Next Article in Journal
Dual-Responsive Alginate Hydrogels for Controlled Release of Therapeutics
Next Article in Special Issue
Novel Cobalt Dichloride Complexes with Hindered Diphenylphosphine Ligands: Synthesis, Characterization, and Behavior in the Polymerization of Butadiene
Previous Article in Journal
Chitinases as Food Allergens
Previous Article in Special Issue
Activity and Thermal Stability of Cobalt(II)-Based Olefin Polymerization Catalysts Adorned with Sterically Hindered Dibenzocycloheptyl Groups
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Homo- and Copolymerization of Ethylene with Norbornene Catalyzed by Vanadium(III) Phosphine Complexes

CNR-Istituto per lo Studio delle Macromolecole (ISMAC), via A. Corti 12, I-20133 Milano, Italy
*
Author to whom correspondence should be addressed.
Present address: Loxeal srl, via G. Marconato 2, I-20811 Cesano Maderno (MB).
Molecules 2019, 24(11), 2088; https://doi.org/10.3390/molecules24112088
Submission received: 10 May 2019 / Revised: 29 May 2019 / Accepted: 30 May 2019 / Published: 31 May 2019
(This article belongs to the Special Issue Well-Defined Metal Complex Catalysts for Olefin Polymerization)

Abstract

:
Herein, we report the homo- and co-polymerization of ethylene (E) with norbornene (NB) catalyzed by vanadium(III) phosphine complexes of the type VCl3(PMenPh3-n)2 [n = 2 (1a), 1 (1b)] and VCl3(PR3)2 [R = phenyl (Ph, 1c), cyclohexyl (Cy, 1d), tert-butyl (tBu, 1e)]. In the presence of Et2AlCl and Cl3CCOOEt (ETA), 1a1e exhibit good activities for the polymerization of ethylene, affording linear, semicrystalline PEs with a melting temperature of approximately 130 °C. Mainly alternating copolymers with high comonomer incorporation were obtained in the E/NB copolymerization. A relationship was found between the electronic and steric properties of the phosphine ligands and the catalytic performance. Overall, the presence of electron-withdrawing ligand substituents increases the productivity, complexes with aryl phosphine (weaker σ–donor character) exhibiting a higher (co)polymerization initiation rate than those with alkyl phosphines (stronger σ–donor character). Steric effects also seem to play a key role since 1d and 1e, having large size phosphines (PCy3 θ = 170° and PtBu3 θ = 182°, respectively) are more active than 1a (PMe2Ph θ = 122°). In this case, the larger size of PtBu3 and PCy3 likely compensates for their higher donor strength compared to PMe2Ph.

Graphical Abstract

1. Introduction

Phosphines are one of the most important families of ligands in organometallic chemistry [1,2]. The synthesis, structure and reactivity of transition metal phosphine complexes continue to attract considerable attention due to their ability to undergo different transformations in which the phosphine is involved in the reactivity or acts as spectator ligand [3,4,5]. Phosphine ligands are typically constructed with combinations of alkyl and aryl groups, and both steric and electronic properties of the ligand can be modulated for specific applications. In addition, the relatively low cost and large variety of phosphines are incentives for the development of metal–phosphine complexes.
Vanadium is an element of growing importance. With multiple oxidation states and coordination number and flexible geometry, applications of vanadium complexes in catalysis and polymer science continue to grow [6,7,8,9,10]. Today, vanadium catalysts are largely used for the production of elastomers [11], high-molecular-weight poly(ethylene)s (PEs) with narrow molecular weight distribution [12,13], syndiotactic poly(propylene)s [14,15], and ethylene copolymers with cyclic olefins (COCs) [16,17] and α-olefins [18,19]. Among them, there is an increased interest in COCs because of the easy availability of the monomers and unique copolymer properties. The most versatile and interesting COCs are the copolymers of ethylene (E) with norbornene (NB), which have gained much attention as high-tech engineering plastics [20,21]. A significant breakthrough in the field of COCs was achieved in 1991 when Kaminsky discovered that C2-symmetric Group 4 metallocenes, in combination with methylalumoxane (MAO), catalyzed the copolymerization of ethylene with cyclic olefins [22]. This discovery also stimulated the development of post-metallocenes with early and late transition metals, expanding the commercially useful transition metals, but also broadening the range of COCs accessible by insertion polymerization [23].
For the reasons mentioned above concerning the versatility of vanadium complexes and the increasing relevance of COCs, in recent years, we became interested in the synthesis of vanadium complexes, and their application in the polymerization of various (di)olefins [16,17,24]. In this framework, recently, we reported that VCl3(PMePh2)2, in combination with Et2AlCl and ETA, catalyzed the E/NB copolymerization [24]. Mainly alternating copolymers, in the presence of a low excess of Et2AlCl and ETA, were successfully obtained. Vanadium–phosphine complexes have been known since the mid 1980s by the pioneer work of Caulton and Larkworthy [25,26,27,28], but have been less investigated in the context of olefin (co)polymerization with respect to vanadium complexes bearing multidentate chelating and anionic ligands having, in different combination, N and O hard donor atoms [29,30,31,32,33,34]. In this paper, we report the homo- and copolymerization of ethylene with NB catalyzed by a series of vanadium(III) complexes with tertiary phosphine of the type PMenPh3-n with n = 2 (1a), n = 1 (1b), and PR3 with R = Ph (1c), Cy (1d), tBu (1e). We found that a relationship exists between the electronic and steric properties of the phosphine ligand and the catalytic performance, and that the substituents on the phosphorus atom strongly affect the molecular weight of the resulting (co)polymers. The homo- and copolymerization of ethylene with NB using VCl3(THF)3 under the same set of conditions were performed as well for comparison.

