Next Article in Journal
Development of an Atomic-Oxygen-Erosion-Resistant, Alumina-Fiber-Reinforced, Fluorinated Polybenzoxazine Composite for Low-Earth Orbital Applications
Next Article in Special Issue
Crosslinked Polynorbornene-Based Anion Exchange Membranes with Perfluorinated Branch Chains
Previous Article in Journal
Structural and Physicochemical Characterization of Extracted Proteins Fractions from Chickpea (Cicer arietinum L.) as a Potential Food Ingredient to Replace Ovalbumin in Foams and Emulsions
Previous Article in Special Issue
Direct Synthesis of Chain-End Toluene Functionalized Hyperbranched Ethylene Oligomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Synthesis of Partially Chain-Straightened Propylene Oligomers and P-MA Co-Oligomers Using Axially Flexible Shielded Iminopyridyl Palladium Complexes

1
School of Chemical and Materials Engineering, Huainan Normal University, Huainan 232038, China
2
Institutes of Physical Science and Information Technology, Key Laboratory of Structure and Functional Regulation of Hybrid Materials of Ministry of Education, Anhui University, Hefei 230601, China
3
McKetta Department of Chemical Engineering, University of Texas at Austin, Austin, TX 78712, USA
4
Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology, Changzhou University, Changzhou 213164, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(1), 111; https://doi.org/10.3390/polym15010111
Submission received: 5 December 2022 / Revised: 15 December 2022 / Accepted: 22 December 2022 / Published: 27 December 2022
(This article belongs to the Special Issue Carbon-Based Functional Polymers: Design, Properties and Applications)

Abstract

:
In this study, a series of partially chain-straightened propylene oligomers and functional propylene–methyl acrylate (P-MA) co-oligomers were synthesized with 8-alkyl-iminopyridyl Pd(II) catalysts. The molecular weight and polar monomer incorporation ratio could be tuned by using Pd(II) catalysts with various 8-alkyl-naphthyl substituents (8-alkyl: H, Me, and n-Bu). In propylene oligomerization, all the 8-alkyl-iminopyridyl Pd(II) catalysts convert propylene to partially chain-straightened (119–136/1000 C) oligomers with low molecular weights (0.3–1.5 kg/mol). Among the catalysts, Pd1 with non-substituent (H) on the ligand showed the highest activity of 5.4 × 104 g/((mol of Pd) h), generating oligomers with the lowest molecular weight (Mn: 0.3 kg/mol). Moreover, polar-functionalized propylene-MA co-oligomers with very high incorporation ratios (22.8–36.5 mol %) could be obtained in the copolymerization using these 8-alkyl-iminopyridyl Pd(II) catalysts. Additionally, Pd1 exhibited the best performance in propylene-MA copolymerization as it displayed the highest MA incorporation ratio of up to 36.5 mol%. All the three catalysts are capable of generating partially chain-straightened P-MA co-oligomers and the activities decrease gradually while the molecular weight increases with the increasing steric hindrance of the alkyl substituent (H < Me < n-Bu). Compared to Pd4 with the rigid 8-aryl substituent, the flexible 8-alkyl-iminopyridyl Pd(II) catalysts (Pd1-3) not only showed much higher activities in the propylene oligomerization, but also yielded P-MA co-oligomers with significantly higher incorporation ratios in the propylene co-oligomerization.

