakeyama, N. Stereospecific and molecular weight-controlled polymerization of 1,3-butadiene with Co(acac)3-MAO catalyst. J. Polym. Sci. Part A Polym. Chem. 2001, 39, 2793–2798. [Google Scholar] [CrossRef]
  22. Gong, D.; Tang, F.; Xu, Y.; Hu, Z.; Luo, W. Cobalt catalysed controlled copolymerization: An efficient approach to bifunctional polyisoprene with enhanced properties. Poly. Chem. 2021, 12, 1653–1660. [Google Scholar] [CrossRef]
  23. Ricci, G.; Forni, A.; Boglia, A.; Motta, T. Synthesis, structure, and butadiene polymerization behavior of alkylphosphine cobalt(II) complexes. J. Mol. Catal. A Chem. 2005, 226, 235–241. [Google Scholar] [CrossRef]
  24. Ricci, G.; Forni, A.; Boglia, A.; Motta, T.; Zannoni, G.; Canetti, M.; Bertini, F. Synthesis and X-ray structure of CoCl2(PiPrPh2)2. A new highly active and stereospecific catalyst for 1,2 polymerization of conjugated dienes when used in association with MAO. Macromolecules 2005, 38, 1064–1070. [Google Scholar] [CrossRef]
  25. Ricci, G.; Forni, A.; Boglia, A.; Sommazzi, A.; Masi, F. Synthesis, structure and butadiene polymerization behavior of CoCl2(PRxPh3–x)2 (R=methyl, ethyl, propyl, allyl, isopropyl, cyclohexyl; x=1, 2). Influence of the phosphorous ligand on polymerization stereoselectivity. J. Organomet. Chem. 2005, 690, 1845–1854. [Google Scholar] [CrossRef]
  26. Ricci, G.; Motta, T.; Boglia, A.; Alberti, E.; Zetta, L.; Bertini, F.; Arosio, P.; Famulari, A.; Meille, S.V. Synthesis, characterization, and crystalline structure of syndiotactic 1,2-polypentadiene:  The trans polymer. Macromolecules 2005, 38, 8345–8352. [Google Scholar] [CrossRef]
  27. Ricci, G.; Boglia, A.; Motta, T.; Bertini, F.; Boccia, A.C.; Zetta, L.; Alberti, E.; Famulari, A.; Arosio, P.; Meille, S.V. Synthesis and structural characterization of syndiotactic trans-1,2 and cis-1,2 polyhexadienes. J. Polym. Sci. Part A Polym. Chem. 2007, 45, 5339–5353. [Google Scholar] [CrossRef]
  28. Ricci, G.; Leone, G.; Boglia, A.; Bertini, F.; Boccia, A.C.; Zetta, L. Synthesis and characterization of isotactic 1,2-poly(E-3-methyl-1,3-pentadiene). Some remarks about the influence of monomer structure on polymerization stereoselectivity. Macromolecules 2009, 42, 3048–3056. [Google Scholar] [CrossRef]
  29. Ricci, G.; Leone, G.; Pierro, I.; Zanchin, G.; Forni, A. Novel cobalt dichloride complexes with hindered diphenylphosphine ligands: Synthesis, characterization, and behavior in the polymerization of butadiene. Molecules 2019, 24, 2308–2315. [Google Scholar] [CrossRef] [PubMed]
  30. Kim, J.S.; Ha, C.-S.; Kim, I. Highly stereospecific polymerizations of 1,3-butadiene with cobalt(II) pyridyl bis(imine) complexes. e-Polymers 2006, 6, 1–9. [Google Scholar] [CrossRef]
  31. Appukuttan, V.; Zhang, L.; Ha, C.-S.; Kim, I. Highly active and stereospecific polymerizations of 1,3-butadiene by using bis(benzimidazolyl)amine ligands derived Co(II) complexes in combination with ethylaluminum sesquichloride. Polymer 2009, 50, 1150–1158. [Google Scholar] [CrossRef]
