Next Article in Journal
5000 h Multi-Factor Accelerated Aging Test of FRP Made Transmission Tower: Characterization, Thermal Decomposition and Reaction Kinetics Study
Next Article in Special Issue
Isothermal and Non-Isothermal Crystallization Studies of Long Chain Branched Polypropylene Containing Poly(ethylene-co-octene) under Quiescent and Shear Conditions
Previous Article in Journal
Soy-Based Adhesive Cross-Linked by Phenol-Formaldehyde-Glutaraldehyde
Previous Article in Special Issue
A Modified Ceramic-Coating Separator with High-Temperature Stability for Lithium-Ion Battery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Ligand Backbone Structure and Connectivity on the Properties of Phosphine-Sulfonate Pd(II)/Ni(II) Catalysts

Chinese Academy of Sciences Key Laboratory of Soft Matter Chemistry, Department of Polymer Science and Engineering, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Polymers 2017, 9(5), 168; https://doi.org/10.3390/polym9050168
Submission received: 1 April 2017 / Revised: 28 April 2017 / Accepted: 3 May 2017 / Published: 9 May 2017
(This article belongs to the Special Issue Olefin Polymerization and Polyolefin)

Abstract

:
Phosphine-sulfonate based palladium and nickel catalysts have been extensively studied in ethylene polymerization and copolymerization reactions. Previously, the majority of the research works focused on the modifications of the substituents on the phosphorous atom. In this contribution, we systematically demonstrated that the change of the ligand backbone from benzene to naphthalene could greatly improve the properties of this class of catalysts. In the palladium system, this change could increase catalyst stability and polyethylene molecular weights. In the nickel system, this change could dramatically increase the polyethylene molecular weights. Most interestingly, the change in the connectivity of phosphine and sulfonate moieties to the naphthalene backbone could also significantly influence the catalyst properties.

Graphical Abstract

1. Introduction

In olefin polymerization, late transition metal catalysts have attracted much attention because of their low oxophilicity, and correspondingly the potentials to incorporation polar functionalized monomers into polyolefins. Among the numerous late transition metal catalysts, the Brookhart type α-diimine Ni(II) and Pd(II) [1,2,3,4,5,6,7,8,9,10,11,12,13,14], phenoxyminato based Ni(II) [15,16,17,18,19,20,21] and the phosphine-sulfonate Pd(II) catalysts [22,23,24,25,26,27,28,29,30,31,32,33,34,35] are the most extensively studied systems. It has been demonstrated that the properties of these catalysts are very sensitive to the ligand sterics. Specifically, it has been well established that the steric bulk on the axial positions in α-diimine systems could decrease the chain transfer rate, and increase the polyolefin molecular weight (Scheme 1, I) [1,2,36,37,38,39]. Similarly, the steric bulk on the axial positions in phenoxyminato (Scheme 1, II) and phosphine-sulfonate (Scheme 1, III) systems is crucial to obtain high-performance catalysts [40,41,42].
The modulation of the steric effect on the axial positions directly affects the steric environment of the metal center. In addition, the ligand backbone structure could indirectly influence the steric environment of the metal center, and correspondingly affect the properties of the metal catalysts. For example, different substituents on the backbone R position could greatly alter the properties of the α-diimine Pd(II) and Ni(II) catalysts (Scheme 1, IV) [43,44,45]. Despite the various efforts to modify the phosphine-sulfonate ligands, there have been very few studies on the modifications of the ligand backbone structures [46]. Recently, our group showed that the catalyst stability and activity could be greatly enhanced by changing the phosphine-sulfonate backbone from a benzene bridge to a naphthalene one (Scheme 1, V) [47]. In this contribution, we hope to further improve the properties of the naphthalene based phosphine-sulfonate Pd(II) and Ni(II) catalysts by: (1) changing the linking position of the phosphine and the sulfonate moieties on the naphthalene backbone; and (2) using a sterically very bulky bi-aryl substituent on the phosphorous atom.