2. Results and Discussion

2.1. Polymerization of Ethylene

First, 1a1e were screened for the polymerization of ethylene under the same set of conditions of 1 atm of ethylene, at 20 °C, in toluene. Et2AlCl was selected as co-catalyst due to exhibiting higher activities than MAO [24]. Indeed, the V–alkyl species formed by some vanadium complexes and Et2AlCl were found to be in equilibrium between chloro-bridged and cationic alkyl species due to the smaller bulkiness and stronger nucleophilic nature of Et2AlCl [6,10], while an isolated cationic species would be formed in the presence of MAO [10,29]. The polymerization experiments were carried out in the presence of ETA (50 equiv. to V) to overcome the main disadvantage of using vanadium(III) complexes, which is the rapid deactivation associated with the reduction to inactive vanadium(II) species. The results are listed in Table 1, and are presented as a function of the phosphine donor ability [as measured by the electronic parameter υCO based on the carbonyl stretching frequencies in Ni(CO)3L complexes (L = phosphine ligand)] and steric properties (measured by cone angle, θ) as defined by Tolman [35]. Table 1 also reports the results for the polymerization of ethylene by VCl3(THF)3 employed in this study for comparison.
All the vanadium phosphine complexes exhibited good activities for the polymerization of ethylene, (Figure 1A) affording linear, semicrystalline PEs with a melting temperature of approximately 130 °C (Figure S1 and S2 in Supplementary Materials). The activities of the vanadium phosphine complexes were lower than VCl3(THF)3, likely due to the faster displacement of the THF ligand (indeed THF is expected to be a weak donor ligand and thus to dissociate faster), but also due to different ratios of the dissociated ligand re-insertion rate relative to the ethylene insertion/coordination rate [24].
Generally, both the activity and the molecular weight of the resulting PEs were strongly affected by the substituents on the phosphorus atom, although the relationship between the catalytic activity, the polymer molecular weight and the phosphine ligand is not straightforward. This is because phosphine ligands are involved in dynamic “on/off” events, which include dissociative and associative pathways, and competitive reactions [24], and electronic and steric effects are intimately connected and difficult to separate [36,37,38,39].
Changing the substituents on the phosphorus atoms yielded large changes in the ligand donor properties: aryl phosphines (weaker σ-donor character) being weakly coordinated to the metal compared to alkyl phosphines (stronger σ-donor character). However, according to the parameters defined by Tolman [35], a general trend seems to emerge from data in Table 1. For complexes having at least one phenyl group on the phosphorus atoms, i.e., VCl3(PMenPh3-n)2 and VCl3(PPh3)2, the activity increased in the order 1a (PMe2Ph) < 1b (PMePh2) ≤ 1c (PPh3) with decreasing ligand donor strength (Table 1 and Figure 1A). Steric effects also play a role. In fact, 1e (PtBu3, the stronger σ-donor, θ = 182°) gave an activity close to that of 1b (PMePh2, θ = 136°): in this case, the bigger size of 1e compared to 1b more than compensates for its higher donor strength (Figure 1A). Moreover, the effect of the steric perturbation is better illustrated by the differences in activity between 1d (PCy3) and 1e (PtBu3). These two complexes have similar electronic properties (1d υ = 2056.1 cm1; 1e υ = 2056.4 cm1), but 1e exhibited an activity higher than 1d likely due to the larger cone angle of PtBu3 (θ = 182°) compared to that of PCy3 (θ = 170°) (Figure 1A).
PEs obtained from the vanadium phosphine complexes had higher molecular weight than the polymer obtained from VCl3(THF)3, and in some cases even two times more. This is likely because VCl3(THF)3 is more prone to β–H elimination, followed by a fast displacement of a vinyl-terminated polymer, than the vanadium phosphine complexes. Complexes containing phosphines with an increased σ-donor character gave PEs with higher molecular weight: 1e (PtBu3, the strongest σ-donor) gave the polymer with the highest molecular weight (Table 1, entry 5 Mw = 84,300 g/mol), while PPh3 (the weakest σ-donor) gave the polymer with the lowest molecular weight (Table 1, entry 3 Mw = 35,300 g/mol), close to that obtained from VCl3(THF)3 (Table 1, entry 6 Mw = 34,800 g/mol). It is worthwhile to note that the molecular weight of PEs obtained from complexes having aryl phosphines increased in the reverse order with respect to the activity; that is 1c (PPh3, Mw = 35,300 g/mol) < 1b (PMePh2, 44,600) < 1a (PMe2Ph, 65,500).