  32. Cariou, R.; Chirinos, J.J.; Gibson, V.C.; Jacobsen, G.; Tomov, A.K.; Britovsek, G.J.P.; White, A.J.P. The effect of the central donor in bis(benzimidazole)-based cobalt catalysts for the selective cis-1,4-polymerisation of butadiene. Dalton Trans. 2010, 39, 9039–9045. [Google Scholar] [CrossRef] [PubMed]
  33. Gong, D.; Jia, X.; Wang, B.; Zhang, X.; Jiang, L. Synthesis, characterization, and butadiene polymerization of iron(III), iron(II) and cobalt(II) chlorides bearing 2,6-bis(2-benzimidazolyl)pyridyl or 2,6-bis(pyrazol)pyridine ligand. J. Organomet. Chem. 2012, 702, 10–18. [Google Scholar] [CrossRef]
  34. Gong, D.; Jia, W.; Chen, T.; Huang, K.-W. Polymerization of 1,3-butadiene catalyzed by pincer cobalt(II) complexes derived from 2-(1-arylimino)-6-(pyrazol-1-yl)pyridine ligands. Appl. Catal. A-Gen. 2013, 464–465, 35–42. [Google Scholar] [CrossRef]
  35. He, A.; Wang, G.; Zhao, W.; Jiang, X.; Yao, W.; Sun, W.-H. High cis-1,4 polyisoprene or cis-1,4/3,4 binary polyisoprene synthesized using 2-(benzimidazolyl)-6-(1-(arylimino)ethyl)pyridine cobalt(II) dichlorides. Polym. Int. 2013, 62, 1758–1766. [Google Scholar] [CrossRef]
  36. Gong, D.; Wang, B.; Jia, X.; Zhang, X. The enhanced catalytic performance of cobalt catalysts towards butadiene polymerization by introducing a labile donor in a salen ligand. Dalton Trans. 2014, 43, 4169–4178. [Google Scholar] [CrossRef]
  37. Wang, G.; Jiang, X.; Zhao, W.; Sun, W.-H.; Yao, W.; He, A. Catalytic behavior of Co(II) complexes with 2-(benzimidazolyl)-6-(1-(arylimino)ethyl)pyridine ligands on isoprene stereospecific polymerization. J. Appl. Polym. Sci. 2014, 131, 1–6. [Google Scholar] [CrossRef]
  38. Alnajrani, M.N.; Mair, F.S. The behaviour of β-triketimine cobalt complexes in the polymerization of isoprene. RSC Adv. 2015, 5, 46372–46385. [Google Scholar] [CrossRef]
  39. Guo, J.; Wang, B.; Bi, J.; Zhang, C.; Zhang, H.; Bai, C.; Hu, Y.; Zhang, X. Synthesis, characterization and 1,3-butadiene polymerization studies of cobalt dichloride complexes bearing pyridine bisoxazoline ligands. Polymer 2015, 59, 124–132. [Google Scholar] [CrossRef]
  40. Jia, X.; Li, W.; Zhao, J.; Yi, F.; Luo, Y.; Gong, D. Dual catalysis of the selective polymerization of biosourced myrcene and methyl methacrylate promoted by salicylaldiminato cobalt(II) complexes with a pendant donor. Organometallics 2019, 38, 278–288. [Google Scholar] [CrossRef]
  41. Yang, D.; Gan, Q.; Chen, H.; Ying, W.; Zhao, J.; Jia, X.; Gong, D. Polymerization of conjugated dienes and olefins promoted by cobalt complexes supported by phosphine oxide ligands. Inorg. Chim. Acta 2019, 496, 119046. [Google Scholar] [CrossRef]
