Next Article in Journal
Nanopaper Properties and Adhesive Performance of Microfibrillated Cellulose from Different (Ligno-)Cellulosic Raw Materials
Next Article in Special Issue
Facile, Efficient Copolymerization of Ethylene with Norbornene-Containing Dienes Promoted by Single Site Non-Metallocene Oxovanadium(V) Catalytic System
Previous Article in Journal
Effect of Bamboo Flour Grafted Lactide on the Interfacial Compatibility of Polylactic Acid/Bamboo Flour Composites
Previous Article in Special Issue
Homo- and Copolymerizations of Ethylene and Norbornene Using Bis(β-ketoamino) Titanium Catalysts Containing Pyrazolone Rings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Use of Vanadium Complexes Bearing Naphthalene-Bridged Nitrogen-Sulfonate Ligands as Catalysts for Copolymerization of Ethylene and Propylene

1
College of Chemistry, Jilin University, Changchun 130012, China
2
Key Laboratory of Synthetic Rubber, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China
3
Department of Wound Repair Center, Burn & Plastic Surgery, First Affiliated Hospital of Chinese PLA General Hospital, Beijing 100037, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this study.
Polymers 2017, 9(8), 325; https://doi.org/10.3390/polym9080325
Submission received: 25 June 2017 / Revised: 19 July 2017 / Accepted: 24 July 2017 / Published: 31 July 2017
(This article belongs to the Special Issue Olefin Polymerization and Polyolefin)

Abstract

:
Vanadium complexes bearing naphthalene-bridged nitrogen-sulfonate ligand ([ê2(N,O)-8-(PhN)-1-naphthalenesulfonato]VOCl (1a) and [ê2(N,O)-8-(PhN)-1-naphthalenesulfonato]VCl2 (1b)) were synthesized. Activated by ethylaluminium sesquichloride (EASC) and in the presence of ethyl trichloroacetate (ETCA) as reactivator, complexes 1a and 1b showed activities of up to 39.1 kg polymer (mol V)−1 h−1, affording the copolymers with high molecular weights (Mw up to 28 × 104) and narrow molecular weight distributions (Mw/Mn ~ 3.0) as well as high propylene incorporation of up to 49.4%. Compared to the traditional VOCl3 system, these complexes exhibited higher propylene incorporation ability and higher catalytic activities especially at high polymerization temperatures of 50 °C and above. Determined by DSC and 13C NMR, the copolymers obtained with 1a and 1b had more random structures than that with the VOCl3 system.

Graphical Abstract

1. Introduction

Design and synthesis of transition metal-based catalysts for olefin polymerization is a long-standing research subject since the pioneering work by Ziegler and Natta, and some of them have been successfully utilized in industrial applications [1,2,3,4,5,6,7]. Vanadium-based catalysts, though less active than those of other systems, exhibit interesting features in olefin (co)polymerization process [8,9,10,11,12,13]. Such catalysts can produce high molecular weight polymers with narrow molecular weight distributions, ethylene/á-olefin copolymers having high á-olefin incorporation, and syndiotactic polypropylene. Particularly, vanadium-based catalysts are widely used worldwide for the industrial production of ethylene/propylene copolymers and ethylene/propylene/diene terpolymers, a rapid growing class of elastomers. However, despite the unique characteristics and practical application of the classical Ziegler-type vanadium catalysts, they suffered from the decay of activity during the polymerization process as a result of reduction of active species to less active or even inactive low-valent vanadium species, which is more prevalent at elevated temperatures. Therefore, the development of vanadium catalysts with high activity and stability has been recognized as an attractive target.
Two strategies are known to improve the catalytic performance of vanadium catalysts. The use of reoxidants such as ethyl trichloroacetate and chlorinated hydrocarbons (so-called “rejuvenators” or “promoters” in the literature) is well-known for maintaining the higher (active) oxidation state of vanadium systems by the reactivation of vanadium (II) center to vanadium (III) species [14,15]. The incorporation of appropriate ancillary ligands modification vanadium-based systems has proved another effective way for stabilizing the catalytically active species [8,9,10,11,12,13,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. For instance, Nomura and coworkers reported that imido-based vanadium complexes containing phenoxyimine or anilidomethylpyridine coligands are efficient catalyst precursors for olefin (co)polymerization [19,20,21]. Redshaw et al. prepared a variety of vanadium complexes bearing chelating aryloxides, which exhibit high activity in ethylene polymerization even at elevated temperatures [10,23,24]. Li et al. reported various [N, O], [N, N], and [O, P=O] ligands such as β-enaminoketonato, iminopyrrolide, and phenoxy-phosphine oxide supported vanadium complexes with high activity and comonomer incorporation ability in ethylene (co)polymerization [9,25,26,27]. Abernethy reported a highly stable N-heterocyclic carbene vanadium (V) trichloride oxide complex even in air, but its catalysis properties was not investigated [28]. Nomura synthesized a series of (imido) vanadium (V) complexes bearing anionic N-heterocyclic carbenes, which exhibit remarkable catalytic activity for ethylene polymerization in the presence of Al(i-Bu)3 [29]. Recently, chelating sulfonated ligands have attracted considerable interest in the design of highly active catalysts because of their remarkable electronic asymmetry as well as geometric flexibility [36,37,38]. Given the impressive activities associated with chelating sulfonated ligand-based Ru, Pd, and Ni catalytic chemistry including olefin metathesis, olefin polymerization, and copolymerization with polar vinyl monomers [39,40,41], such hybrid ligands would be attractive candidates for preparation of vanadium catalysts with high performance. As mentioned above, many efforts have been made to develop novel vanadium catalysts, and a variety of vanadium complexes with high activities and incorporation ability have been successfully prepared. VOCl3 has been widely used for the industrial production of ethylene/propylene copolymers and ethylene/propylene/diene terpolymers. Therefore, in the present study, two novel vanadium complexes bearing naphthalene-bridged nitrogen-sulfonate ligand, [ê2(N,O)-8-(PhN)-1-naphthalenesulfonato]VOCl and [ê2(N,O)-8-(PhN)-1-naphthalenesulfonato]VCl2, were synthesized, and due to the ease of preparation, these complexes might be promising candidates for practical applications. Their catalytic behaviors toward ethylene/propylene copolymerization were investigated by comparing with the results of the traditional VOCl3-based catalyst as a control system. Furthermore, the effects of cocatalyst quantity, polymerization temperature, and time on the polymerization were examined in detail.