2. Results and Discussion

Literature procedure was employed to prepare the ligands [47]. First, the 2-naphthalenesulfonic acid was converted to the toluidinium salt from the reaction with excess amount of p-toluidine (Scheme 2, see Supplementary Materials, Experimental Sections). The corresponding lithium salt was generated from the reaction with 1 equivalent of nBuLi, and dehydrated using a dean-stark apparatus in refluxing toluene. Subsequently, ligands L1-L3 were obtained in 39–47% yields from the reaction of 1 equivalent of nBuLi with the lithium salt in THF (tetrahydrofuran) followed by the addition of R2PCl. These ligands were characterized by by 1H, 13C, and 31P NMR (Nuclear Magnetic Resonance) spectroscopy (Bruker, Karlsruhe, Germany) (see Supplementary Materials, Figures S1–S9), elemental analysis and Mass spectrometry (EI+, Bruker Daltonics Inc., Billerica, MA, USA).
The reactions of ligands L1-L3 with (TMEDA)PdMe2 (TMEDA = tetramethylethylenediamine) in DMSO (dimethylsulfoxide) led to the formation of the Pd(II) complexes (Pd1-Pd3) at 41–48% yields (Scheme 2). The reaction of ligands L1-L3 with Na2CO3 and trans-[(PPh3)2Ni(Cl)Ph] afforded the desired Ni(II) complexes (Ni1-Ni3) at 34–76% yields. These metal complexes were characterized using 1H, 13C, and 31P NMR (see Supplementary Materials, Figures S10–S20), elemental analysis and MALDI-TOF-MS (Matrix Assisted Laser Desorption Ionization-Time of Flight-Mass Spectrometry). For comparison purpose, the palladium complex Pd2′/Pd2″ and the nickel complex Ni2′/Ni1″ were prepared according to literature procedures [48,49].
The molecular structures of Pd2 and Ni1 were determined by X-ray diffraction analysis (Figure 1; see CIF files and Supplementary Materials, Tables S1 and S2). The geometry at both the palladium and the nickel center is square planar with the methyl or phenyl substituent cis to the phosphine group. Clearly, the hydrogen atom on the C15 in Pd2 and C7 in Ni1 could exert some steric influence to the substituents on the phosphorous atom. Especially, this steric effect could be enhanced when the substituents are bulky. Most importantly, this interaction could potentially influence the properties of these catalysts in ethylene polymerization and copolymerization reactions.
The palladium catalysts are highly active in ethylene polymerization, with activities well above 105 g·mol−1·h−1 (Table 1, entries 1–6). Catalyst Pd3 with the biaryl substituent showed almost 10-fold increase in polymer molecular weight comparing with catalyst Pd1. The nickel catalysts are also highly active in ethylene polymerization, with activities comparable with those of the palladium catalysts (Table 1, entries 7–9). The palladium catalysts completely lost activity at 25 °C. However, the nickel catalysts could maintain high activity at 25 °C. Most importantly, the polyethylene molecular weight could be dramatically increased at lower polymerization temperature. For the case of Ni3, molecular weight of up to 142,300 could be achieved (Table 1, entry 12). Similar with the palladium case, catalyst Ni3 with the biaryl substituent showed much higher polymer molecular weight than catalysts Ni1 and Ni2.
Some very interesting results were obtained from the comparisons between catalysts Pd2/Ni2 and our previously reported catalysts Pd2′/Ni2′. Catalyst Pd2 showed similar activity and similar polymer molecular weight with catalyst Pd2′ (Table 1, entry 2 versus 13). The polyethylene molecular weights for catalysts Pd2 and Pd2′ are much higher than that of the conventional catalyst Pd2″ with the benzene as ligand backbone. In terms of catalyst stability, catalyst Pd2 showed slightly better stability than catalyst Pd2′, both of which are much more stable than conventional catalyst Pd2″ (Figure 2). Clearly, the change in the ligand backbone from benzene to naphthalene could significantly improve the performance of phosphine-sulfonate palladium catalysts. However, the change in the connectivity of the phosphine moiety and the sulfonate moiety to the naphthalene backbone does not influence the properties of these palladium catalysts.
In the nickel system, catalyst Ni2 showed polyethylene molecular weight twice as much as that of catalyst Ni2′ (Table 1, entry 8 versus 14, entry 11 versus 15). This suggested that the very small perturbations in ligand sterics could exert significant effect on the properties of the phosphine-sulfonate nickel catalysts. Most interestingly, more dramatically differences were observed for the cases of catalyst Ni1 versus catalyst Ni1″ bearing phenyl substituent on the phosphorous atom. No isolate solid polymer product was generated by catalyst Ni1″ in ethylene polymerization at either 80 or 25 °C. This agrees well with literature results [9]. In contrast, catalyst Ni1 demonstrated high activity (up to 1.9 × 105 g·mol−1·h−1), high polymer molecular weight (up to 37,200) and high melting temperature (134 °C) in ethylene polymerization. Cleary, the ligand backbone structure also plays an important role in determining the catalyst properties.
The palladium catalysts Pd1-Pd3 can also initiate efficient ethylene/metal acrylate copolymerization, with comonomer incorporation ratios ranging between 3% and 27% (Table 2, entries 1–6). The polymer molecular weights were dramatically reduced comparing with those in ethylene homopolymerization. Because of the great performance of catalyst Ni3 in ethylene homopolymerization, its properties in ethylene/polar monomer copolymerization were also investigated. Recently, Coates et al. showed that α-diimine nickel catalyst could mediate ethylene/methyl 10-undecenoate copolymerization in the presence of MAO (methylaluminoxane) [50]. Our groups showed that some sterically very bulky phosphine-sulfonate nickel catalysts could copolymerize ethylene with various polar monomers [51]. Here, Ni3 could also achieve moderate catalytic activity, along with moderate comonomer incorporation and high copolymer molecular weights in ethylene copolymerization with methyl 10-undecenoate and 6-chloro-1-hexene (Table 2, entries 7 and 8). The palladium complex Pd2″ with benzene backbone showed much lower activity and copolymer molecular weight than the corresponding complex Pd2 with the naphthalene backbone (Table 2, entries 9 and 10). Moreover, the nickel complex Ni1″ with benzene backbone is not active in the copolymerization (Table 2, entry 11).
Clearly, great enhancement in the polymerization properties was achieved by changing the ligand backbone from benzene to naphthalene (Scheme 3). Ligand electronic effect may play an important role. 1-Naphthalenesulfonic acid (pKa = 0.17 at 298 K in aqueous solution) is more acidic than benzenesulfonic acid (pKa = 0.70 at 298 K in aqueous solution) [52], suggesting that naphthalene based ligand is electronically more withdrawing than the benzene based ligand. Moreover, the bigger size of the naphthalene backbone may help to prevent catalyst deactivation reactions such as bis-ligation [53]. Furthermore, the potential interaction of the hydrogen atom on the 8 position of the naphthalene backbone with the phosphine or sulfonate substituent may also influence the catalyst properties. This steric effect may be the key factor in the differences between the two sets of catalysts with different connectivity to the naphthalene backbone.