  42. Xu, Y.; Zhao, J.; Gan, Q.; Ying, W.; Hu, Z.; Tang, F.; Luo, W.; Luo, Y.; Jian, Z.; Gong, D. Synthesis and properties investigation of hydroxyl functionalized polyisoprene prepared by cobalt catalyzed co-polymerization of isoprene and hydroxylmyrcene. Polym. Chem. 2020, 11, 2034–2043. [Google Scholar] [CrossRef]
  43. Chen, X.; Huang, L.; Gao, W. One-pot synthesis of cobalt complexes with 2,6-bis(arylimino)phenoxyl/phenthioxyl ligands and catalysis on isoprene polymerization. Dalton Trans. 2021, 50, 5218–5225. [Google Scholar] [CrossRef] [PubMed]
  44. Gong, D.; Ying, W.; Zhao, J.; Li, W.; Xu, Y.; Luo, Y.; Zhang, X.; Capacchione, C.; Grassi, A. Controlling external diphenylcyclohexylphosphine feeding to achieve cis-1,4-syn-1,2 sequence controlled polybutadienes via cobalt catalyzed 1,3-butadiene polymerization. J. Catal. 2019, 377, 367–377. [Google Scholar] [CrossRef]
  45. Chen, L.; Ai, P.; Gu, J.; Jie, S.; Li, B.-G. Stereospecific polymerization of 1,3-butadiene catalyzed by cobalt complexes bearing N-containing diphosphine PNP ligands. J. Organom. Chem. 2012, 716, 55–61. [Google Scholar] [CrossRef]
  46. Chen, H.; Pan, W.; Huang, K.-W.; Zhang, X.; Gong, D. Controlled polymerization of isoprene promoted by a type of hemilabile X PN3 (X = O, S) ligand supported cobalt(ii) complexes: The role of a hemilabile donor on the level of control. Polym. Chem. 2017, 8, 1805–1814. [Google Scholar] [CrossRef]
  47. Li, W.; Zhao, J.; Zhang, X.; Gong, D. Capability of PN3-type cobalt complexes toward selective (Co-)polymerization of myrcene, butadiene, and isoprene: Access to biosourced polymers. Ind. Eng. Chem. Res. 2019, 58, 2792–2800. [Google Scholar] [CrossRef]
  48. Endo, K.; Kitagawa, T.; Nakatani, K. Effect of an alkyl substituted in salen ligands on 1,4-cis selectivity and molecular weight control in the polymerization of 1,3-butadiene with (salen)Co(II) complexes in combination with methylaluminoxane. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 4088–4094. [Google Scholar] [CrossRef]
  49. Jia, X.; Liu, H.; Hu, Y.; Dai, Q.; Bi, J.; Bai, C.; Zhang, X. Highly active and cis-1,4 selective polymerization of 1,3-butadiene catalyzed by cobalt(II) complexes bearing α-diimine ligands. Chin. J. Catal. 2013, 34, 1560–1569. [Google Scholar] [CrossRef]
  50. Jie, G.; Feng, B.; Ying, C.; Tao, C.; Rui, M.; Pan, L.; Xiaobing, H.; Zhenzhen, L.; Jianlan, M.; Chunjie, Y. Cobalt complex bearing β-ketoamine ligand: Syntheses, structure, catalytic behavior for styrene and methyl methacrylate polymerization. Appl. Organomet. Chem. 2014, 28, 584–588. [Google Scholar] [CrossRef]
  51. Alnajrani, M.N.; Alshmimri, S.A.; Alsager, O.A. Alpha and beta diimine cobalt complexes in isoprene polymerization: A comparative study. RSC Adv. 2016, 6, 113803–113814. [Google Scholar] [CrossRef]
  52. Alnajrani, M.N.; Mair, F.S. Bidentate forms of β-triketimines: Syntheses, characterization and outstanding performance of enamine–diimine cobalt complexes in isoprene polymerization. Dalton Trans. 2016, 45, 10435–10446. [Google Scholar] [CrossRef] [PubMed]