2. Experimental Section

2.1. General Considerations and Materials

All the manipulations were carried out under nitrogen atmosphere using standard Schlenk techniques. Hexane and tetrahydrofuran were refluxed over sodium/diphenyl ketyl under nitrogen and then distilled prior to use. Vanadium (V) trichloride oxide (VOCl3) and vanadium tetrachloride (VCl4) were purchased from Nanjing XingHong Chemical Co (Nanjing, China) and used as received. 8-(Phenylamino)naphthalene-1-sulfonic acid and ethyl trichloroacetate (ETCA) were purchased from Aladdin Co (Shanghai, China). Ethylaluminium sesquichloride (EASC) was purchased from Sigma-Aldrich Co (Saint Louis, MO, USA) and diluted to 1.0 mol/L solution by hexane. Polymerization grade ethylene and propylene were further purified by passing through columns of 5 Å molecular sieves and MnO. Other chemicals were commercially available and used without further purification.
The NMR spectra of polymers were recorded on a Bruker 300 MHz spectrometer at 135 °C with 1,2-dichlorobenzene-d4 as a solvent. Elemental analysis was conducted with Carlo Erba 1106 and ST02 apparatus. Differential scanning calorimetry (DSC) measurements were performed with a TA instrument DSC 2920 at a heating rate of 10 °C/min. The molecular weight and molecular weight distribution of the polymers were determined by gel permeation chromatography (GPC) at 140 °C on a PL-GPC 220 high temperature chromatograph equipped with three PL gel 10 mm Mixed-B LS type columns. 1,2,4-Trichlorobenzene was employed as eluant at a flow rate of 1.0 mL/min. The calibration was made by polystyrene standard (Easi-Cal PS-1 PL Ltd.).

2.2. Synthesis of Complexes

2.2.1. Synthesis of Complex 1a

A 100 mL three-necked flask was equipped with a three-way stopcock and a magnetic stirring bar. After the flask was flushed with nitrogen, tetrahydrofuran (40 mL) and 8-(phenylamino) naphthalene-1-sulfonic acid (1.5 g, 5.0 mmol) was added, and then vanadium (V) trichloride oxide (0.88 g, 5.1 mmol) was added at −20 °C. The reaction mixture was warmed to room temperature and stirred for 48 h. The solvent was removed in vacuo, the resulting solid was repeatedly washed with hexane and dried under vacuum to give a brown solid; yield (1.2 g, 3.0 mmol), 60%. IR (KBr, cm−1): 1618, 1602, 1483, 1251, 1141, 1037, 999, 762, 747, 681. 1H NMR (CDCl3) δ (ppm): 8.28 (s, 1H, Ar), 7.82 (s, 1H, Ar), 7.61−7.12 (m, 8H, Ar), 6.78 (s, 1H, Ar). Anal. Calcd for C16H11ClNO4SV: C, 48.08; H, 2.77; N, 3.50; S, 8.02. Found: C, 50.80; H, 2.56; N, 2.99; S, 8.50.