1. Introduction

Since polypropylene was first synthesized in 1950s, this type of thermoplastic polymer has become one of the most used plastic products worldwide [1,2]. The excellent chemical and mechanical properties of polypropylene enable it to be used in packaging, manufacturing, the automotive industry, and household appliances [3]. Regardless of its widespread usage, the nonpolar properties of polypropylene limits its utilization. To improve the polarity of polypropylene for adhesion, dye-ability and compatibility, functional groups were always introduced into the material [4,5]. A post-modification technique was commonly used in industry. However, a deterioration of the accompanying properties always happened due to the processes occurring under harsh conditions, such as extreme heat, irradiation with high energy, and excessive etching. Additionally, the uncontrollable functional group distribution and side reactions would always diminish the material’s quality [4,5]. Therefore, propylene and functional vinyl monomers copolymerization using transition metal catalysts is the most straightforward and economic strategy to generate functional polypropylenes. Early-transition-metal catalysts exhibit highly efficient propylene polymerization. Additionally, Ziegler-Natta catalysts, metallocene catalysts, and other early-transition-metal catalysts were always easily deactivated by functional-group-containing polar monomers, although several cases have shown they could be applied in polar functional monomer and olefin copolymerization when the incorporated polar monomers were protected or tied up by Lewis acid as a “masking reagent” [6,7,8,9,10,11,12,13]. Recently, several cases showed a few cationic metallocene and (pyridylamido)Hf complexes could directly copolymerize propylene and polar monomers bearing the special structure without a “masking reagent” [14,15,16,17].
In the last three decades, the development of late-transition-metal catalysts in olefin polymerization has been witnessed owing to their great capability of olefin–polar monomer copolymerization. Compared to early-transition metals, late-transition metal (palladium and nickel) complexes have higher polar functional group tolerance [18,19,20]. Since Brookhart et al. initially reported α-diimine bearing Ni(II) and Pd(II) complexes could be utilized in olefin (co)polymerization in the 1990s, a series of ligands with imine structures have been developed [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36]. The polyolefin characters including molecular weight, polymer dispersity index (PDI), and microstructure not only depend on the polymerization conditions, but the coordinated complex ligands [37,38,39,40,41]. Among the α-diimine-based analogs, iminopyridyl-based Ni(II) and Pd(II) complexes have been noticed. The obtained oligomers using Ni(II) and Pd(II) catalysts bearing the iminopyridyl ligands have been reported in most ethylene (co)polymerizations [42,43,44,45,46,47,48]. When the ligands with an iminopyridyl structure are applied in propylene polymerization, the unilateral axial block provided by the imine motif would partially shield the metal ion center and result in the low molecular weight of the polymer [42,43,44,45,46,47,48]. Recently, we introduced a 1,5-di(dibenzosuberyl)amine structure into the iminopyridyl catalysts. The generated Ni(II)/Pd(II) complexes were apt to retard the chain transfer in polymerization. Therefore, the high-molecular-weight (co)polymers were produced [49,50,51,52]. We also found the Ni(II) and Pd(II) complexes bearing iminopyridyl ligands with dibenzhydryl and 8-alkyl/aryl-naphthyl substituents could induce ethylene (co)polymerization to form high-molecular-weight (co)polymers [52,53,54]. Moreover, these imine-based catalysts exhibited significant advantages in the copolymerization of propylene with polar comonomers and generated functionalized polypropylene materials (Chart 1) [55,56,57,58]. In this work, we describe the catalytic performance of propylene (co)polymerization using iminopyridyl Pd(II) catalysts (Chart 1h) composed of various 8-alkylnaphthyl substituents.

2. Experimental Section

2.1. General Procedures and Materials

All the reactions and polymerizations were carried out using standard Schlenk techniques or a glovebox under a dry nitrogen atmosphere. Anhydrous n-hexane, toluene, dichloromethane (DCM) and methyl acrylate were purchased from Ennegy Chemicals (Shanghai, China) and Sinopharm (Shanghai, China). All the solvents and monomers are dried with CaH2. The propylene gas was obtained from Nanjing Special Gas Co. (Nanjing, China) and applied in reaction without purification. The complexes Pd1-4 were synthesized according to the methods reported before [52,54]. The deuterated solvents were dried and distilled with CaH2. Nuclear magnetic resonance (NMR) spectra (1H: 600 MHz, 13C: 150 MHz) were recorded on a JNM-ECZ600R NMR instrument (Nippon Electronic Company, Tokyo, Japan) at ambient temperature unless otherwise stated. The chemical shifts observed in the 1H and 13C NMR spectra were referenced to the residual resonance of deuterated solvents (the coupling constants are reported in Hz, CDCl3: 1H, 7.26 ppm, 13C, 77.16 ppm). The molecular weight and molecular-weight distribution of polypropylene (PP), propylene oligomers, and propylene–MA co-oligomers (P–MA) were determined by size exclusion chromatography (SEC) analyses, which were carried out with two linear Styragel SEC columns containing Tosoh equipment (Tosoh Corporation, Tokyo, Japan) and eluted with THF. The system was calibrated with the polystyrene standards and operated at 40 °C at a flow rate of 0.35 mL/min with THF.

2.2. Procedures of Propylene (Co)Polymerization

2.2.1. Propylene Polymerization

A 350 mL thick-walled pressure vessel together with a magnetic stirrer was charged with 40 mL DCM in the glovebox. The pressure vessel was connected to a Schlenk line bearing high pressure. A solution of Pd1-4 (10 μmol) with NaBArF (20 μmol, 2.0 equiv.) in 2 mL of CH2Cl2 was introduced to the vessel via a syringe at room temperature. The reactor was pressurized and maintained at 4 atm propylene pressure with rapid stirring. The mixture was stirred for 3 h at the reaction temperature. After cooling down, the reactor was quenched in air. Then, the synthesized polymer was obtained via evaporation under vacuum until the weight remained constant.

2.2.2. Copolymerization of Propylene and MA

A 350 mL thick-walled pressure vessel with a stirring bar was charged with an appropriate amount of DCM, MA, and NaBArF in the glovebox. The pressure vessel was connected to a Schlenk line bearing high pressure. The Pd catalysts (20 μmol) in 2 mL CH2Cl2 were syringed into the well-stirred solution with the total reaction volume kept at 20 mL. The reactor was pressurized with propylene and maintained at 4 atm pressure. The copolymerization was terminated by passing through a pad of silica after continuous stirring for 12 h. Then, the formed copolymers were obtained under vacuum to an unchanged weight. XMA% = 3/(Iall − 3)×100%. Iall: total H integral; IOCH3 = 3: OCH3 hydrogen integral (ca. 3.5–5.0 ppm).