  53. Dai, Q.; Jia, X.; Yang, F.; Bai, C.; Hu, Y.; Zhang, X. Iminopyridine-based Cobalt(II) and Nickel(II) complexes: Synthesis, characterization, and their catalytic behaviors for 1,3-butadiene polymerization. Polymers 2016, 8, 12. [Google Scholar] [CrossRef]
  54. Wang, X.; Fan, L.; Huang, C.; Liang, T.; Guo, C.-Y.; Sun, W.-H. Highly cis-1,4 selective polymerization of isoprene promoted by α-diimine cobalt(II) chlorides. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 3609–3615. [Google Scholar] [CrossRef]
  55. Fang, L.; Zhao, W.; Han, C.; Liu, H.; Hu, Y.; Zhang, X. Isoprene polymerization with pyrazolylimine Cobalt(II) complexes: Manipulation of 3,4-selectivities by ligand design and use of triphenylphosphine. Eur. J. Inorg. Chem. 2019, 2019, 609–616. [Google Scholar] [CrossRef]
  56. Zhang, X.; Zhu, G.; Mahmood, Q.; Zhao, M.; Wang, L.; Jing, C.; Wang, X.; Wang, Q. Iminoimidazole-based Co(II) and Fe(II) complexes: Syntheses, characterization, and catalytic behaviors for isoprene polymerization. J. Polym. Sci. Part A Polym. Chem. 2019, 57, 767–775. [Google Scholar] [CrossRef]
  57. Liu, L.; Wang, F.; Zhang, C.; Liu, H.; Wu, G.; Zhang, X. Thermally robust alpha-diimine nickel and cobalt complexes for cis-1,4 selective 1,3-butadiene polymerizations. Mol. Catal. 2022, 517, 112044. [Google Scholar] [CrossRef]
  58. Chen, B.; Gong, D. Polymerization of butadiene and isoprene using α-diimine cobalt catalysts bearing electron donor at the N-aryl of imine. J. Organomet. Chem. 2023, 994, 122727. [Google Scholar] [CrossRef]
  59. Zhao, M.; Mahmood, Q.; Jing, C.; Wang, L.; Zhu, G.; Zhang, X.; Wang, Q. Isoprene polymerization: Catalytic performance of iminopyridine vanadium(III) chloride versus vanadium(III) chloride. Polymers 2019, 11, 1122. [Google Scholar] [CrossRef]
  60. Zhao, M.; Wang, L.; Mahmood, Q.; Jing, C.; Zhu, G.; Zhang, X.; Wang, X.; Wang, Q. Controlled isoprene polymerization mediated by iminopyridine-iron (II) acetylacetonate pre-catalysts. Appl. Organometal. Chem. 2019, 33, e4836. [Google Scholar] [CrossRef]
  61. Zhao, M.; Wang, L.; Mahmood, Q.; Jing, C.; Zhu, G.; Zhang, X.; Chen, X.; Wang, Q. Highly active and thermo-stable iminopyridyl vanadium oxychloride catalyzed isoprene polymerization. J. Polym. Sci. 2020, 58, 2708–2717. [Google Scholar] [CrossRef]
  62. Jing, C.; Wang, L.; Zhu, G.; Hou, H.; Zhou, L.; Wang, Q. Enhancing thermal stability in aminopyridine iron(II)-catalyzed polymerization of conjugated dienes. Organometallics 2020, 39, 4019–4026. [Google Scholar] [CrossRef]
  63. Wang, L.; Wang, X.; Hou, H.; Zhu, G.; Han, Z.; Yang, W.; Chen, X.; Wang, Q. An unsymmetrical binuclear iminopyridine-iron complex and its catalytic isoprene polymerization. Chem. Commun. 2020, 56, 8846–8849. [Google Scholar] [CrossRef] [PubMed]