2.2. Copolymerization of Ethylene with Norbornene

Then, 1a1e were used for the E/NB copolymerization under the same set of conditions of 1 atm of ethylene, at 20 °C, in toluene, with Et2AlCl (500 equiv to V), ETA (50 equiv to V) and for a NB/E feed ratio of 4. The copolymerizations were performed at short reaction times to ensure the uniformity of the catalytic process, even in the presence of a very high comonomer consumption in the early instants of the reaction, and thus to obtain copolymers with uniform composition [16,17,24]. The results are summarized in Table 2.
None of the investigated complexes was active in the homopolymerization of NB under the conditions employed, while high activities were obtained for the copolymerization of E/NB in the first seconds of the reaction. When both the comonomers are available for the active intermediates, all the complexes are instantaneously activated and the copolymerization proceeds at significant rates, particularly at the initial stage. The increased activity could be due to the coordination of the nucleophilic and sterically demanding NB that, in some cases, is expected to stabilize the active site [40].
In general, we found the same trend in activity observed for the homopolymerization of ethylene, complexes with aryl phosphines being more active than those with alkyl phosphines (Figure 1B vs. Figure 1A). For complexes having at least one phenyl group on the phosphorus atoms, i.e., 1a1c, the activity increased in the order 1a (PMe2Ph) < 1b (PMePh2) < 1c (PPh3), 1c being the most active according to the lower donor strength of PPh3 (Figure 1B). Moving to complexes bearing alkyl phosphines, i.e., 1d and 1e, the activity increased with respect to 1a, likely due to the larger cone angle of PtBu3 and PCy3 (θ = 182 and 170°, respectively) despite the higher donor strength of PtBu3 and PCy3 compared to PMe2Ph (θ = 145°). In this case, the bigger size of 1d and 1e compared to 1b more than compensates for their higher donor strength (Figure 1B). Also, of note, moving from 1d to 1e, the activity did not increase to the same extent as found for the homopolymerization of ethylene (Figure 1A vs. Figure 1B). This may be likely due to the increased steric interaction between the bulky NB and the bulky tert-butyl substituents on the phosphorous atoms in 1e, which would offer a significant impediment to the coordination and insertion of the bulkier comonomer, thus depressing propagation rate.
As observed in the case of ethylene polymerization, the average molecular weight of the copolymers generally increased with increasing the ligand σ-donor character, in the order 1c < 1b < 1a1d < 1e, 1c (PPh3) giving the copolymer with the lowest molecular weight (Mw = 42,600 g/mol) close to that obtained with VCl3(THF)3 (Mw = 43,600 g/mol). Overall, the increase in the ligand σ–donor ability has a positive effect on the (co)polymer molecular weight: a stronger electron-donating ligand may stabilize the electron-deficient vanadium intermediate and reduce the rate of β–H elimination and subsequent chain-transfer [17,41]. Conversely, electron-withdrawing ligand substituents reduce the electron density at the metal center, resulting in higher rates of β–H elimination, but significantly enhancing catalytic performance. This could be explained assuming that β–H elimination is the rate-determining step. Since the cyclic olefin prevents β–H elimination because of Bredt’s rule, the molecular weight of the copolymers is generally higher than that of PEs.
On the other hand, the type of phosphine ligand had no significant effect on the NB incorporation and copolymer microstructure. All the copolymers have a NB content ranging from 37.8 to 39.3 mol%, regardless of the steric and electronic properties of the phosphine ligand. The same result was found by some of us [16,17], and by Yue-Sheng Li using V(III) complexes bearing β-enaminoketonato ligands [33] and different tridentate salicyladiminato ligands [34].
The copolymer microstructure was investigated by 13C-NMR. The signals of each chemical shift region were assigned, according to the literature [42], as follows: 42.0–54.0 ppm, C2/C3; 34.5–42.0 ppm, C1/C4; 31.0–34.5 ppm, C7; 26.0–31.0 ppm, C5/C6 and ethylene CH2 (Figure 2). The 13C-NMR spectra of all the copolymers show the typical pattern of alternating copolymers from an addition-type copolymerization of NB with cis exo-exo enchainment. The dominant signals are those of isolated and alternating NB units, with traces of diads. As an example, the 13C-NMR spectrum of entry 8 (Table 2, NB = 37.8 mol%) is shown in Figure 2.
The relative intensities of peaks at 45.8 and 45.2 ppm, assigned to C2/C3 of NB in the alternating isotactic and syndiotactic NB–E–NB–E–NB sequences, respectively, also revealed that the copolymers have a random tacticity, the ratio of isotactic and syndiotactic units being in the range from 35:32 to 28:33.
The analysis of copolymer thermal properties revealed no melting events, but only a unique glass transition temperature (Tg) in the range from 73 to 93 °C, thus suggesting that the copolymers are amorphous possibly with a homogeneous composition [43,44]. It is expected that for random copolymers with comparable Mw, Tg increases with increasing NB incorporation in the copolymers [45,46]. However, as it can be seen in Table 2 and Figure 3, the copolymer generated with VCl3(THF)3 (sample 12, Mw = 43,600 g/mol), despite the highest content of NB (41.2 mol%), has the lowest Tg (73 °C) even when compared with sample 9 of the same molecular weight (Mw = 42,600 g/mol, Tg = 81 °C). We attribute this result to the fact that sample 12 is probably less homogenous in terms of composition distribution, likely due to a faster initiation efficiency of VCl3(THF)3 and hence to a more pronounced compositional drift. This leads to the formation of NB-rich copolymer chains at the beginning of the copolymerization and E-rich chains at the end, causing a non-homogeneous distribution of the comonomer that helps lower the Tg [17,24].

3. Materials and Methods

3.1. General Procedures and Materials

The manipulation of air- and/or moisture-sensitive reagents was performed under an inert atmosphere using a dual vacuum/nitrogen line and standard Schlenk-line techniques. Ethylene and nitrogen were purified by passage over columns of CaCl2 and molecular sieves. Oxygen was removed by fluxing the gases through BTS catalysts. Toluene (Aldrich, St. Louis, MO, USA, 99.5%) was refluxed over Na for 8 h and then distilled and stored over molecular sieves. ETA (Aldrich, 97%) was stirred over CaH2 for 4 h and then distilled trap-to-trap. Et2AlCl (Aldrich), and VCl3(THF)3 (Aldrich) were used as received without further purification. NB (Aldrich, 99%) was stirred over molten potassium at 80 °C under nitrogen for 4 h and then distilled. A stock solution was prepared by dissolving 50 g of freshly distilled NB in 86.2 mL of toluene. A deuterated solvent for NMR measurements (C2D2Cl4) (Aldrich, >99.5% atom D) was used as received. All the vanadium phosphine complexes were synthesized according to that already reported in the literature [47].

3.2. General (co)polymerization Procedure

Polymerizations were carried out in a 25-mL round-bottomed Schlenk flask. Prior to starting polymerization, the reactor was heated to 110 °C under vacuum for 1 h and backfilled with nitrogen. The reactor was charged with toluene, NB, ETA and Et2AlCl in that order, and then the solution was degassed, and ethylene was added until saturation. Polymerization was started by adding a toluene solution (2 mg/mL) of a vanadium complex via syringe under continuous flow of ethylene. Polymerizations were stopped with methanol containing a small amount of hydrochloric acid; the precipitated polymers were collected by filtration, repeatedly washed with fresh methanol and finally dried in vacuum at room temperature to constant weight.

3.3. Characterization

NMR spectra were recorded on a Bruker (Bruker Italia Srl, Milano, Italy) NMR advance 400 Spectrometer operating at 400 MHz (1H) and 100.58 MHz (13C), working in the PFT mode at 103 °C. The 13C experiments were performed with a 10-mm probe in C2D2Cl4 and referred to hexamethyldisiloxane as the internal standard. Thermal analysis was performed with a differential scanning calorimeter (DSC) Perkin Elmer 8000 (Waltham, MA, USA)). The scans were recorded in a flowing nitrogen atmosphere at scanning rates of 10 °C/min. The molecular weight average (Mw) and the molecular weight distribution (Mw/Mn) were obtained by a high temperature Waters GPCV2000 (Milford, MA, USA) size-exclusion chromatography (SEC) system equipped with a refractometer detector. The experimental conditions consisted of three PL Gel Olexis columns, ortho-dichlorobenzene (o-DCB) as the mobile phase, a 0.8-mL/min flow rate, and 145 °C temperature. The calibration of the SEC system was constructed using eighteen narrow Mw/Mn poly(styrene) standards with Mws ranging from 162 to 5.6 × 106 g/mol. For SEC analysis, approximately 12 mg of polymer was dissolved in 5 mL of DCB with 0.05% of BHT as antioxidant.

4. Conclusions

Vanadium(III) complexes bearing commercially available monodentate phosphines exhibit good activity for the polymerization of ethylene and the E/NB copolymerization in the presence of Et2AlCl and ETA. Introduction of NB in the polymerization mixture accelerates the propagation rate and restrains chain termination to some extent. Alternating copolymers with high NB incorporation were produced within the first copolymerization seconds so that copolymers with a homogeneous composition were obtained only by keeping a short reaction time to limit the undesired compositional drift.
A relationship was found to exist between the phosphine electronic and steric properties and the catalytic performance. Overall, electron-withdrawing ligand substituents significantly enhance catalytic performance, vanadium complexes with aryl phosphine (weaker σ–donor character) exhibiting a higher initiation rate than those with alkyl phosphines (stronger σ–donor character). Steric properties also have a relevant effect, with complexes with great steric hindrance, such as PCy3 and PtBu3 ligands, exhibiting a higher initiation rate than PMe2Ph (weaker σ–donor). In this case, the larger cone angle of PCy3 and PtBu3 more than compensates for their higher donor strength compared to PMe2Ph. The type of phosphine ligand also strongly affects the (co)polymer molecular weight. Complexes with electron-donating phosphines afford (co)polymers with higher molecular weight, likely because an electron-donating ligand stabilizes the electron-deficient catalytic intermediate, reducing the rate of chain transfer.

Supplementary Materials

The following are available online. Figure S1: 13C and 1H-NMR spectra of a selected PE. Figure S2: (a) DSC curve, and (b) XRD pattern of a selected PE.

Author Contributions

G.Z., G.R. and G.L. conceived and designed the experiments; G.Z., A.G. and F.B. performed the experiments; G.Z., A.G., F.B. and G.L. analyzed the data; G.Z. and G.L. wrote the paper.

Funding

This research received no external funding.

Acknowledgments

The authors are grateful to F. Greco for the acquisition of NMR spectra and D. Piovani for SEC measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Burt, J.; Levason, W.; Reid, G. Coordination chemistry of the main group elements with phosphine, arsine and stibine ligands. Coord. Chem. Rev. 2009, 260, 65–115. [Google Scholar] [CrossRef]
  2. Cheng, C.; Hartwig, J.F. Rhodium-catalyzed intermolecular C-H silylation of arenes with high steric regiocontrol. Science 2014, 343, 853–857. [Google Scholar] [CrossRef]
  3. Scheuermann, M.L.; Johnson, E.J.; Chirik, P.J. Alkene Isomerization–Hydroboration Promoted by Phosphine-Ligated Cobalt Catalysts. Org. Lett. 2015, 17, 2716–2719. [Google Scholar] [CrossRef] [PubMed]
  4. Murai, M.; Takeshima, H.; Morita, H.; Kuninobu, Y.; Takai, K. Acceleration Effects of Phosphine Ligands on the Rhodium-Catalyzed Dehydrogenative Silylation and Germylation of Unactivated C(sp3)–H Bonds. J. Org. Chem. 2015, 80, 5407–5414. [Google Scholar] [CrossRef]
  5. Ricci, G.; Sommazzi, A.; Masi, F.; Ricci, M.; Boglia, A.; Leone, G. Well Defined Transition Metal Complexes with Phosphorus and Nitrogen Ligands for 1,3-Dienes Polymerization. Coord. Chem. Rev. 2010, 254, 661–676. [Google Scholar] [CrossRef]
  6. Nomura, K.; Zhang, S. Design of Vanadium Complex Catalysts for Precise Olefin Polymerization. Chem. Rev. 2011, 111, 2342–2362. [Google Scholar] [CrossRef] [PubMed]
  7. Gambarotta, S. Vanadium-based Ziegler–Natta: challenges, promises, problems. Coord. Chem. Rev. 2003, 237, 229–243. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  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. Christman, D.L.; Keim, G.I. Reactivities of Nonconjugated Dienes Used in Preparation of Terpolymers in Homogeneous Systems. Macromolecules 1968, 1, 358–363. [Google Scholar] [CrossRef]
  12. Zambelli, A.; Pasquon, I.; Signorini, R.; Natta, G. Polymerization of propylene to syndiotactic polymer. III. Behaviour of the catalyst system VCl4–Al(C2H5)2Cl in the presence of Lewis bases. Makromol. Chem. 1968, 112, 60–182. [Google Scholar]
  13. Junghanns, E.; Gumboldt, O.; Bier, G. Polymerization of ethylene and propylene to amorphous copolymers with catalysts of vanadium oxychloride and alkyl aluminum halides. Makromol. Chem. 1962, 58, 18–42. [Google Scholar] [CrossRef]
  14. Murphy, V.J.; Turner, H. Low-Coordinate (Arylimido) vanadium(V) Alkyls:  Synthesis and Reactivity of V(NAr)(CH2Ph)3 (Ar = C6H3-2,6-iPr2). Organometallics 1997, 16, 2495–2497. [Google Scholar] [CrossRef]
  15. Doi, Y.; Suzuki, S.; Soga, K. Living coordination polymerization of propene with a highly active vanadium-based catalyst. Macromolecules 1986, 19, 2896–2900. [Google Scholar] [CrossRef]
  16. Zanchin, G.; Vendier, L.; Pierro, I.; Bertini, F.; Ricci, G.; Lorber, C.; Leone, G. Homo- and co-polymerization of ethylene with cyclic olefins catalyzed by phosphine adducts of (imido)vanadium(IV) complexes. Organometallics 2018, 37, 3181–3195. [Google Scholar] [CrossRef]
  17. Zanchin, G.; Pierro, I.; Parisini, E.; Martì-Rujas, J.; Ricci, G.; Leone, G. Synthesis, structure and behavior of vanadium(III) diphosphine complexes in the homo- and co-polymerization of ethylene with norbornene: The ligand donor strength and bite angle make the difference. J. Organomet. Chem. 2018, 861, 142–150. [Google Scholar] [CrossRef]
  18. Zhang, S.; Zhang, W.C.; Shang, D.D.; Wu, Y.X. Synthesis of ultra-high-molecular-weight ethylene-propylene copolymer via quasi-living copolymerization with N-heterocyclic carbene ligated vanadium complexes. J. Polym. Sci. A: Polym. Chem. 2019, 57, 553–561. [Google Scholar] [CrossRef]
  19. Bialek, M.; Bisz, E. Dichlorovanadium(IV) diamine-bis(phenolate) complexes for ethylene (co)polymerization and 1-olefin isospecific polymerization. J. Catal. 2018, 362, 65–73. [Google Scholar] [CrossRef]
  20. Lamonte, R.R.; Mc Nally, D. Cyclic olefin copolymers. Adv. Mater. Process. 2001, 3, 33–36. [Google Scholar]
  21. Nunes, S.P.; Ohlsson, P.D.; Ordeig, O.; Kutter, P.J. Cyclic olefin polymers: Emerging materials for lab-on-a-chip applications. Microfluid. Nanofluid. 2010, 9, 145–161. [Google Scholar] [CrossRef]
  22. Kaminsky, W.; Bark, A.; Arndt, M. New polymers by homogenous zirconocene/aluminoxane catalysts. Macromol. Symp. 1991, 47, 83–93. [Google Scholar] [CrossRef]
  23. Li, X.; Hou, Z. Organometallic catalysts for copolymerization of cyclic olefins. Coord. Chem. Rev. 2008, 252, 1842–1869. [Google Scholar] [CrossRef]
  24. Leone, G.; Pierro, I.; Zanchin, G.; Forni, A.; Bertini, F.; Rapallo, A.; Ricci, G. Vanadium(III)–catalyzed copolymerization of ethylene with norbornene: Microstructure at tetrad level and reactivity ratios. J. Mol. Catal. A: Chem. 2016, 424, 220–231. [Google Scholar] [CrossRef]
  25. Bansemer, R.L.; Huffman, J.C.; Caulton, K.G. Synthesis and characterization of VCl3(PMePh2)2. Inorg. Chem. 1985, 24, 3003–3006. [Google Scholar] [CrossRef]
  26. Holt, D.G.L.; Larkworthy, L.F.; Povey, D.C.; Smith, G.W.; Leigh, G.J. Synthesis and structural studies of complexes of vanadium(II) and vanadium(III) halides with tertiary phosphines. Inorg. Chim. Acta 1993, 207, 11–19. [Google Scholar] [CrossRef]
  27. Bultitude, J.; Larkworthy, L.F.; Povey, D.C.; Smith, G.W.; Dilworth, J.R. The preparation and crystal and molecular structures of trichlorobis(methyldiphenylphosphine)vanadium(III) and its acetonitrile adduct. J. Chem. Soc. Dalton Trans. 1986, 2253–2258. [Google Scholar] [CrossRef]
  28. Nieman, J.; Teuben, J.H.; Huffman, J.C.; Caulton, K.G. Preparation and characterization of mono-cyclopentadienylvanadium dihalide bis-phosphine complexes -Crystal-structure of (η-5-C5H5)VCl2(PMe3)2. J. Org. Chem. 1983, 255, 193–204. [Google Scholar] [CrossRef]
  29. Wang, W.; Nomura, K. Notable effects of aluminum alkyls and solvents for highly efficient ethylene (co)polymerizations catalyzed by (arylimido)-(aryloxo)vanadium complexes. Adv. Synth. Catal. 2006, 348, 743–750. [Google Scholar] [CrossRef]
  30. Wang, J.B.; Lu, L.P.; Liu, J.Y.; Li, Y.S. [ONN]-type amine pyridine (s) phenolate-based oxovanadium (v) catalysts for ethylene homo-and copolymerization. Dalton Trans. 2014, 43, 12926–12934. [Google Scholar] [CrossRef]
  31. Wu, J.Q.; Mu, J.S.; Zhang, S.W.; Li, Y.S. Vanadium(V) complexes containing tetradentate amine trihydroxy ligands as catalysts for copolymerization of cyclic olefins. J. Polym. Sci. Part A Pol. Chem. 2010, 48, 1122–1132. [Google Scholar] [CrossRef]
  32. Lorber, C.; Wolff, F.; Choukroun, R.; Vendier, L. Ethylene Homo- and Copolymerization Activity of a Series of [ONNO]-Type Amine Bis(phenolate) Based Vanadium(II–V) Catalysts. Eur. J. Inorg. Chem. 2005, 2850–2859. [Google Scholar] [CrossRef]
  33. Tang, L.M.; Wu, J.Q.; Duan, Y.Q.; Pan, L.; Li, Y.G.; Li, Y.S. Ethylene polymerizations, and the copolymerizations of ethylene with hexene or norbornene with highly active mono(β-enaminoketonato) vanadium(III) catalysts. J. Polym. Sci. Part A Pol. Chem. 2008, 46, 2038–2048. [Google Scholar] [CrossRef]
  34. Wu, J.Q.; Pan, L.; 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]
  35. Tolman, A.C. Steric effects of phosphorus ligands in organometallic chemistry and homogeneous catalysis. Chem. Rev. 1977, 77, 313–348. [Google Scholar] [CrossRef]
  36. Fernandez, A.; Reyes, C.; Wilson, M.R.; Woska, D.C.; Prock, A.; Giering, W.P. Examination of Drago’s EB and CB parameters for phosphines through the Quantitative Analysis of Ligand Effects (QALE). Organometallics 1997, 16, 342–348. [Google Scholar] [CrossRef]
  37. Woska, D.C.; Prock, A.; Giering, W.P. Determination of the stereoelectronic parameters of PF3, PCl3, PH3, and P(CH2CH2CN)(3). The quantitative analysis of ligand effects (QALE). Organometallics 2000, 19, 4629–4638. [Google Scholar] [CrossRef]
  38. Joerg, S.; Drago, R.S.; Sales, J. Reactivity of phosphorus donors. Organometallics 1998, 17, 589–599. [Google Scholar] [CrossRef]
  39. Bent, H.A. An Appraisal of Valence-bond Structures and Hybridization in Compounds of the First-row elements. Chem. Rev. 1961, 61, 275–311. [Google Scholar] [CrossRef]
  40. Yoshida, Y.; Mohri, J.I.; Ishii, S.I.; Mitani, M.; Saito, J.; Matsui, S.; Makio, H.; Nakano, T.; Tanaka, H.; Onda, M.; et al. Living Copolymerization of Ethylene with Norbornene Catalyzed by Bis(Pyrrolide−Imine) Titanium Complexes with MAO. J. Am. Chem. Soc. 2004, 126, 12023–12032. [Google Scholar] [CrossRef] [PubMed]
  41. Popeney, C.S.; Guan, Z. Effect of Ligand Electronics on the Stability and Chain Transfer Rates of Substituted Pd(II) α-Diimine Catalysts. Macromolecules 2010, 43, 4091–4097. [Google Scholar] [CrossRef]
  42. Leone, G.; Mauri, M.; Losio, S.; Bertini, F.; Ricci, G.; Porri, L. Copolymerization of ethylene with α-olefins and cyclic olefins catalyzed by a Ti(IV) diisopropoxy complex bearing a tridentate [O,S,O]-type bis(phenolato) ligand. Polim. Chem. 2014, 5, 3412–3423. [Google Scholar] [CrossRef]
  43. Simanke, A.G.; Alamo, R.G.; Galland, G.B.; Mauler, R. Wide-Angle X-ray Scattering of Random Metallocene−Ethylene Copolymers with Different Types and Concentration of Comonomer. Macromolecules 2001, 34, 6959–6971. [Google Scholar] [CrossRef]
  44. Arndt, M.; Beulich, I. C1-symmetric metallocenes for olefin polymerisation, 1. Catalytic performance of [Me2C(3-tertBuCp)(Flu)]ZrCl2 in ethene/norbornene copolymerisation. Macromol. Chem. Phys. 1998, 199, 1221–1232. [Google Scholar] [CrossRef]
  45. Hasan, T.; Ikeda, T.; Shiono, T. Ethene−Norbornene Copolymer with High Norbornene Content Produced by ansa-Fluorenylamidodimethyltitanium Complex Using a Suitable Activator. Macromolecules 2004, 37, 8503–8509. [Google Scholar] [CrossRef]
  46. Kiesewetter, J.; Arikan, B.; Kaminsky, W. Copolymerization of ethene with norbornene using palladium(II) α-diimine catalysts: Influence of feed composition, polymerization temperature, and ligand structure on copolymer properties and microstructure. Polymer 2006, 47, 3302–3314. [Google Scholar] [CrossRef]
  47. Leone, G.; Zanchin, G.; Pierro, I.; Sommazzi, A.; Forni, A.; Ricci, G. Synthesis, structure and 1,3-Butadiene polymerization behavior of Vanadium(III) phosphine complexes. Catalysts 2017, 7, 369. [Google Scholar] [CrossRef]
Sample Availability: Not available.
Figure 1. Plot of activity, calculated as gpol/molV vs. the vanadium complexes employed [PPh3 (1c), PMePh2 (1b), PMe2Ph (1a), PCy3 (1d) and PtBu3 (1e)] for (A) the polymerization of ethylene and (B) the copolymerization of ethylene with norbornene (NB). In each bar, the phosphine cone angle (°) is reported.
Figure 1. Plot of activity, calculated as gpol/molV vs. the vanadium complexes employed [PPh3 (1c), PMePh2 (1b), PMe2Ph (1a), PCy3 (1d) and PtBu3 (1e)] for (A) the polymerization of ethylene and (B) the copolymerization of ethylene with norbornene (NB). In each bar, the phosphine cone angle (°) is reported.
Molecules 24 02088 g001
Figure 2. 13C-NMR spectrum of entry 8 (Table 2, NB = 37.8 mol%).
Figure 2. 13C-NMR spectrum of entry 8 (Table 2, NB = 37.8 mol%).
Molecules 24 02088 g002
Figure 3. DSC curves (second heating) of E/NB copolymers obtained with (a) VCl3(THF)3 (entry 12, NB = 41.2 mol%), (b) 1c (entry 9, NB = 38.7 mol%), (c) 1e (entry 11, NB = 38.8 mol%), (d) 1a (entry 7, NB = 38.0 mol%), (e) 1b (entry 8, NB = 37.8 mol%) and (f) 1d (entry 10, NB = 39.3 mol%).
Figure 3. DSC curves (second heating) of E/NB copolymers obtained with (a) VCl3(THF)3 (entry 12, NB = 41.2 mol%), (b) 1c (entry 9, NB = 38.7 mol%), (c) 1e (entry 11, NB = 38.8 mol%), (d) 1a (entry 7, NB = 38.0 mol%), (e) 1b (entry 8, NB = 37.8 mol%) and (f) 1d (entry 10, NB = 39.3 mol%).
Molecules 24 02088 g003
Table 1. Polymerization of ethylene catalyzed by 1a1e in combination with Et2AlCl.1.
Table 1. Polymerization of ethylene catalyzed by 1a1e in combination with Et2AlCl.1.
EntryV-catPhosphineYield (mg)Activity 4 (×10 2)Mw5 (×10 3)Mw/Mn5
(type)υCO 2 (cm−1)θ3 (°)
11aPMe2Ph2065.312219539065.56.8
21bPMePh22067.013628156244.63.9
31cPPh32068.914530260435.34.2
41dPCy32056.417022144256.96.3
51ePtBu32056.118229559084.27.9
6VCl3(THF)3 36172234.75.0
1 Polymerization conditions: ethylene pressure, 1.01 bar; total volume, 25 mL (toluene); V-complex, 5 μmol; Al/V = 500; ETA/V = 50; temperature, 22 °C; time, 2 min; 2 measure of the electron-donating properties of phosphine ligands, as reported by Tolman [35]; 3 phosphine cone angle, as reported by Tolman [35]; 4 activity in gpol/molV; 5 determined by size-exclusion chromatography (SEC).
Table 2. Copolymerization of E/NB catalyzed by 1a1e in combination with Et2AlCl and ETA.1.
Table 2. Copolymerization of E/NB catalyzed by 1a1e in combination with Et2AlCl and ETA.1.
EntryV-catPhosphineYield (mg)Activity 4 (×10 2)NB 5 (mol%)Mw6 (×10 3)Mw/Mn6Tg7 (°C)
(type)υCO 2 (cm−1)θ3 (°)
71aPMe2Ph2065.312216032038.090.34.290
81bPMePh22067.013624148237.861.43.890
91cPPh32068.914536773438.742.64.281
101dPCy32056.417022645239.386.45.993
111ePtBu32056.118224749438.8130.07.085
12VCl3(THF)3 39579041.243.64.473
1 Polymerization conditions: ethylene pressure, 1.01 bar; total volume, 25 mL (toluene); V-complex, 5 μmol; Al/V = 500; ETA/V = 50 temperature, 22 °C; time, 10 s; [NB]/[E] feed ratio, 4 mol/mol; NB feedstock concentration, 0.57 mol/L; 2 measure of the electron-donating properties of phosphine ligands, as reported by Tolman [35]; 3 phosphine cone angle, as reported by Tolman [35]; 4 activity in gpol/molV; 5 determined by 13C-NMR; 6 determined by SEC; 7 determined by differential scanning calorimeter (DSC).

Share and Cite

MDPI and ACS Style

Zanchin, G.; Gavezzoli, A.; Bertini, F.; Ricci, G.; Leone, G. Homo- and Copolymerization of Ethylene with Norbornene Catalyzed by Vanadium(III) Phosphine Complexes. Molecules 2019, 24, 2088. https://doi.org/10.3390/molecules24112088

AMA Style

Zanchin G, Gavezzoli A, Bertini F, Ricci G, Leone G. Homo- and Copolymerization of Ethylene with Norbornene Catalyzed by Vanadium(III) Phosphine Complexes. Molecules. 2019; 24(11):2088. https://doi.org/10.3390/molecules24112088

Chicago/Turabian Style

Zanchin, Giorgia, Alessia Gavezzoli, Fabio Bertini, Giovanni Ricci, and Giuseppe Leone. 2019. "Homo- and Copolymerization of Ethylene with Norbornene Catalyzed by Vanadium(III) Phosphine Complexes" Molecules 24, no. 11: 2088. https://doi.org/10.3390/molecules24112088

Article Metrics

Back to TopTop