2.2.2. Synthesis of Complex 1b

A 100 mL three-necked flask was equipped with a three-way stopcock and a magnetic stirring bar. After the flask was flushed with nitrogen, tetrahydrofuran (40 mL) and 8-(phenylamino) naphthalene-1-sulfonic acid (1.5 g, 5.0 mmol) was added, and then vanadium (IV) tetrachloride (0.98 g, 5.1 mmol) was added at −20 °C. The reaction mixture was warmed to room temperature and stirred for 48 h. The solvent was removed in vacuo, the resulting solid was repeatedly washed with hexane and dried under vacuum to give a brown solid; yield 58%, IR (KBr, cm−1): 1618, 1604, 1479, 1247, 1199, 1137, 1033, 1003, 740, 680. Anal. Calcd for C16H11Cl2NO3SV: C, 45.84; H, 2.65; N, 3.34; S, 7.65. Found: C, 46.50; H, 2.01; N, 2.96; S, 8.11.

2.3. General Procedure for Copolymerization of Ethylene and Propylene

Copolymerization was carried out in a glass reactor with vigorous stirring. A typical polymerization procedure is described as following. After the reactor was repeatedly evacuated and flushed with nitrogen, hexane (100 mL) was added, and premixed ethylene/propylene gas (1:2 mole ratio) was then introduced until saturation. Copolymerization was initiated by the sequential addition of certain amounts of EASC, vanadium complex, and ETCA and carried out at 1 atm of mixture gas pressure and a desired temperature. The polymerization was quenched by addition of excessive methanol containing a small amount of hydrochloric acid, and the precipitated polymer was filtered, repeatedly washed with methanol, and dried under vacuum at 40 °C to constant weight.

3. Results and Discussion

3.1. Synthesis of Vanadium Complexes

The synthetic route for the vanadium complexes is shown in Scheme 1. The reaction of VCl4 or VOCl3 with 1.0 equivalent of 8-(phenylamino)naphthalene-1-sulfonic acid in THF afforded complexes 1a and 1b in moderate yields (60% and 58%, respectively). While the attempt to obtain the crystals of these two complexes was failed, these complexes were identified by 1H NMR (Figure S1) and element analysis. The signal at 11.2 ppm assignable to NH group of the ligand disappeared in the H1 NMR spectrum of complex, and other signals meanwhile remained intact. This result combined with the elemental analysis data indicates the formation of complex. These two complexes are relatively stable in air and soluble in common organic solvents such as CHCl3, CH2Cl2, and toluene.

3.2. Ethylene/Propylene Copolymerization

It is known that vanadium compounds could be reduced by alkylaluminum to different oxidation state, V(IV), V(III), and V(II), depending on the Al/V ratio [8], and in turn, the dosage of cocatalyst remarkably influenced the activity of vanadium catalysts. In order to investigate the effect of Al/V mole ratio on the catalytic behavior, the ethylene/propylene copolymerization was conducted with complexes 1a and 1b in the presence of the cocatalyst EASC and the reactivator ETCA along with commercially used VOCl3 for comparison. As shown in Figure 1, in all cases, the catalytic activity increased as the Al/V mole ratio increased from 10 to 40, and the activity of VOCl3-based system was higher than that of 1a-system whilst somewhat lower than the 1b-system. However, a further increased Al/V ratio to 50 resulted in a decrease of activity of VOCl3 due to the reduction of the active species by the cocatalyst, whereas the 1a and 1b catalyst system exhibited an increase in activities. These results indicate that the naphthalene-bridged nitrogen-sulfonate ligand protects 1a and 1b catalysts against catalyst deactivation typically through reduction by an aluminum cocatalyst to the inactive divalent vanadium species. Since the highest activity of traditional VOCl3-based system was obtained at an Al/V mole ratio of 40, it was used in the further experiments to compare the catalytic performance between the newly synthesized vanadium complexes and VOCl3.
In the cases of vanadium-catalyzed olefin polymerization, it is well established that the activity can be enhanced in the presence of reactivator ETCA, which reactivates the low-valent, less active or inactive V(II) species to active vanadium (III) species [15,42]. As shown in Table 1, in all cases, the activities of complexes 1a, 1b, and VOCl3 in the presence of ETCA (38.4~78.4 kg polymer (mol V)−1 h−1 bar−1) were much higher than those of ETCA-free systems (18.6~41.2 kg polymer (mol V)−1 h−1 bar−1). Similar results were observed in other vanadium catalyst systems [15,43]. Especially, in the absence of ETCA, the activities of complexes 1a and 1b were more or less than that of VOCl3, further suggesting that the nitrogen-sulfonate ligand protects active species generated by 1a and 1b from deactivation. The highest activity was observed at ETCA/V ratio of 10 in all cases, and further increased the level of ETCA resulted in a slightly decreased activity. Among three systems, 1b gave the highest catalytic activity of 78.4 kg polymer (mol V)−1 h−1 bar−1, which is higher than those of 1a and VOCl3 (52.8 and 64.0 kg polymer (mol V)−1 h−1 bar−1, respectively). Although the detailed reason for this observation is not clear yet, it is speculated that the role played by ETCA to regenerate active vanadium species is more pronounced in the case of 1b, in which the presence of ETCA led to ca. a 4-fold increase in activity, much higher than that in the cases of 1a and VOCl3. On the other hand, the molecular weight of the resulting copolymers decreased with the increase in the ETCA/V ratio. This result might be due to the higher concentration of active species generated in the presence of the reactivator. It is worth noting that the molecular weights of the resulting copolymers obtained with complexes 1a and 1b were higher than that with commercially used VOCl3. Whereas VOCl3 system produced copolymers with rather broad molecular weight distributions (Mw/Mn = 5.46–7.69), the unimodal GPC curves and the narrow molecular weight distributions suggested that the formation of single-site active species in 1a and 1b systems. These facts indicate that EASC does not abstract the naphthalene-bridged nitrogen-sulfonate ligand from vanadium center to the aluminum cocatalyst, as that proposed in the cases of tridentate (OSO) ligand-supported vanadium complexes [44]. Meanwhile, the presence of ETCA also influenced the propylene incorporation in the resulting copolymers. The increase in ETCA/V ratio led to an enhancement in propylene incorporation. For instance, the incorporation of propylene in the copolymers obtained with 1b was 49.4% at ETCA/V ratio of 20, which is much higher than that of in the absence of ETCA (33.1%). Noticeably, under identical conditions the incorporation of propylene in the cases of 1a and 1b was higher than that of VOCl3.
The influence of polymerization temperature on the activity of these systems was further investigated, and the results are listed in Figure 2. The catalytic activities of complexes 1a and 1b as well as VOCl3 decreased with an increase in temperature as commonly observed in other vanadium catalysts, and the reason might be that the increase of polymerization temperature resulted in the accelerated reduction of active high valent vanadium species. Complexes 1a and 1b exhibited higher activities at high temperatures of 50 °C and above compared with VOCl3, suggesting the potential high thermal stability of the two complexes, although they showed more or less lower catalytic activities than that of VOCl3 at 0 °C. The fact of maintaining high activity at high temperatures is important from the viewpoint of practical application.
Catalyst lifetime studies were carried out with complexes 1a and 1b, as well as VOCl3-based systems at 30 °C, and the results are shown in Figure 3. The three catalysts displayed similar trends in catalytic performance with respect to the activity against polymerization time. In all cases, the catalytic activity increased at the first stage and was found to peak at ca. 30 min, 53, 78, and 64 kg polymer (mol V)−1 h−1 for 1a, 1b, and VOCl3, respectively, whereas further prolonging polymerization time resulted in a greater or lesser decrease in activity.

3.3. Polymer Characterization

The thermal behaviors of the copolymer samples were similar. As shown in Figure 4, the DSC thermograms of the copolymers exhibit only a glass transition temperature. The Tgs of the copolymers obtained with 1a and 1b are −56.9 °C and −56.5 °C, respectively, which are slightly lower than that with VOCl3 (−55.9 °C). The copolymers obtained with 1a and 1b did not display melting points, whereas the ethylene-propylene copolymer obtained with VOCl3 had weak melting endotherms at 100 to 120 °C, although the X-ray spectrum indicates that the polymer is amorphous (Figure S2). These results might be due to the more randomly distributed monomer units and/or the increasing propylene incorporation amount in the copolymers than that obtained with VOCl3 system.
Since the sequence distribution is one of the important factors determining the mechanical properties of ethylene/propylene copolymer, the microstructures of ethylene/propylene copolymers were further characterized by 13C NMR. Figure 5 shows the typical 13C NMR spectrum of the resulting copolymer, and the peaks are assigned according to the literature [45,46,47,48]. The composition of the copolymers was estimated from the 13C NMR spectra according to the calculation method proposed by Randall [49], and the results are shown in Figure 6. The copolymer obtained with VOCl3 possessed higher content of EEE triads than those with 1a and 1b, while the content of EPE triad is lower, suggesting that in the cases of 1a and 1b (Figure S3) propylene inserted more randomly along the polymer chain. On the other hand, compared to VOCl3 (Figure S4), 1a and 1b afforded polymers with more or less higher PPP triad contents, which might be due to the higher propylene incorporation.

4. Conclusions

In summary, we have synthesized two novel naphthalene-bridged nitrogen-sulfonate ligand-supported vanadium complexes. The complexes exhibited high catalytic activities in ethylene/propylene copolymerization in the presence of EASC as a cocatalyst and ETCA as a reactivator, affording with high molecular weight copolymers. The naphthalene-bridged nitrogen-sulfonate ligand protects catalysts against catalyst deactivation typically through reduction by an aluminum cocatalyst to the inactive divalent vanadium species. Compared to the traditional VOCl3 system, the newly synthesized vanadium complex-based catalysts showed higher activities at high polymerization temperatures. The unimodal GPC curves and narrow molecular distributions of the resulting polymers suggested the presence of single active species in the polymerization systems. Meanwhile, the complexes displayed higher propylene incorporation ability, and the resulting copolymers possessed more random structures. Taking the advantages of facile preparation, high activity especially at elevated temperatures and propylene incorporation ability as well as high molecular weight and uniform composition of the resulting polymers, these vanadium complexes are promising as practically useful catalysts for ethylene/propylene copolymerization.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4360/9/8/325/s1, Figure S1: 1H NMR spectra of the ligand and complex 1a. Figure S2: XRD spectrum of ethylene/propylene copolymer obtained with 1a/EASC catalyst (sample from run 3 in Table 1). Figure S3: 13C NMR spectrum of ethylene/propylene copolymer obtained with 1b/EASC catalyst (sample from run 6 in Table 1). Figure S4: 13C NMR spectrum of ethylene/propylene copolymer obtained with VOCl3/EASC catalyst (sample from run 10 in Table 1).

Acknowledgments

This work was supported by the 973 Program, Grant No. 2015CB654700 (2015CB654702), the Capital Characteristic Clinical Application Research (No. Z141107002514167), Health Bureau of Logistical Support Department of the Central Military Commission (No. BWS14J049), and the National Natural Science Foundation of China (No. 51073067, No. u1462124).

Author Contributions

X.H. and X.Z. conceived and designed the experiments; C.Z., L.L. and H.Z. performed the experiments; D.H. contributed analysis tools; Y.H. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ziegler, K.; Holzkamp, E.; Breil, H.; Martin, H. Das mülheimer normaldruck-polyäthylen-verfahren. Angew. Chem. 1955, 67, 541–547. [Google Scholar] [CrossRef]
  2. Natta, G. Stereospezifische katalysen und isotaktische polymere. Angew. Chem. 1956, 68, 393–403. [Google Scholar] [CrossRef]
  3. Coates, G.W. Precise control of polyolefin stereochemistry using single-site metal catalysts. Chem. Rev. 2000, 100, 1223–1252. [Google Scholar] [CrossRef] [PubMed]
  4. Ittel, S.D.; Johnson, L.K.; Brookhart, M. Late-metal catalysts for ethylene homo- and copolymerization. Chem. Rev. 2000, 100, 1169–1204. [Google Scholar] [CrossRef] [PubMed]
  5. Gibson, V.C.; Spitzmesser, S.K. Advances in non-metallocene olefin polymerization catalysis. Chem. Rev. 2003, 103, 283–316. [Google Scholar] [CrossRef] [PubMed]
  6. Makio, H.; Terao, H.; Iwashita, A.; Fujita, T. FI catalysts for olefin polymerization—A comprehensive treatment. Chem. Rev. 2011, 111, 2363–2449. [Google Scholar] [CrossRef] [PubMed]
  7. Baier, M.C.; Zuideveld, M.A.; Mecking, S. Post-metallocenes in the industrial production of polyolefins. Angew. Chem. Int. Ed. 2014, 53, 9722–9744. [Google Scholar] [CrossRef] [PubMed]
  8. Nomura, K.; Zhang, S. Design of vanadium complex catalysts for precise olefin polymerization. Chem. Rev. 2011, 111, 2342–2362. [Google Scholar] [CrossRef] [PubMed]
  9. Wu, J.Q.; Li, Y.S. Well-defined vanadium complexes as the catalysts for olefin polymerization. Coord. Chem. Rev. 2011, 255, 2303–2314. [Google Scholar] [CrossRef]
  10. Redshaw, C. Vanadium procatalysts bearing chelating aryloxides: Structure–activity trends in ethylene polymerisation. Dalton Trans. 2010, 39, 5595–5604. [Google Scholar] [CrossRef] [PubMed]
  11. Matsugi, T.; Fujita, T. High-performance olefin polymerization catalysts discovered on the basis of a new catalyst design concept. Chem. Soc. Rev. 2008, 37, 1264–1277. [Google Scholar] [CrossRef] [PubMed]
  12. Gambarotta, S. Vanadium-based Ziegler-Natta: Challenges, promises, problems. Coord. Chem. Rev. 2003, 237, 229–243. [Google Scholar] [CrossRef]
  13. Hagen, H.; Boersma, J.; van Koten, G. Homogeneous vanadium-based catalysts for the Ziegler-Natta polymerization of α-olefins. Chem. Soc. Rev. 2002, 31, 357–364. [Google Scholar] [CrossRef] [PubMed]
  14. Gumboldt, A.; Helberg, J.; Schleitzer, G. Über die reaktivierung der bei der äuthylen/propylen-copolymerisation verwendeten vanadium-katalysatoren. Makromol. Chem. 1967, 101, 229–245. [Google Scholar] [CrossRef]
  15. Christman, D.L. Preparation of polyethylene in solution. J. Polym. Sci. Part A 1972, 10, 471–487. [Google Scholar] [CrossRef]
  16. Zhang, S.; Zhang, W.C.; Shang, D.D.; Zhang, Z.Q.; Wu, Y.X. Ethylene/propylene copolymerization catalyzed by vanadium complexes containing N-heterocyclic carbenes. Dalton Trans. 2015, 44, 15264–15270. [Google Scholar] [CrossRef] [PubMed]
  17. Nomura, K.; Igarashi, A.; Katao, S.; Zhang, W.J.; Sun, W.H. Synthesis and structural analysis of (imido)vanadium(V) complexes containing chelate (anilido)methyl-imine ligands: Ligand effect in ethylene dimerization. Inorg. Chem. 2013, 52, 2607–2614. [Google Scholar] [CrossRef] [PubMed]
  18. Bialek, M.; Czaja, K.; Pietruszka, A. Ethylene/1-olefin copolymerization behaviour of vanadium and titanium complexes bearing salen-type ligand. Polym. Bull. 2013, 70, 1499–1517. [Google Scholar] [CrossRef]
  19. Igarashi, A.; Zhang, S.; Nomura, K. Ethylene dimerization/polymerization catalyzed by (adamantylimido)vanadium(V) complexes containing (2-anilidomethyl)pyridine ligands: Factors affecting the ethylene reactivity. Organometallics 2012, 31, 3575–3581. [Google Scholar] [CrossRef]
  20. Zhang, S.; Katao, S.; Sun, W.H.; Nomura, K. Synthesis of (arylimido)vanadium(V) complexes containing (2-anilidomethyl)pyridine ligands and their Use as the catalyst precursors for olefin polymerization. Organometallics 2012, 28, 5925–5933. [Google Scholar] [CrossRef]
  21. Onishi, Y.; Katao, S.; Fujiki, M.; Nomura, K. Synthesis and structural analysis of (arylimido)vanadium(V) complexes containing phenoxyimine ligands: New, efficient catalyst precursors for ethylene polymerization. Organometallics 2008, 27, 2590–2596. [Google Scholar] [CrossRef]
  22. Smit, T.M.; Tomov, A.K.; Britovsek, G.J.P.; Gibson, V.C.; White, A.J.P.; Williams, D.J. The effect of imine-carbon substituents in bis(imino)pyridine-based ethylene polymerization catalysts across the transition series. Catal. Sci. Technol. 2012, 2, 643–655. [Google Scholar] [CrossRef]
  23. Redshaw, C.; Clowes, L.; Hughes, D.L.; Elsegood, M.R.J.; Yamato, T. Ethylene polymerization catalysis by vanadium-based systems bearing sulfur-bridged calixarenes. Organometallics 2011, 30, 5620–5624. [Google Scholar] [CrossRef]
  24. Clowes, L.; Redshaw, C.; Hughes, D.L. Vanadium-based pro-catalysts bearing depleted 1,3-calix[4]arenes for ethylene or ε-caprolactone polymerization. Inorg. Chem. 2011, 50, 7838–7845. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, S.W.; Lu, L.P.; Long, Y.Y.; Li, Y.S. Synthesis, structural characterization, and ethylene polymerization behavior of (arylimido)vanadium(V) complexes bearing tridentate Schiff base ligands. J. Polym. Sci. Part A 2014, 52, 2633–2642. [Google Scholar]
  26. Lu, L.P.; Wang, J.B.; Liu, J.Y.; Li, Y.S. Ethylene polymerization and ethylene/hexene copolymerization by vanadium(III) complexes bearing bidentate phenoxy-phosphine oxide ligands. J. Polym. Sci. Part A 2013, 51, 5298–5306. [Google Scholar]
  27. Zhang, S.W.; Zhang, G.B.; Lu, L.P.; Li, Y.S. Novel vanadium(III) complexes with tridentate phenoxy-phosphine [O,P(O),O] ligands: Synthesis, characterization, and catalytic behavior of ethylene polymerization and copolymerization with 10-undecen-1-ol. J. Polym. Sci. Part A 2013, 51, 844–854. [Google Scholar] [CrossRef]
  28. Abernethy, C.D.; Codd, G.M.; Spicer, M.D.; Taylor, M.K. A highly stable N-heterocyclic carbene complex of trichloro-oxo-vanadium(V) displaying novel Cl−Ccarbene bonding interactions. J. Am. Chem. Soc. 2003, 125, 1128–1129. [Google Scholar] [CrossRef] [PubMed]
  29. Igarashi, A.; Kolychev, E.L.; Tamm, M.; Nomura, K. Synthesis of (imido)vanadium(V) dichloride Complexes Containing Anionic N-Heterocyclic Carbenes That Contain a Weakly Coordinating Borate moiety: new MAO-free ethylene polymerization catalysts. Organometallics 2016, 35, 1778–1784. [Google Scholar] [CrossRef]
  30. Tsuchiya, Y.; Endo, K. Vanadium alkoxide catalyzed polymerization of vinyl chloride. J. Polym. Sci. Part A 2011, 49, 1006–1012. [Google Scholar] [CrossRef]
  31. Wu, J.Q.; Pan, L.; Li, Y.G.; Liu, S.R.; Li, Y.S. Synthesis, structural characterization, and olefin polymerization behavior of vanadium(III) complexes bearing tridentate schiff base ligands. Organometallics 2009, 28, 1817–1825. [Google Scholar] [CrossRef]
  32. Tomov, A.K.; Gibson, V.C.; Zaher, D.; Elsegood, M.; Dale, S.H. Bis(benzimidazole)amine vanadium catalysts forolefin polymerization and co-polymerisation: thermally robust, single-site catalysts activated by simple alkylaluminium reagents. Chem. Commun. 2004, 17, 1956–1957. [Google Scholar] [CrossRef] [PubMed]
  33. Nomura, K.; Mitsudome, T.; Lgarashi, A. Synthesis of (adamantylimido)vanadium(V) dimethyl complex containing (2-anilidomethyl)pyridine ligand and selected reactions: exploring the oxidation state of the catalytically active species in ethylene dimerization. Organometallics 2017, 36, 530–542. [Google Scholar] [CrossRef]
  34. Nomura, K.; Hou, X. Synthesis of vanadium-alkylidene complexes and their use as catalysts for ring opening metathesis polymerization. Dalton Trans. 2017, 46, 12–24. [Google Scholar] [CrossRef] [PubMed]
  35. Diteepeng, N.; Tang, X.; Hou, X. Ethylene polymerisation and ethylene/norbornene copolymerisation by using aryloxo-modified vanadium(V) complexes containing 2,6-difluoro-, dichloro-phenylimido complexes. Dalton Trans. 2015, 44, 12273–12281. [Google Scholar] [CrossRef] [PubMed]
  36. Nakamura, A.; Ito, S.; Nozaki, K. Coordination−insertion copolymerization of fundamental polar monomers. Chem. Rev. 2009, 109, 5215–5244. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, E.Y.X. Coordination polymerization of polar vinyl monomers by single-site metal catalysts. Chem. Rev. 2009, 109, 5157–5214. [Google Scholar] [CrossRef] [PubMed]
  38. 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]
  39. Carrow, B.P.; Nozaki, K. Transition-metal-catalyzed functional polyolefin synthesis: Effecting control through chelating ancillary ligand design and mechanistic insights. Macromolecules 2014, 47, 2541–2555. [Google Scholar] [CrossRef]
  40. Wu, Z.; Chen, M.; Chen, C. Ethylene polymerization and copolymerization by palladium and nickel catalysts containing naphthalene-bridged phosphine–sulfonate ligands. Organometallics 2016, 35, 1472–1479. [Google Scholar] [CrossRef]
  41. Bashir, O.; Piche, L.; Claverie, J.P. 18-electron ruthenium phosphine sulfonate catalysts for olefin metathesis. Organometallics 2014, 33, 3695–3701. [Google Scholar] [CrossRef]
  42. Ma, Y.; Reardon, D.; Gambarotta, S.; Yap, G.; Zahalka, H.; Lemay, C. Vanadium-catalyzed ethylene-propylene copolymerization:  the question of the metal oxidation state in Ziegler-Natta polymerization promoted by (β-diketonate)3V. Organometallics 1999, 18, 2773–2781. [Google Scholar] [CrossRef]
  43. Adisson, E.; Deffieux, A.; Fontanille, M.; Bujadoux, K. Polymerization of ethylene at high temperature by vanadium-based heterogeneous Ziegler—Natta catalysts. II. Study of the activation by halocarbons. J. Polym. Sci. Part A 1994, 32, 1033–1041. [Google Scholar] [CrossRef]
  44. Janas, Z.; Wisniewska, D.; Jerzykiewicz, L.B.; Sobota, P.; Drabent, K.; Szczegot, K. Synthesis, structural studies and reactivity of vanadium complexes with tridentate (OSO) ligand. Dalton Trans. 2007, 20, 2065–2069. [Google Scholar] [CrossRef] [PubMed]
  45. Carman, C.J.; Harrington, R.A.; Wilkes, C.E. Microstructure and physical properties of hydrochlorinated 1,4-polyisoprene prepared by butyllithium in nonpolar solvent. Macromolecules 1977, 10, 149–153. [Google Scholar]
  46. Randall, J.C. Methylene sequence distributions and number average sequence lengths in ethylene-propylene copolymers. Macromolecules 1978, 11, 33–36. [Google Scholar] [CrossRef]
  47. Cheng, H.N. Carbon-13 NMR analysis of ethylene-propylene rubbers. Macromolecules 1984, 17, 1950–1955. [Google Scholar] [CrossRef]
  48. Smith, W.V. Sequence distribution in ethylene-propylene copolymers. I. Relations between multads and between multads and the 13C NMR spectrum. J. Polym. Sci. B Polym. Phys. 1980, 18, 1573–1585. [Google Scholar] [CrossRef]
  49. Randall, J.C. Sequence distributions versus catalyst site behavior of in situ blends of polypropylene and poly(ethylene-co-propylene). J. Polym. Sci. Part A 1998, 36, 1527–1542. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of vanadium complexes 1a and 1b.
Scheme 1. Synthesis of vanadium complexes 1a and 1b.
Polymers 09 00325 sch001
Figure 1. Effect of Al/V ratio on on activity of vanadium-based catalysts in ethylene/propylene polymerization. Reaction conditions: [V] = 0.5 mmol/L, [ETAC]/[V] = 10, hexane 100 mL, 30 °C, 30 min, 1 bar of ethylene/propylene (1/2, molar ratio) pressure.
Figure 1. Effect of Al/V ratio on on activity of vanadium-based catalysts in ethylene/propylene polymerization. Reaction conditions: [V] = 0.5 mmol/L, [ETAC]/[V] = 10, hexane 100 mL, 30 °C, 30 min, 1 bar of ethylene/propylene (1/2, molar ratio) pressure.
Polymers 09 00325 g001
Figure 2. Effect of polymerization temperature on activity of vanadium-based catalysts in ethylene/propylene polymerization.
Figure 2. Effect of polymerization temperature on activity of vanadium-based catalysts in ethylene/propylene polymerization.
Polymers 09 00325 g002
Figure 3. Effect of polymerization time on activity of vanadium-based catalysts in ethylene/propylene polymerization.
Figure 3. Effect of polymerization time on activity of vanadium-based catalysts in ethylene/propylene polymerization.
Polymers 09 00325 g003
Figure 4. DSC curves of ethylene/propylene copolymers obtained with vanadium-based catalysts (left, VOCl3, Run 3; 1a, Run 6; 1b, Run 10 in Table 1) and partly enlarged curves (right).
Figure 4. DSC curves of ethylene/propylene copolymers obtained with vanadium-based catalysts (left, VOCl3, Run 3; 1a, Run 6; 1b, Run 10 in Table 1) and partly enlarged curves (right).
Polymers 09 00325 g004
Figure 5. 13C NMR spectrum of ethylene/propylene copolymer obtained with 1a/EASC catalyst (sample from Run 3 in Table 1).
Figure 5. 13C NMR spectrum of ethylene/propylene copolymer obtained with 1a/EASC catalyst (sample from Run 3 in Table 1).
Polymers 09 00325 g005
Figure 6. Monomer sequence distribution of ethylene/propylene copolymers determined from 13C NMR spectra (VOCl3, Run 3; 1a, Run 6; 1b, Run 10 in Table 1).
Figure 6. Monomer sequence distribution of ethylene/propylene copolymers determined from 13C NMR spectra (VOCl3, Run 3; 1a, Run 6; 1b, Run 10 in Table 1).
Polymers 09 00325 g006
Table 1. Ethylene/propylene copolymerization with V complex/Et3Al2Cl3 a.
Table 1. Ethylene/propylene copolymerization with V complex/Et3Al2Cl3 a.
RunCat.ETCA/VYield (g)Activity bMw × 10−4 cMw/Mn cC3 Incorp. (mol%) d
11a01.0341.223.13.1636.0
21a51.3052.012.72.0939.0
31a101.3252.88.33.1740.6
41a201.2248.84.02.6648.2
51b00.5020.028.13.0133.1
61b101.9678.419.62.6337.3
71b201.5562.06.12.7749.4
8VOCl300.9318.619.55.6925.1
9VOCl350.9638.410.37.2528.9
10VOCl3101.6064.06.47.6933.0
11VOCl3201.5361.22.25.4640.2
a Polymerization conditions: 100 mL of hexane, 30 °C, 30 min; catalyst, 50 μmol; Al/V = 40; propylene/ethylene (2/1, mole ratio) pressure, 1 atm; ETCA, ethyl trichloroacetate. b kg polymer (mol V)−1 h−1. c Determined by GPC. d Determined by 13C NMR.

Share and Cite

MDPI and ACS Style

Hao, X.; Zhang, C.; Li, L.; Zhang, H.; Hu, Y.; Hao, D.; Zhang, X. Use of Vanadium Complexes Bearing Naphthalene-Bridged Nitrogen-Sulfonate Ligands as Catalysts for Copolymerization of Ethylene and Propylene. Polymers 2017, 9, 325. https://doi.org/10.3390/polym9080325

AMA Style

Hao X, Zhang C, Li L, Zhang H, Hu Y, Hao D, Zhang X. Use of Vanadium Complexes Bearing Naphthalene-Bridged Nitrogen-Sulfonate Ligands as Catalysts for Copolymerization of Ethylene and Propylene. Polymers. 2017; 9(8):325. https://doi.org/10.3390/polym9080325

Chicago/Turabian Style

Hao, Xiufeng, Chundi Zhang, Lin Li, Hexin Zhang, Yanming Hu, Daifeng Hao, and Xuequan Zhang. 2017. "Use of Vanadium Complexes Bearing Naphthalene-Bridged Nitrogen-Sulfonate Ligands as Catalysts for Copolymerization of Ethylene and Propylene" Polymers 9, no. 8: 325. https://doi.org/10.3390/polym9080325

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