  64. Zhao, M.; Zhang, X.; Wang, L.; Wang, L.; Zhu, G.; Wang, Q. Pyridine-oxazoline ligated iron complexes: Synthesis, characterization, and catalytic activity for isoprene polymerization. Appl. Organomet. Chem. 2022, 36, e6848. [Google Scholar] [CrossRef]
  65. Zhao, M.; Ma, Y.; Zhang, X.; Wang, L.; Zhu, G.; Wang, Q. Synthesis, characterization and catalytic property studies for isoprene polymerization of iron complexes bearing unionized pyridine-oxime ligands. Polymers 2022, 14, 3612. [Google Scholar] [CrossRef] [PubMed]
  66. Ghachtouli, S.E.; Guillot, R.; Brisset, F.; Aukauloo, A. Cobalt-based particles formed upon electrocatalytic hydrogen production by a cobalt pyridine oxime complex. ChemSusChem 2013, 6, 2226–2230. [Google Scholar] [CrossRef] [PubMed]
  67. Du, J.; Meng, X.-G.; Zeng, X.-C. Copper(II) and Cobalt(II) dinuclear complexes with oxime-based ligand as catalysts for the hydrolysis of phosphate diester. Chin. J.Chem. 2008, 26, 421–425. [Google Scholar] [CrossRef]
  68. Yousuf, N.; Ma, Y.P.; Mahmood, Q.; Zhang, W.J.; Liu, M.; Yuan, R.Y.; Sun, W.-H. Structurally rigid (8-(arylimino)-5,6,7-trihydroquinolin-2-yl)-methyl acetate cobalt complex catalysts for isoprene polymerization with high activity and cis-1,4 selectivity. Catalysts 2023, 13, 1120. [Google Scholar] [CrossRef]
Scheme 1. The structures of the precursors employed.
Scheme 1. The structures of the precursors employed.
Polymers 15 04660 sch001
Figure 1. Molecular structure of Co2 (hydrogen atoms were omitted for clarity). Selected bond distances (Å) and angles (deg): Co-N(1) 2.188(17), Co-N(2) 2.098(16), Co-N(3) 2.219(16), Co-N(4) 2.132 (16), Co-Cl(1) 2.470(7), Co-Cl(2) 2.455(6), O(1)-H(1) 0.8200, O(2)-H(2) 0.8200, N(2)-Co-N(4) 172.1(5), N(2)-Co-N(1) 77.2(6), N(4)-Co-N(1) 109.6(6), N(2)-Co-N(3) 109.5(6), N(4)-Co-N(3) 74.7(6), N(1)-Co-N(3) 94.0(4), Cl(2)-Co-Cl(1) 98.68(15). The gray sphere represents the carbon atom, and the hydrogen atom is omitted.
Figure 1. Molecular structure of Co2 (hydrogen atoms were omitted for clarity). Selected bond distances (Å) and angles (deg): Co-N(1) 2.188(17), Co-N(2) 2.098(16), Co-N(3) 2.219(16), Co-N(4) 2.132 (16), Co-Cl(1) 2.470(7), Co-Cl(2) 2.455(6), O(1)-H(1) 0.8200, O(2)-H(2) 0.8200, N(2)-Co-N(4) 172.1(5), N(2)-Co-N(1) 77.2(6), N(4)-Co-N(1) 109.6(6), N(2)-Co-N(3) 109.5(6), N(4)-Co-N(3) 74.7(6), N(1)-Co-N(3) 94.0(4), Cl(2)-Co-Cl(1) 98.68(15). The gray sphere represents the carbon atom, and the hydrogen atom is omitted.
Polymers 15 04660 g001
Figure 2. 1H and 13C NMR spectra of the representative sample of polyisoprene obtained at [Co]/[AlEt2Cl] = 1/10 (Table 2, entry 2). 1H NMR of the 1,4 =CH signals at a and the 3,4 =CH2 signal at b.
Figure 2. 1H and 13C NMR spectra of the representative sample of polyisoprene obtained at [Co]/[AlEt2Cl] = 1/10 (Table 2, entry 2). 1H NMR of the 1,4 =CH signals at a and the 3,4 =CH2 signal at b.