3. Results and Discussions

Propylene Polymerization

All of the Pd(II) complexes (Pd1-4) (Chart 2) were active in propylene polymerization after in situ dechlorination with 2.0 equiv. sodium tetrakis(3,5-bis(trifluoromethyl)phenyl)borate(NaBArF). The polymerization results are summarized in Table 1. All the 8-alkylnaphthyl Pd(II) complexes showed moderate catalytic activities (21.7–54 kg mol−1 h−1), and highly branched propylene oligomers (119–136/1000 C) with various molecular weights (0.3–1.5 kg mol−1) were obtained at given conditions (Table 1, Figure 1). Compared to Pd4 with the rigid 8-arylnaphthyl structure, Pd1-3 complexes bearing the 8-alkylnaphthyl structure with higher flexibility exhibited up to ten times higher activities with significantly lower molecular weights. This result is close to the ethylene polymerization results, in which palladium complexes bearing the 8-alkyl group also exhibited activities that were orders of magnitude higher [54]. The axial steric hindrance provided by the flexible substituents would probably facilitate the coordination and insertion of the propylene molecule [35,36,57]. The further comparison of Pd1, Pd2 and Pd3 revealed the higher molecular weight of propylene oligomers could be obtained using the complexes bearing the longer the alkyl group chain (entries 5–6 vs. 3–4 vs. 1–2, Table 1). This may be due to the axial steric effect. A more effective axial steric hindrance provided by longer alkyl substitutes would retard the chain transfer reaction more effectively in polymerization [38,54]. In addition, the activities of all the complexes, Pd14, increased 1.6–2.7 times when the polymerization temperature increased, while the molecular weight decreased (entries 1–8, Table 2). The significantly reduced energy barrier of propylene insertion into the active palladium species with these alkyl groups, and the increased ratio of chain transfer rate to chain growth rate are responsible for the results observed above.
Most interestingly, a chain walking phenomenon was also found in iminopyridyl palladium-catalyzed propylene polymerization (Figure 2). In contrast to the chain-walking-lead branch formation in ethylene polymerization, chain walking drives partial chain straightening in propylene polymerization (Figure 3). Theoretically, being derived from the methyl group containing monomer, polypropylene has a branching density of 333/1000 C. All the obtained propylene oligomers or polypropylenes in this study are 119–136/1000 C, which are significantly lower than 333/1000 C. This indicates that a remarkable chain straightening occurs during the polymerization. The calculated value of 1,3 chain straightening is around 0.49–0.54 (Table 1) for all the obtained propylene oligomers or polypropylenes, which indicates almost half of the propylene molecules experience 1,3 chain straightening. The chain walking in polymerization would facilitate the propylene insertion (Figures S1–S13). The driving force for chain walking is mainly due to the reduction in the steric hindrance during the insertion of propylene molecules in the polymerization process [59]. The 13C NMR spectrum analysis of a propylene oligomer yield with Pd3 at 30 °C is shown in Figure 3 [59,60]. Branch structures observed include methyl, 2-methylpropyl+, 2-methyl(CH2)n+, and 2,4-dimethylpentyl+. The adjacent-methyl and isolated-methyl branches are also observed.
The performance of the iminopyridyl Pd(II) complexes in propylene-MA copolymerization was also explored (Table 2). In this study, polar-functionalized propylene-MA co-oligomers with relatively high values of incorporation ratios (up to 36.5 mol %) were obtained (Table 2). The Pd(II) catalysts showed low activities (level of 103 g mol−1 h−1), which were remarkably lower than those in propylene homo-polymerization. As expected, the higher MA concentration resulted in higher MA incorporation, but reduced activities. Compared to propylene homo-polymerization, the molecular weight was dramatically decreased when MA was supplied. However, the molecular weight was not further decreased with the extra MA supplement (Table 2, entries 1, 3, 5, 7 vs. 2, 4, 6, 8). The substantial reduction in molecular weights implied the insertion of excess polar monomers might greatly suppress the growth of the polymer chain and facilitate chain transfer during polymerization (Figures S14–S19) [49,52]. The propylene-MA co-oligomers obtained with Pd2-4 containing substituted alkyl/aryl group tended to produce co-oligomers with higher molecular weights compared to Pd1 (entries 1–2 vs. 3–8, Table 2). It demonstrates the steric effect is crucial for polymer chain propagation over chain transfer, which is similar to the propylene homo-polymerization result. The propylene-MA co-oligomer obtained with the 8-alkylnaphthyl Pd1-3 tended to generate co-oligomers with higher incorporation ratios compared to 8-arylnaphthyl Pd4 (entries 1–6 vs. 7–8, Table 2), which suggests the complexes with flexible alkyl substituents have the advantage of MA insertion.

4. Conclusions

In conclusion, we described propylene (co)polymerization using a series of 8-alkyl-1-naphthyl iminopyridyl Pd(II) catalysts. Compared to the exhibition of 8-arylnaphthyl Pd4 in propylene polymerization, all the 8-alkylnaphthyl Pd1-3 catalysts showed up to ten times higher activities, but generated propylene oligomers with significantly lower molecular weights. In the corresponding Pd(II)-catalyzed propylene-MA copolymerization, the 8-alkylnaphthyl Pd1-3 catalysts displayed higher MA incorporation ratios with similar molecular weights to 8-arylnaphthyl Pd4. Most notably, all the produced polypropylenes and propylene (co)oligomers are partial chain straightening, with almost half of the inserted propylene monomers undergoing 1,3 chain straightening. These propylene oligomers and propylene-MA co-oligomers could possibly be utilized as functional additives in surface modifiers or lubricants.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15010111/s1, Figures S1–S13: 1H and 13C NMR of some representative propylene oligomers and P-MA co-oligomers; Figures S14–S19: SEC of some representative propylene oligomers and P-MA co-oligomers.

Author Contributions

Conceptualization, S.D.; methodology, H.S.; validation, S.D., H.S. and W.Z.; formal analysis, S.D. and W.Z.; investigation, H.S.; resources, S.D. and C.Z.; data curation, S.D.; writing—original draft preparation, W.Z. and H.F.; writing—review and editing, S.D.; supervision, S.D.; project administration, S.D.; funding acquisition, S.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Natural Science Foundation of Anhui Province (2108085Y06), Anhui Provincial Key Laboratory Open Project Foundation (LCECSC-01), and Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology (BM2012110).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Galli, P.; Vecellio, G. Technology: Driving force behind innovation and growth of polyolefins. Prog. Polym. Sci. 2001, 26, 1287. [Google Scholar] [CrossRef]
  2. Coates, G.W. Precise control of polyolefin stereochemistry using single-site metal catalysts. Chem. Rev. 2000, 100, 1223. [Google Scholar] [CrossRef] [PubMed]
  3. Stuerzel, M.; Mihan, S.; Muelhaupt, R. From multisite polymerization catalysis to sustainable materials and all-polyolefin composites. Chem. Rev. 2016, 116, 1398. [Google Scholar] [CrossRef] [PubMed]
  4. Boaen, N.K.; Hillmyer, M.A. Post-polymerization functionalization of polyolefins. Chem. Soc. Rev. 2005, 34, 267. [Google Scholar] [CrossRef]
  5. Nakamura, A.; Ito, S.; Nozaki, K. Coordination-insertion copolymerization of fundamental polar monomers. Chem. Rev. 2009, 109, 5215. [Google Scholar] [CrossRef]
  6. Zhang, M.; Yuan, X.; Wang, L.; Chung, T.C.M.; Huang, T.; De Groot, W. Synthesis and characterization of well-controlled isotactic polypropylene ionomers containing ammonium ion groups. Macromolecules 2014, 47, 571. [Google Scholar] [CrossRef]
  7. Zhang, G.; Li, H.; Antensteiner, M.; Chung, T.C.M. Synthesis of functional polypropylene containing hindered phenol stabilizers and applications in metallized polymer film capacitors. Macromolecules 2015, 48, 2925. [Google Scholar] [CrossRef]
  8. Dong, J.Y.; Wang, Z.M.; Hong, H.; Chung, T.C. Synthesis of isotactic polypropylene containing a terminal Cl, OH, or NH2 group via metallocene-mediated polymerization/chain transfer reaction. Macromolecules 2002, 35, 9352. [Google Scholar] [CrossRef]
  9. Hagihara, H.; Tsuchihara, K.; Sugiyama, J.; Takeuchi, K.; Shiono, T. Copolymerization of propylene and polar allyl monomer with zirconocene/methylaluminoxane catalyst: Catalytic synthesis of amino-terminated isotactic polypropylene. Macromolecules 2004, 37, 5145. [Google Scholar] [CrossRef]
  10. Hagihara, H.; Ishihara, T.; Ban, H.T.; Shiono, T. Precise control of microstructure of functionalized polypropylene synthesized by the ansa-zirconocene/MAO catalysts. J. Polym. Sci. Part A: Polym. Chem. 2008, 46, 1738. [Google Scholar] [CrossRef]
  11. Hagihara, H.; Tsuchihara, K.; Sugiyama, J.; Takeuchi, K.; Shiono, T. Copolymerization of 3-buten-1-ol and propylene with an isospecific zirconocene/methylaluminoxane catalyst. J. Polym. Sci. Part A Polym. Chem. 2004, 42, 5600. [Google Scholar] [CrossRef]
  12. Kaya, A.; Jakisch, L.; Komber, H.; Pompe, G.; Pionteck, J.; Voit, B.; Schulze, U. Synthesis of oxazoline functionalized polypropene using metallocene catalysts. Macromol. Rapid Commun. 2000, 21, 1267. [Google Scholar] [CrossRef]
  13. Wilen, C.E.; Nasman, J.H. Polar activation in copolymerization of propylene and 6-tert-butyl-[2-(1,1-dimethylhept-6-enyl)]-4-methylphenol over a racemic [1,1′-(dimethylsilylene)bis(η5-4,5,6,7-tetrahydro-1-indenyl)]zirconium dichloride/methylalumoxane catalyst system. Macromolecules 1994, 27, 4051. [Google Scholar] [CrossRef]
  14. Wang, X.; Wang, Y.; Shi, X.; Liu, J.; Chen, C.; Li, Y. Syntheses of well-defined functional isotactic polypropylenes via efficient copolymerization of propylene with ω-halo-α-alkenes by post-metallocene hafnium catalyst. Macromolecules 2014, 47, 552. [Google Scholar] [CrossRef]
  15. Wang, X.; Long, Y.; Wang, Y.; Li, Y. Insights into propylene/ω-halo-α-alkenes copolymerization promoted by rac-Et(Ind)2ZrCl2 and (pyridyl-amido)hafnium catalysts. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 3421. [Google Scholar] [CrossRef]
  16. Shang, R.; Gao, H.; Luo, F.; Li, Y.; Wang, B.; Ma, Z.; Pan, L.; Li, Y. Functional isotactic polypropylenes via efficient direct copolymerizations of propylene with various amino-functionalized α-olefins. Macromolecules 2019, 52, 9280. [Google Scholar] [CrossRef]
  17. Huang, M.; Chen, J.; Wang, B.; Huang, W.; Chen, H.; Gao, Y.; Marks, T.J. Polar isotactic and syndiotactic polypropylenes by organozirconium-catalyzed masking-reagent-free propylene and amino-olefin copolymerization. Angew. Chem. Int. Ed. 2020, 59, 20522. [Google Scholar] [CrossRef]
  18. Ittel, S.D.; Johnson, L.K.; Brookhart, M. Late-metal catalysts for ethylene homo- and copolymerization. Chem. Rev. 2000, 100, 1169. [Google Scholar] [CrossRef]
  19. Mu, H.; Zhou, G.; Hu, X.; Jian, Z. Recent advances in nickel mediated copolymerization of olefin with polar monomers. Coord. Chem. Rev. 2021, 435, 213802. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Zhang, Y.; Hu, X.; Wang, C.; Jian, Z. Advances on controlled chain walking and suppression of chain transfer in catalytic olefin polymerization. ACS Catal. 2022, 12, 14304. [Google Scholar] [CrossRef]
  21. Johnson, L.K.; Killian, C.M.; Brookhart, M. New Pd(II)- and Ni(II)-based catalysts for polymerization of ethylene and α-olefins. J. Am. Chem. Soc. 1995, 117, 6414. [Google Scholar] [CrossRef]
  22. Liao, Y.; Cai, Q.; Dai, S. Synthesis of high molecular weight polyethylene and E-MA copolymers using iminopyridine Ni(II) and Pd(II) complexes containing a flexible backbone and rigid axial substituents. Chin. J. Polym. Sci. 2022. [Google Scholar] [CrossRef]
  23. Lu, Z.; Liao, Y.; Fan, W.; Dai, S. Efficient suppression of chain transfer reaction in ethylene coordination polymerization with dibenzosuberyl substituents. Polym. Chem. 2022, 13, 4090. [Google Scholar] [CrossRef]
  24. Lu, W.; Liao, Y.; Dai, S. Facile access to ultra-highly branched polyethylenes using hybrid “sandwich” Ni(II) and Pd(II) catalysts. J. Catal. 2022, 411, 54. [Google Scholar] [CrossRef]
  25. Lu, W.; Wang, H.; Fan, W.; Dai, S. Exploring the relationship between polyethylene microstructure and spatial structure of α-diimine Pd(II) catalysts via a hybrid steric strategy. Inorg. Chem. 2022, 61, 6799. [Google Scholar] [CrossRef] [PubMed]
  26. Zhong, L.; Li, G.; Liang, G.; Gao, H.; Wu, Q. Enhancing thermal stability and living fashion in α-diimine-nickel-catalyzed (co)polymerization of ethylene and polar monomer by increasing the steric bulk of ligand backbone. Macromolecules 2017, 50, 2675. [Google Scholar] [CrossRef]
  27. Zhong, S.; Tan, Y.; Zhong, L.; Gao, J.; Liao, H.; Jiang, L.; Gao, H.; Wu, Q. Precision Synthesis of Ethylene and Polar Monomer Copolymers by Palladium-Catalyzed Living Coordination Copolymerization. Macromolecules 2017, 50, 5661. [Google Scholar] [CrossRef]
  28. Hu, X.; Zhang, Y.; Li, B.; Jian, Z. Horizontally and vertically concerted steric strategy in α-diimine nickel promoted ethylene (Co)polymerization. Chin. J. Chem. 2021, 39, 2829. [Google Scholar] [CrossRef]
  29. Xia, J.; Zhang, Y.; Kou, S.; Jian, Z. A concerted double-layer steric strategy enables an ultra-highly active nickel catalyst to access ultrahigh molecular weight polyethylenes. J. Catal. 2020, 390, 30. [Google Scholar] [CrossRef]
  30. 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. [Google Scholar] [CrossRef]
  31. Liu, Y.-S.; Harth, E. Distorted sandwich α-diimine Pd(II) catalyst: Linear polyethylene and synthesis of ethylene/acrylate elastomers. Angew. Chem., Int. Ed. 2021, 60, 24107. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, D.; Nadres, E.T.; Brookhart, M.; Daugulis, O. Synthesis of highly branched polyethylene using “sandwich” (8-p-tolyl naphthyl α-diimine)nickel(II) catalysts. Organometallics 2013, 32, 5136. [Google Scholar] [CrossRef]
  33. Li, S.; Dai, S. 8-Arylnaphthyl substituent retarding chain transfer in insertion polymerization with unsymmetrical α-diimine systems. Polym. Chem. 2020, 11, 7199. [Google Scholar] [CrossRef]
  34. Lu, Z.; Chang, G.; Jing, K.; Dai, S. A dual steric enhancement strategy in α-diimine nickel and palladium catalysts for ethylene polymerization and copolymerization. Organometallics 2022, 41, 124. [Google Scholar] [CrossRef]
  35. Guo, L.; Sun, W.; Li, S.; Xu, G.; Dai, S. Bulky yet flexible substituents in insertion polymerization with α-diimine nickel and palladium systems. Polym. Chem. 2019, 10, 4866. [Google Scholar] [CrossRef]
  36. 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. [Google Scholar] [CrossRef]
  37. Gates, D.P.; Svejda, S.A.; Oñate, 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. [Google Scholar] [CrossRef]
  38. Gong, Y.; Li, S.; Gong, Q.; Zhang, S.; Liu, B.; Dai, S. Systematic investigations of ligand steric effects on α-diimine nickel catalyzed olefin polymerization and copolymerization. Organometallics 2019, 38, 2919. [Google Scholar] [CrossRef]
  39. Wiedemann, T.; Voit, G.; Tchernook, A.; Roesle, P.; Go, I.; Mecking, S. Monofunctional hyperbranched ethylene oligomers. J. Am. Chem. Soc. 2014, 136, 2078. [Google Scholar] [CrossRef]
  40. Wang, H.; Duan, G.; Fan, H.; Dai, S. Second coordination sphere effect of benzothiophene substituents on chain transfer and chain walking in ethylene insertion polymerization. Polymer 2022, 245, 124707. [Google Scholar] [CrossRef]
  41. Guo, L.; Liu, Y.; Lian, K.; Sun, W.; Zhu, H.; Du, Q.; Liu, Z.; Chen, X.; Dai, S. Electronic effects of the backbone on bis(imino)pyridyliron(II)-catalyzed ethylene polymerization. Eur. J. Inorg. Chem. 2018, 2018, 4887. [Google Scholar] [CrossRef]
  42. Wu, X.; Xu, G.; Lu, W.; Li, Z.-Y.; Dai, S. Ethylene (Co)oligomerization in alkane solvents facilitated by rigid-flexible double-layer steric strategy. Eur. Polym. J. 2022, 177, 111459. [Google Scholar] [CrossRef]
  43. Chen, J.; Yan, Z.; Li, Z.; Dai, S. Direct synthesis of chain-end toluene functionalized hyperbranched ethylene oligomers. Polymers 2022, 14, 3049. [Google Scholar] [CrossRef] [PubMed]
  44. Ding, B.; Chang, G.; Yan, Z.; Dai, S. Ethylene (Co)oligomerization using iminopyridyl Ni(II) and Pd(II) complexes bearing benzocycloalkyl moieties to access hyperbranched ethylene oligomers and ethylene-MA co-oligomers. Front. Chem. 2022, 10, 961426. [Google Scholar] [CrossRef] [PubMed]
  45. Fan, H.; Xu, G.; Wang, H.; Dai, S. Direct synthesis of hyperbranched ethylene oligomers and ethylene-MA co-oligomers using iminopyridyl systems with weak neighboring group interactions. J. Polym. Sci. 2022, 60, 1944. [Google Scholar] [CrossRef]
  46. Yan, Z.; Bi, H.; Ding, B.; Wang, H.; Xu, G.; Dai, S. A rigid-flexible double-layer steric strategy for ethylene (Co)oligomerization with pyridine-imine Ni(II) and Pd(II) complexes. New J. Chem. 2022, 46, 8669. [Google Scholar] [CrossRef]
  47. Fan, H.; Chang, G.; Bi, H.; Gui, X.; Wang, H.; Xu, G.; Dai, S. Facile Synthesis of hyperbranched ethylene oligomers and ethylene/methyl acrylate co-oligomers with different microscopic chain architectures. ACS Polym. Au 2022, 2, 88. [Google Scholar] [CrossRef]
  48. Li, S.; Lu, Z.; Fan, W.; Dai, S. Efficient incorporation of a polar comonomer for direct synthesis of hyperbranched polar functional ethylene oligomers. New J. Chem. 2021, 45, 4024. [Google Scholar] [CrossRef]
  49. Ge, Y.; Li, S.; Wang, H.; Dai, S. Synthesis of branched polyethylene and ethylene-MA copolymers using unsymmetrical iminopyridyl nickel and palladium complexes. Organometallics 2021, 40, 3033. [Google Scholar] [CrossRef]
  50. Ge, Y.; Cai, Q.; Wang, Y.; Gao, J.; Chi, Y.; Dai, S. Synthesis of high-molecular-weight branched polyethylene using a hybrid “sandwich” pyridine-imine Ni(II) catalyst. Front. Chem. 2022, 10, 886888. [Google Scholar] [CrossRef]
  51. Peng, H.; Li, S.; Li, G.; Dai, S.; Ji, M.; Liu, Z.; Guo, L. Rotation-restricted strategy to synthesize high molecular weight polyethylene using iminopyridyl nickel and palladium catalyst. Appl. Organomet. Chem. 2021, 35, e6140. [Google Scholar] [CrossRef]
  52. Li, S.; Dai, S. Highly efficient incorporation of polar comonomers in copolymerizations with ethylene using iminopyridyl palladium system. J. Catal. 2021, 393, 51. [Google Scholar] [CrossRef]
  53. Dai, S.; Li, S. Effect of aryl orientation on olefin polymerization in iminopyridyl catalytic system. Polymer 2020, 200, 122607. [Google Scholar] [CrossRef]
  54. Ge, Y.; Lu, Z.; Dai, S. Flexible axial shielding strategy for the synthesis of high-molecular-weight polyethylene and polar functionalized polyethylene with pyridine-imine Ni(II) and Pd(II) complexes. Organometallics 2022, 41, 2042. [Google Scholar] [CrossRef]
  55. Luckham, S.L.J.; Nozaki, K. Toward the copolymerization of propylene with polar comonomers. Acc. Chem. Res. 2021, 54, 344. [Google Scholar] [CrossRef]
  56. Johnson, L.K.; Mecking, S.; Brookhart, M. Copolymerization of ethylene and propylene with functionalized vinyl monomers by palladium(II) catalysts. J. Am. Chem. Soc. 1996, 118, 267. [Google Scholar] [CrossRef]
  57. Fan, H.; Liao, Y.; Dai, S. Propylene polymerization and copolymerization with polar monomers facilitated by flexible cycloalkyl substituents in α-diimine systems. Polymer 2022, 254, 125076. [Google Scholar]
  58. Lu, Z.; Wang, H.; Li, S.; Dai, S. Direct synthesis of various polar functionalized polypropylene materials with tunable molecular weights and high incorporation ratios. Polym. Chem. 2021, 12, 5495. [Google Scholar] [CrossRef]
  59. McCord, E.F.; McLain, S.J.; Nelson, L.T.J.; Arthur, S.D.; Coughlin, E.B.; Ittel, S.D.; Johnson, L.K.; Tempel, D.; Killian, C.M.; Brookhart, M. 13C and 2D NMR analysis of propylene polymers made with α-diimine late metal catalysts. Macromolecules 2001, 34, 362. [Google Scholar] [CrossRef]
  60. Svejda, S.A.; Johnson, L.K.; Brookhart, M. Low-temperature spectroscopic observation of chain growth and migratory insertion barriers in (α-diimine)Ni(II) olefin polymerization catalysts. J. Am. Chem. Soc. 1999, 121, 10634. [Google Scholar] [CrossRef]
Chart 1. The α-diimine (ad) and iminopyridyl (eh) palladium complexes used for propylene (co)polymerization and this work (h).
Chart 1. The α-diimine (ad) and iminopyridyl (eh) palladium complexes used for propylene (co)polymerization and this work (h).
Polymers 15 00111 ch001
Chart 2. The iminopyridyl Pd(II) catalysts with various substituents at 8 position of naphthyl iminopyridyl ligands (Pd1: H; Pd2: Me; Pd3: butyl; Pd4: p-tolyl) were used for the propylene (co)polymerization.
Chart 2. The iminopyridyl Pd(II) catalysts with various substituents at 8 position of naphthyl iminopyridyl ligands (Pd1: H; Pd2: Me; Pd3: butyl; Pd4: p-tolyl) were used for the propylene (co)polymerization.
Polymers 15 00111 ch002
Figure 1. Comparisons on yield (a), molecular weight (b), and branching density (c) of propylene oligomers or polypropylene yielded with catalysts Pd1–Pd4 at 30 °C (blue) and 50 °C (red).
Figure 1. Comparisons on yield (a), molecular weight (b), and branching density (c) of propylene oligomers or polypropylene yielded with catalysts Pd1–Pd4 at 30 °C (blue) and 50 °C (red).
Polymers 15 00111 g001
Figure 2. The general mechanism of propylene chain walking polymerization with iminopyridyl Pd(II) catalysts.
Figure 2. The general mechanism of propylene chain walking polymerization with iminopyridyl Pd(II) catalysts.
Polymers 15 00111 g002
Figure 3. 13C NMR spectrum analysis of the partially chain-straightened propylene oligomer obtained by using Pd3 at 30 °C (Table 1, entry 5). Branches are labeled as xBy, where y is the branch length and x is the carbon, starting from the methyl end with 1. The methine groups for the different branch lengths are labeled with brBy. The asterisk represents the situation when the methyl groups are in adjacent positions to each other.
Figure 3. 13C NMR spectrum analysis of the partially chain-straightened propylene oligomer obtained by using Pd3 at 30 °C (Table 1, entry 5). Branches are labeled as xBy, where y is the branch length and x is the carbon, starting from the methyl end with 1. The methine groups for the different branch lengths are labeled with brBy. The asterisk represents the situation when the methyl groups are in adjacent positions to each other.
Polymers 15 00111 g003
Table 1. Effect of Catalysts and Temperatures on Propylene Polymerization. a.
Table 1. Effect of Catalysts and Temperatures on Propylene Polymerization. a.
Ent.Precat.T/°CYield/gAct. bMncMw/MncB d[CH3]/[CH2] e%1,3 f
1Pd1300.973.230.62.001190.2254
2Pd1501.625.400.31.271250.2353
3Pd2300.652.171.01.511360.2649
4Pd2501.103.670.91.531290.2451
5Pd3300.732.431.51.781210.2254
6Pd3501.294.301.41.591300.2451
7Pd4300.110.3713.62.431200.2254
8Pd4500.260.8710.01.571240.2353
a Reaction conditions: Pd catalyst (10 μmol), NaBArF (2.0 equiv.), propylene (4 atm), CH2Cl2 (40 mL), polymerization time (3 h). b Activity = 104 g/((mol of Pd) h). c Mn is in units of kg mol−1. Determined by SEC in THF at 40 °C vs. polystyrene standards. d B = Number of branches per 1000 C, as determined by 1H NMR spectroscopy. e Determined by 1H NMR spectroscopy. f 1,3-enchainment, calculated from the equation: %1,3-enchainment = [(1 − R)/(1 + 2R)] × 100, where R = [CH3]/[CH2].
Table 2. Propylene-MA copolymerization. a.
Table 2. Propylene-MA copolymerization. a.
Ent.Precat.[MA]Yield (g)Act. bXMAc(%)MndMw/MndB e
1Pd11 0.441.8334.30.31.28119
2Pd120.371.5436.50.31.28152
3Pd210.341.4222.80.51.37108
4Pd220.220.9225.10.51.21154
5Pd310.110.4625.30.61.32126
6Pd320.090.3833.80.51.20142
7Pd410.170.7113.80.51.21105
8Pd420.130.5423.00.51.16151
a General conditions: Pd catalysts (20 μmol), NaBArF (2.0 equiv.), propylene (4 atm), polymerization time (12 h), the total volume of CH2Cl2 and MA (20 mL), polymerization temperature (40 °C), b Activity = 103 g/((mol of Pd) h), c XMA = Incorporation of MA, d Mn is in units of kg mol–1. Determined by SEC in THF at 40 °C vs. polystyrene standards, e B = Number of branches per 1000 C, as determined by 1H NMR spectroscopy. The branches ending with functional groups are added to the total branches.
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

Sun, H.; Fan, H.; Zhu, C.; Zou, W.; Dai, S. Direct Synthesis of Partially Chain-Straightened Propylene Oligomers and P-MA Co-Oligomers Using Axially Flexible Shielded Iminopyridyl Palladium Complexes. Polymers 2023, 15, 111. https://doi.org/10.3390/polym15010111

AMA Style

Sun H, Fan H, Zhu C, Zou W, Dai S. Direct Synthesis of Partially Chain-Straightened Propylene Oligomers and P-MA Co-Oligomers Using Axially Flexible Shielded Iminopyridyl Palladium Complexes. Polymers. 2023; 15(1):111. https://doi.org/10.3390/polym15010111

Chicago/Turabian Style

Sun, Huayin, Huijun Fan, Chuangao Zhu, Wenping Zou, and Shengyu Dai. 2023. "Direct Synthesis of Partially Chain-Straightened Propylene Oligomers and P-MA Co-Oligomers Using Axially Flexible Shielded Iminopyridyl Palladium Complexes" Polymers 15, no. 1: 111. https://doi.org/10.3390/polym15010111

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