3. Conclusions

To conclude, a series of phosphine-sulfonate based palladium and nickel catalysts were prepared and characterized. A naphthalene bridge was installed in the ligand framework. These palladium and nickel catalysts showed very high activities in ethylene polymerization. For the case of nickel system, very high polyethylene molecular weights could be achieved. Both the palladium and the nickel catalysts could initiate ethylene-polar monomer copolymerization. We clearly demonstrated the importance of ligand backbone structure and connectivity in determining the properties of these metal catalysts. This work provides an alternative strategy to modify/improve the properties of group 10 phosphine-sulfonate catalysts.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4360/9/5/168/s1. Experimental procedures, characterization for ligands, palladium complexes, nickel complexes (Figures S1–S20), polyethylene and copolymers (Figures S21–S56), and CIF files for complexes Pd2 and Ni1.

Acknowledgments

This work was supported by National Natural Science Foundation of China (NSFC, 21690071, and 51522306), the Fundamental Research Funds for the Central Universities (WK3450000001), and the Recruitment Program of Global Experts.

Author Contributions

Changle Chen conceived and designed the experiments; Zixia Wu, Changwen Hong, Hongxu Du, and Wenmin Pang performed the experiments; and Zixia Wu and Changle Chen analyzed the data and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, L.H.; Chen, C.L. (α-Diimine) palladium catalyzed ethylene polymerization and copolymerization with polar comonomers. Sci. China Chem. 2015, 58, 1663–1673. [Google Scholar] [CrossRef]
  2. Guo, L.H.; Dai, S.Y.; Sui, X.L.; Chen, C.L. Palladium and nickel catalyzed chain walking olefin polymerization and copolymerization. ACS Catal. 2016, 6, 428–441. [Google Scholar] [CrossRef]
  3. Guan, Z.; Cotts, P.M.; McCord, E.F.; McLain, S.J. Chain walking: A new strategy to control polymer topology. Science 1999, 283, 2059–2061. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, C.; Luo, S.; Jordan, R.F. Multiple Insertion of a Silyl Vinyl Ether by (α-Diimine)PdMe+ Species. J. Am. Chem. Soc. 2008, 130, 12892–12893. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, C.; Jordan, R.F. Palladium-catalyzed dimerization of vinyl ethers to acetals. J. Am. Chem. Soc. 2010, 132, 10254–10255. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, C.; Luo, S.; Jordan, R.F. Cationic Polymerization and Insertion Chemistry in the Reactions of Vinyl Ethers with (α-Diimine)PdMe+ Species. J. Am. Chem. Soc. 2010, 132, 5273–5284. [Google Scholar] [CrossRef] [PubMed]
  7. Zhu, L.; Fu, Z.S.; Pan, H.J.; Feng, W.; Chen, C.L.; Fan, Z.Q. Synthesis and application of binuclear α-diimine nickel/palladium catalysts with a conjugated backbone. Dalton Trans. 2014, 43, 2900–2906. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, R.K.; Sui, X.L.; Pang, W.M.; Chen, C.L. Ethylene polymerization by xanthene-bridged dinuclear α-diimine NiII complexes. ChemCatChem 2016, 8, 434–440. [Google Scholar] [CrossRef]
  9. Na, Y.N.; Wang, X.; Lian, K.; Zhu, Y.; Li, W.; Luo, Y.; Chen, C.L. Dinuclear α–diimine Ni(II) and Pd(II) Catalyzed Ethylene Polymerization and Copolymerization. ChemCatChem 2017, 9, 1062–1066. [Google Scholar] [CrossRef]
  10. Wang, R.K.; Zhao, M.H.; Chen, C.L. Influence of ligand second coordination sphere effects on the olefin (co)polymerization properties of α-diimine Pd(II) catalysts. Polym. Chem. 2016, 7, 3933–3938. [Google Scholar] [CrossRef]
  11. Guo, L.H.; Dai, S.Y.; Chen, C.L. Investigations of the ligand electronic effects on α-diimine nickel(II) catalyzed ethylene polymerization. Polymers 2016, 8, 37. [Google Scholar] [CrossRef]
  12. Takano, S.; Takeuchi, D.; Osakada, K.; Akamatsu, N.; Shishido, A. Dipalladium Catalyst for Olefin Polymerization: Introduction of Acrylate Units into the Main Chain of Branched Polyethylene. Angew. Chem. Int. Ed. 2014, 53, 9246–9250. [Google Scholar] [CrossRef] [PubMed]
  13. Allen, K.E.; Campos, J.; Daugulis, O.; Brookhart, M. Living Polymerization of Ethylene and Copolymerization of Ethylene/Methyl Acrylate Using “Sandwich” Diimine Palladium Catalysts. ACS Catal. 2015, 5, 456–464. [Google Scholar] [CrossRef]
  14. Zhou, C.; Liu, W.J.; Daugulis, O.; Brookhart, M. Mechanistic Studies of Pd(II)-Catalyzed Copolymerization of Ethylene and Vinylalkoxysilanes: Evidence for a β-Silyl Elimination Chain Transfer Mechanism. J. Am. Chem. Soc. 2016, 138, 16120–16129. [Google Scholar]
  15. Younkin, T.R.; Connor, E.F.; Henderson, J.I.; Friedrich, S.K.; Grubbs, R.H.; Bansleben, D.A. Neutral, Single-Component Nickel(II) Polyolefin Catalysts That Tolerate Heteroatoms. Science 2000, 287, 460–462. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, C.; Friedrich, S.; Younkin, T.R.; Li, R.T.; Grubbs, R.H.; Bansleben, D.A.; Day, M.W. Neutral Nickel(II)-Based Catalysts for Ethylene Polymerization. Organometallics 1998, 17, 3149–3151. [Google Scholar] [CrossRef]
  17. Weberski, M.P.; Chen, C.; Delferro, M.; Zuccaccia, C.; Macchioni, A.; Marks, T.J. Suppression of β-hydride chain transfer in nickel(II)-catalyzed ethylene polymerization via weak fluorocarbon ligand–product interactions. Organometallics 2012, 31, 3773–3789. [Google Scholar] [CrossRef]
  18. Weberski, M.P.; Chen, C.; Delferro, M.; Marks, T.J. Ligand Steric and Fluoroalkyl Substituent Effects on Enchainment Cooperativity and Stability in Bimetallic Nickel(II) Polymerization Catalysts. Chem. Eur. J. 2012, 18, 10715–10732. [Google Scholar] [CrossRef] [PubMed]
  19. Stephenson, C.J.; McInnis, J.P.; Chen, C.; Weberski, M.P.; Motta, A.; Delferro, M.; Marks, T.J. Ni(II) Phenoxyiminato Olefin Polymerization Catalysis: Striking Coordinative Modulation of Hyperbranched Polymer Microstructure and Stability by a Proximate Sulfonyl Group. ACS Catal. 2014, 4, 999–1003. [Google Scholar] [CrossRef]
  20. Ölscher, F.; Göttker-Schnetmann, I.; Monteil, V.; Mecking, S. Role of Radical Species in Salicylaldiminato Ni(II) Mediated Polymer Chain Growth: A Case Study for the Migratory Insertion Polymerization of Ethylene in the Presence of Methyl Methacrylate. J. Am. Chem. Soc. 2015, 137, 14819–14828. [Google Scholar] [CrossRef] [PubMed]
  21. Godin, A.; Göttker-Schnetmann, I.; Mecking, S. Nanocrystal Formation in Aqueous Insertion Polymerization. Macromolecules 2016, 49, 8825–8837. [Google Scholar] [CrossRef]
  22. Nakamura, A.; Anselment, T.M.J.; Claverie, J.; Goodall, B.; Jordan, R.F.; Mecking, S.; Rieger, B.; Sen, A.; van Leeuwen, P.W.N.M.; Nozaki, K. Ortho-phosphinobenzenesulfonate: A superb ligand for palladium-catalyzed coordination–insertion copolymerization of polar vinyl monomers. Acc. Chem. Res. 2013, 46, 1438–1449. [Google Scholar] [CrossRef] [PubMed]
  23. Piche, L.; Daigle, J.C.; Rehse, G.; Claverie, J.P. Structure–Activity Relationship of Palladium Phosphanesulfonates: Toward Highly Active Palladium-Based Polymerization Catalysts. Chem. Eur. J. 2012, 18, 3277–3285. [Google Scholar] [CrossRef] [PubMed]
  24. Ota, Y.; Ito, S.; Kuroda, J.; Okumura, Y.; Nozaki, K. Quantification of the Steric Influence of Alkylphosphine–Sulfonate Ligands on Polymerization, Leading to High-Molecular-Weight Copolymers of Ethylene and Polar Monomers. J. Am. Chem. Soc. 2014, 136, 11898–11901. [Google Scholar] [CrossRef] [PubMed]
  25. Nakano, R.; Nozaki, K. Copolymerization of propylene and polar monomers using Pd/IzQO catalysts. J. Am. Chem. Soc. 2015, 137, 10934–10937. [Google Scholar] [CrossRef] [PubMed]
  26. Jian, Z.B.; Moritz, B.C.; Mecking, S. Suppression of Chain Transfer in Catalytic Acrylate Polymerization via Rapid and Selective Secondary Insertion. J. Am. Chem. Soc. 2015, 137, 2836–2839. [Google Scholar] [CrossRef] [PubMed]
  27. Jian, Z.B.; Mecking, S. Insertion Homo- and Copolymerization of Diallyl Ether. Angew. Chem. Int. Ed. 2015, 54, 15845–15849. [Google Scholar] [CrossRef]
  28. Zhang, Y.L.; Cao, Y.C.; Leng, X.B.; Chen, C.; Huang, Z. Cationic Palladium(II) Complexes of Phosphine–Sulfonamide Ligands: Synthesis, Characterization, and Catalytic Ethylene Oligomerization. Organometallics 2014, 33, 3738–3745. [Google Scholar] [CrossRef]
  29. Sui, X.L.; Dai, S.Y.; Chen, C.L. Ethylene polymerization and copolymerization with polar monomers by cationic phosphine phosphonic amide palladium complexes. ACS Catal. 2015, 5, 5932–5937. [Google Scholar] [CrossRef]
  30. Chen, M.; Zou, W.P.; Cai, Z.G. Norbornene homopolymerization and copolymerization with ethylene by phosphine-sulfonate nickel catalysts. Polym. Chem. 2015, 6, 2669–2676. [Google Scholar] [CrossRef]
  31. Chen, M.; Yang, B.P.; Chen, C.L. Redox-controlled olefin (co)polymerization catalyzed by ferrocene-bridged phosphine-sulfonate palladium complexes. Angew. Chem. Int. Ed. 2015, 54, 15520–15744. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, M.; Yang, B.P.; Chen, C.L. Redox Control in Olefin Polymerization and Copolymerization. Synlett 2016, 27, 1297–1302. [Google Scholar] [CrossRef]
  33. Na, Y.N.; Zhang, D.; Chen, C.L. Modulating the Polyolefin Properties through the Incorporation of Nitrogen-Containing Polar Monomers. Polym. Chem. 2017, 8, 2405–2409. [Google Scholar] [CrossRef]
  34. Jian, Z.; Falivene, L.; Boffa, G.; Ortega Sánchez, S.; Caporaso, L.; Grassi, A.; Mecking, S. Direct Synthesis of Telechelic Polyethylene by Selective Insertion Polymerization. Angew. Chem. Int. Ed. 2016, 55, 14378–14595. [Google Scholar] [CrossRef] [PubMed]
  35. Wada, S.; Jordan, R.F. Olefin Insertion into a Pd–F Bond: Catalyst Reactivation Following β-F Elimination in Ethylene/Vinyl Fluoride Copolymerization. Angew. Chem. Int. Ed. 2017, 56, 1820–1824. [Google Scholar] [CrossRef] [PubMed]
  36. Rhinehart, J.L.; Brown, L.A.; Long, B.K. A robust Ni(II) α-diimine catalyst for high temperature ethylene polymerization. J. Am. Chem. Soc. 2013, 135, 16316–16319. [Google Scholar] [CrossRef] [PubMed]
  37. Dai, S.Y.; Sui, X.L.; Chen, C.L. Highly Robust Palladium(II) α-Diimine Catalysts for Slow-Chain-Walking Polymerization of Ethylene and Copolymerization with Methyl Acrylate. Angew. Chem. Int. Ed. 2015, 54, 9948–9953. [Google Scholar] [CrossRef] [PubMed]
  38. Dai, S.Y.; Chen, C.L. Direct Synthesis of Functionalized High-Molecular-Weight Polyethylene by Copolymerization of Ethylene with Polar Monomers. Angew. Chem. Int. Ed. 2016, 55, 13281–13285. [Google Scholar] [CrossRef] [PubMed]
  39. Dai, S.Y.; Zhou, S.X.; Zhang, W.; Chen, C.L. Systematic Investigations of Ligand Steric Effects on α-Diimine Palladium Catalyzed Olefin Polymerization and Copolymerization. Macromolecules 2016, 49, 8855–8862. [Google Scholar] [CrossRef]
  40. Hu, X.H.; Dai, S.Y.; Chen, C.L. Ethylene polymerization by salicylaldimine nickel(II) complexes containing a dibenzhydryl moiety. Dalton Trans. 2016, 45, 1496–1503. [Google Scholar] [CrossRef] [PubMed]
  41. Guironnet, D.; Roesle, P.; Runzi, T.; Gottker-Schnetmann, I.; Mecking, S. Insertion polymerization of acrylate. J. Am. Chem. Soc. 2009, 131, 422–423. [Google Scholar] [CrossRef] [PubMed]
  42. Perrotin, P.; McCahill, J.S.J.; Wu, G.; Scott, S.L. Linear, high molecular weight polyethylene from a discrete, mononuclear phosphinoarenesulfonate complex of nickel(II). Chem. Commun. 2011, 47, 6948–6950. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, F.; Hu, H.; Xu, Y.; Guo, L.; Zai, S.; Song, K.; Gao, H.; Zhang, L.; Zhu, F.; Wu, Q. Thermostable α-Diimine Nickel(II) Catalyst for Ethylene Polymerization: Effects of the Substituted Backbone Structure on Catalytic Properties and Branching Structure of Polyethylene. Macromolecules 2009, 42, 7789–7796. [Google Scholar] [CrossRef]
  44. Guo, L.; Gao, H.; Guan, Q.; Hu, H.; Deng, J.; Liu, J.; Liu, F.; Wu, Q. Substituent effects of the backbone in α-diimine palladium catalysts on homo-and copolymerization of ethylene with methyl acrylate. Organometallics 2012, 31, 6054–6062. [Google Scholar] [CrossRef]
  45. Zou, W.P.; Chen, C.L. Influence of backbone substituents on the ethylene (co)polymerization properties of α-diimine Pd (II) and Ni (II) catalysts. Organometallics 2016, 35, 1794–1801. [Google Scholar] [CrossRef]
  46. Ota, Y.; Ito, S.; Kobayashi, M.; Kitade, S.; Sakata, K.; Tayano, T.; Nozaki, K. Crystalline Isotactic Polar Polypropylene from the Palladium-Catalyzed Copolymerization of Propylene and Polar Monomers. Angew. Chem. Int. Ed. 2016, 55, 7505–7509. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, Z.X.; Chen, M.; Chen, C.L. Ethylene Polymerization and Copolymerization by Palladium and Nickel Catalysts Containing Naphthalene-Bridged Phosphine–Sulfonate Ligands. Organometallics 2016, 35, 1472–1479. [Google Scholar] [CrossRef]
  48. Nowack, R.J.; Hearley, A.K.; Rieger, B. New Phenylnickel-(2-phosphinobenzenesulfonate) Triphenylphosphine Complexes as Highly Active Ethylene Polymerization Catalysts. Anorg. Allg. Chem. 2005, 631, 2775–2781. [Google Scholar] [CrossRef]
  49. Ito, S.; Munakata, K.; Nakamura, A.; Nozaki, K. Copolymerization of Vinyl Acetate with Ethylene by Palladium/Alkylphosphine–Sulfonate Catalysts. J. Am. Chem. Soc. 2009, 131, 14606–14607. [Google Scholar] [CrossRef] [PubMed]
  50. Long, B.K.; Eagan, J.M.; Mulzer, M.; Coates, G.W. Semi-Crystalline Polar Polyethylene: Ester-Functionalized Linear Polyolefins Enabled by a Functional-Group-Tolerant, Cationic Nickel Catalyst. Z. Angew. Chem. Int. Ed. 2016, 55, 7222–7226. [Google Scholar] [CrossRef]
  51. Chen, M.; Chen, C.L. Rational Design of High-Performance Phosphine Sulfonate Nickel Catalysts for Ethylene Polymerization and Copolymerization with Polar Monomers. ACS Catal. 2017, 7, 1308. [Google Scholar] [CrossRef]
  52. Cooper, C. Organic Chemist’s Desk Reference, 2nd ed.; Chapman and Hall/CRC: Boca Raton, FL, USA, 1995; p. 151. [Google Scholar]
  53. Runzi, T.; Tritschler, U.; Roesle, P.; Gottker-Schnetmann, I.; Moller, H.M.; Caporaso, L.; Poater, A.; Cavallo, L.; Mecking, S. Activation and Deactivation of Neutral Palladium(II) Phosphinesulfonato Polymerization Catalysts. Organometallics 2012, 31, 8388–8408. [Google Scholar] [CrossRef]
Scheme 1. The α-diimine, phenoxyminato and phosphine-sulfonate based olefin polymerization catalysts.
Scheme 1. The α-diimine, phenoxyminato and phosphine-sulfonate based olefin polymerization catalysts.
Polymers 09 00168 sch001
Scheme 2. Synthesis of the phosphine-sulfonate ligands and the palladium and nickel complexes.
Scheme 2. Synthesis of the phosphine-sulfonate ligands and the palladium and nickel complexes.
Polymers 09 00168 sch002
Figure 1. Molecular structures of: (a) Pd2; and (b) Ni1. Hydrogen atoms have been omitted for clarity. Selected bond lengths (Å) and angles (°) for Pd2: Pd1-C23 = 2.067(11), Pd1-P1 = 2.211(2), Pd1-O3 = 2.143(7), Pd1-O4 = 2.131(7), S1-O3 = 1.469(8), S1-O1 = 1.436(8), S1-C22 = 1.781(9), P1-C13 = 1.864(3), C13-C22 = 1.387(4), P1-Pd1-O32 = 91.3(1), P1-Pd1-O4 = 177.3(2), P1-Pd1-C23 = 94.7(3), Pd1-P1-C13 = 115.5(3), S1-O3-Pd1 = 112.9(4); for Ni1: Ni1-C41 = 1.8873(15), Ni1-P1 = 2.2221(4), Ni1-P2 = 2.2100(4), Ni1-O1 = 1.9553(11), S1-O1 = 1.4352(13), P2-C23 = 1.8309(15), P2-C17 = 1.8118(17), P1-Ni1-O1 = 93.63(3), P1-Ni1-C41 = 88.18(5), P2-Ni1-O1 = 89.47(3), P2-Ni1-C41 = 89.62(5), O1-Ni1-C41 = 174.92(6).
Figure 1. Molecular structures of: (a) Pd2; and (b) Ni1. Hydrogen atoms have been omitted for clarity. Selected bond lengths (Å) and angles (°) for Pd2: Pd1-C23 = 2.067(11), Pd1-P1 = 2.211(2), Pd1-O3 = 2.143(7), Pd1-O4 = 2.131(7), S1-O3 = 1.469(8), S1-O1 = 1.436(8), S1-C22 = 1.781(9), P1-C13 = 1.864(3), C13-C22 = 1.387(4), P1-Pd1-O32 = 91.3(1), P1-Pd1-O4 = 177.3(2), P1-Pd1-C23 = 94.7(3), Pd1-P1-C13 = 115.5(3), S1-O3-Pd1 = 112.9(4); for Ni1: Ni1-C41 = 1.8873(15), Ni1-P1 = 2.2221(4), Ni1-P2 = 2.2100(4), Ni1-O1 = 1.9553(11), S1-O1 = 1.4352(13), P2-C23 = 1.8309(15), P2-C17 = 1.8118(17), P1-Ni1-O1 = 93.63(3), P1-Ni1-C41 = 88.18(5), P2-Ni1-O1 = 89.47(3), P2-Ni1-C41 = 89.62(5), O1-Ni1-C41 = 174.92(6).
Polymers 09 00168 g001
Figure 2. Polyethylene yield versus polymerization time at 80 °C for catalysts Pd2, Pd2′ and Pd2″.
Figure 2. Polyethylene yield versus polymerization time at 80 °C for catalysts Pd2, Pd2′ and Pd2″.
Polymers 09 00168 g002
Scheme 3. Comparison of the three ligand frameworks.
Scheme 3. Comparison of the three ligand frameworks.
Polymers 09 00168 sch003
Table 1. Ethylene polymerization catalyzed by Pd(II) and Ni(II) complexes a.
Table 1. Ethylene polymerization catalyzed by Pd(II) and Ni(II) complexes a.
EntryCatalyst[cat] (μmol)T (°C)Yield (g)Activity bMw cPolydispersity cTm (°C) d
1Pd1108022.021001.15109
2Pd210802.42.451002.36117
3Pd3108055.027,8002.56129
4Pd12800.52.532002.06114
5Pd22800.42.057002.18124
6Pd32801.26.024,8001.84131
7Ni110801.91.936001.58128
8Ni210802.22.291002.18129
9Ni310805.75.717,5002.04133
10Ni12250.71.737,2003.64134
11Ni22250.51.358,6001.99135
12Ni32251.43.5142,3001.51136
13Pd2′10802.12.142002.28123
14Pd2″10801.41.418001.44115
15Ni2′10801.51.541001.46124
16Ni2′2250.42.027,8002.57133
17Ni1″1080trace----
18Ni1″225trace----
a Polymerization conditions: toluene = 22 mL, CH2Cl2 = 3 mL, ethylene = 9 atm, 80 °C, time = 1 h. b Activity is in unit of 105 g·mol−1·h−1. c Determined by Gel Permeation Chromatograph (GPC) in trichlorobenzene at 150 °C (see Supplementary Materials, Figures S36–S49). d Melting temperature was determined by differential scanning calorimetry (DSC) (see Supplementary Materials, Figures S26–S35).
Table 2. Ethylene copolymerization catalyzed by Pd(II) and Ni(II) complexes a.
Table 2. Ethylene copolymerization catalyzed by Pd(II) and Ni(II) complexes a.
EntryCatalystP (bar)T (oC)Comonomer[M] mol/LYield (mg)Activity bX c (%)Mw dPolydispersity d
1Pd1980 Polymers 09 00168 i0011.250025310001.19
2Pd2980 Polymers 09 00168 i0021.234017544001.92
3Pd3980 Polymers 09 00168 i0031.2950471555001.55
4Pd1980 Polymers 09 00168 i0042.520010816001.43
5Pd2980 Polymers 09 00168 i0052.5300151221001.35
6Pd3980 Polymers 09 00168 i0062.5510252736001.58
7Ni3925 Polymers 09 00168 i0071.01507.51.561,5002.64
8Ni3925 Polymers 09 00168 i0081.0500250.5124,6002.25
9Pd2″925 Polymers 09 00168 i0091.21105.52.519501.95
10Pd2″925 Polymers 09 00168 i0102.5804611001.63
11Ni1″925 Polymers 09 00168 i0111.000---
a Polymerization conditions: total volume toluene + polar monomer = 25 mL, catalyst = 20 μmol. b Activity in unit of 103 g·mol−1·h−1. c Amount of polar monomer incorporated (mol %), determined by 1H NMR spectroscopy (see Supplementary Materials, Figures S21–S25). d Determined by GPC in trichlorobenzene at 150 °C (see Supplementary Materials, Figures S49–S56).

Share and Cite

MDPI and ACS Style

Wu, Z.; Hong, C.; Du, H.; Pang, W.; Chen, C. Influence of Ligand Backbone Structure and Connectivity on the Properties of Phosphine-Sulfonate Pd(II)/Ni(II) Catalysts. Polymers 2017, 9, 168. https://doi.org/10.3390/polym9050168

AMA Style

Wu Z, Hong C, Du H, Pang W, Chen C. Influence of Ligand Backbone Structure and Connectivity on the Properties of Phosphine-Sulfonate Pd(II)/Ni(II) Catalysts. Polymers. 2017; 9(5):168. https://doi.org/10.3390/polym9050168

Chicago/Turabian Style

Wu, Zixia, Changwen Hong, Hongxu Du, Wenmin Pang, and Changle Chen. 2017. "Influence of Ligand Backbone Structure and Connectivity on the Properties of Phosphine-Sulfonate Pd(II)/Ni(II) Catalysts" Polymers 9, no. 5: 168. https://doi.org/10.3390/polym9050168

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