Polymers 15 04660 g002
Table 1. Isoprene polymerization using Co1 and Co4 with different temperature a.
Table 1. Isoprene polymerization using Co1 and Co4 with different temperature a.
EntryCat.Temp.
(°C)
Yield b
(%)
Act. c
(×10−5)
Mnd
(×10−4)
Mw/MndMicrostructure(%) e
trans-1,4cis-1,43,4
1Co125957.712.21.606931
2Co425836.810.21.706832
3Co170836.817.82.006436
4Co470564.64.92.016633
5Co1−30-------
6Co4−30-------
a General conditions: total volume = 8 mL; [isoprene] = 2.5 mol/L; [Co]/[isoprene]/[AlEt2Cl] = 1/2000/500; reaction time = 10 min. b Isolated yield. c gPolymer (mol of Co)−1(h)−1. d Determined by gel permeation chromatography (GPC). e Determined by 1H and 13C NMR.
Table 2. Isoprene polymerization using Co1Co3 with AlEt2Cl as cocatalyst a.
Table 2. Isoprene polymerization using Co1Co3 with AlEt2Cl as cocatalyst a.
EntryCat.[Co]/[AlEt2Cl]Time
(min)
Yield b
(%)
Act. c
(×10−5)
Mn d
(×10−4)
Mw/Mn dMicrostructure(%) e
cis-1,43,4
1Co11/5010>998.215.91.76634
2Co11/1010>998.217.91.96436
3Co11/5057111.612.81.96634
4Co11/1058313.510.12.16436
5Co11/560380.517.51.76337
6Co21/5010>998.216.71.86634
7Co21/1010957.822.01.76436
8Co31/5010897.317.12.06733
9Co31/1010>998.218.42.06436
a General conditions: total volume = 8 mL; [isoprene] = 2.5 mol/L; [Co]/[isoprene] = 1/2000; 25 °C. b Isolated yield. c gPolymer (mol of Co)−1(h)−1. d Determined by gel permeation chromatography (GPC). e Determined by 1H and 13C NMR.
Table 3. Isoprene polymerization using Co4 with AlEt2Cl as cocatalyst a.
Table 3. Isoprene polymerization using Co4 with AlEt2Cl as cocatalyst a.
Entry[Co]/[AlEt2Cl]Time
(min)
Yield b
(%)
Act. c
(×10−5)
Mn d
(×10−4)
Mw/Mn dMicrostructure (%) e
cis-1,43,4
11/10010937.610.21.76535
21/505>9916.314.51.86535
31/1059715.812.32.06337
41/51200-----
a General conditions: total volume = 8 mL; [isoprene] = 2.5 mol/L; [Co]/[isoprene] = 1/2000; 25 °C. b Isolated yield. c gPolymer (mol of Co)−1h−1. d Determined by gel permeation chromatography (GPC). e Determined by 1H and 13C NMR.
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

Du, Y.; Gao, S.; Ma, H.; Lu, S.; Zhang, Z.; Zhao, M. Catalytic Behavior of Cobalt Complexes Bearing Pyridine–Oxime Ligands in Isoprene Polymerization. Polymers 2023, 15, 4660. https://doi.org/10.3390/polym15244660

AMA Style

Du Y, Gao S, Ma H, Lu S, Zhang Z, Zhao M. Catalytic Behavior of Cobalt Complexes Bearing Pyridine–Oxime Ligands in Isoprene Polymerization. Polymers. 2023; 15(24):4660. https://doi.org/10.3390/polym15244660

Chicago/Turabian Style

Du, Yuanxu, Shuo Gao, Hui Ma, Siqi Lu, Zhenhua Zhang, and Mengmeng Zhao. 2023. "Catalytic Behavior of Cobalt Complexes Bearing Pyridine–Oxime Ligands in Isoprene Polymerization" Polymers 15, no. 24: 4660. https://doi.org/10.3390/polym15244660

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop