Next Article in Journal
Feature Papers to Celebrate the Landmarks of Catalysts
Next Article in Special Issue
Pyrazole Supported Zinc(II) Benzoates as Catalysts for the Ring Opening Copolymerization of Cyclohexene Oxide and Carbon Dioxide
Previous Article in Journal
Effects of Catalyst Preparation on Hydrocarbon Product Distribution in Hydrocracking of the Fischer-Tropsch Product with Low Pt-Loaded Catalysts
Previous Article in Special Issue
Molybdenum (VI) Imido Complexes Derived from Chelating Phenols: Synthesis, Characterization and ɛ-Caprolactone ROP Capability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Ethylene or Propylene/1,3-Butadiene Copolymers Possessing Pendant Vinyl Groups with Virtually No Internal Olefins

1
Mitsui Chemicals Inc., 1-2, Waki 6-chome, Waki-cho, Kuga-gun, Yamaguchi 740-0061, Japan
2
Mitsui Chemicals Inc., 580-32 Nagaura, Sodegaura, Chiba 299-0265, Japan
*
Authors to whom correspondence should be addressed.
Catalysts 2015, 5(4), 2001-2017; https://doi.org/10.3390/catal5042001
Submission received: 7 October 2015 / Accepted: 12 November 2015 / Published: 20 November 2015
(This article belongs to the Special Issue Molecular Catalysis for Precise Olefin Polymerization and ROP 2015)

Abstract

:
In general, ethylene/1,3-butadiene copolymerizations provides copolymers possessing both pendant vinyls and vinylenes as olefinic moieties. We, at MCI, studied the substituent effects of C2-symmetric zirconocene complexes, rac-[Me2Si(Indenyl’)2]ZrCl2 (Indenyl’ = generic substituted indenyl), after activation on the ratio of the pendant vinyls and vinylenes of the resultant copolymers. Complexes examined in this study were rac-dimethylsilylbis(1-indenyl)zirconium dichloride (1), rac-dimethylsilyl-bis [1-(2-methyl-4,5-benzoindenyl)]zirconium dichloride (2), rac-dimethylsilyl-bis[l-(2-methyl -4-phenylindenyl)]zirconium dichloride (3), rac-dimethy1si1y1-bis(2-ethyl-4-phenylindenyl) zirconium dichloride (4), rac-dimethylsilyl-bis[l-(2-n-propyl -4-(1-naphthyl)indenyl)]zirconium dichloride (5), rac-dimethylsilyl-[1-(2-ethyl-4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta [a]naphthalenyl)indenyl))][1-(2-n-propyl-4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta[a] naphthalenyl)indenyl))]zirconium dichloride (6), rac-dimethylsilyl-bis[1-(2-ethyl -4-(9-phenanthryl)indenyl)]zirconium dichloride (7), and rac-dimethylsilyl-bis [l-(2-n-propyl-4-(9-phenanthryl)indenyl)]zirconium dichloride (8). We found that the ratio of the pendant vinyls and vinylenes is strongly affected by the bulkiness of the substituent on the complexes examined. The vinyl content increased linearly in the following order, 8 > 7 > 6 > 5 > 4 > 3 > 2 > 1. Notably, complex 8/DMAO formed ethylene/1,3-butadiene copolymers possessing predominant vinyl groups, which can be crucial precursors for functionalized polyolefins. Likewise, complex 8/DMAO afforded propylene/1,3-butadiene copolymers with predominant vinyl groups.

Graphical Abstract

1. Introduction

Polyolefins, represented by polyethylene, polypropylene, and amorphous ethylene/α-olefin copolymers, can display useful mechanical properties (high impact strength, hardness/softness, durability), chemical stability (low reactivity, oil resistance, weathering resistance), thermal stability (heat/cold resistance), processability, lightweight, cost-effectiveness, and recyclability. However, due to their non-polar nature, they show poor affinity for polar materials such as polar organic polymers, glasses, and metals. To overcome this drawback, several attempts have been made in an effort to introduce polar groups into polyolefins, which include copolymerization of ethylene with polar monomers via a high pressure radical polymerization and the modification of polyolefins by a radical reaction [1]. However, these methods have disadvantages to control over polymer architecture (e.g., contents and distributions of polar groups, product molecular weights, molecular weight distributions, and branching extent) due to inherently poor reaction selectivity and unfavorable side reactions. Conversely, the direct synthesis of functionalized polyolefins by the anionic insertion copolymerization of olefins and polar CH2=CHX monomers (X = polar groups) using organometallic complexes has extensively been investigated [1]. Although these investigations have made an impact on polymer synthesis, these are far from being of practical use because of low productivity and a low degree of polar monomer incorporation.
In this study, we focused on the copolymerization of butadiene with olefins for the purpose of creating new functionalized polyolefinic materials using cost-competitive feedstocks. Because butadiene is a conventional monomer, yet it is able to generate a vinyl group onto the polymer backbone. The pendant vinyl group can be converted to functional groups, or utilized by itself. Thus it is that we explored catalysts that can copolymerize butadiene with olefin(s), where the insertion reaction of the butadiene proceeds preferentially via a 1,2-insertion without the accompanying subsequent cyclization, since the formation of the internal olefin, the cause of the degradation of the polymer generated by the 1,4-insertion, is unfavorable.
However, upon copolymerizing butadiene with olefins, several kinds of reaction pathways are possible (Scheme 1) [2,3,4,5]. η2 primary coordination of butadiene leads to formation of vinyl groups or rings through a primary vinyl insertion, whereas η2 secondary and η4-cis coordination of butadiene leads to trans and cis 1,4-butadiene units through the formation of a syn-η3-coordinated growing chain and an anti-η3-coordinated growing chain, respectively [6,7,8,9]. Until now, reports on olefin/butadiene copolymerization, where the insertion reaction of the 1,3-butadiene proceeds via 1,2-insertion preferentially to form a vinyl group, are scarce.
Scheme 1. Reaction pathways of ethylene/1,3-butadiene copolymerization.
Scheme 1. Reaction pathways of ethylene/1,3-butadiene copolymerization.
Catalysts 05 02001 g007
A rare example of such polymerization is propylene/butadiene copolymerization using rac-[Me2Si (2-methyl-4-phenylindenyl)2]ZrCl2 in the presence of hydrogen [10,11]. However, while this method provides a copolymer having pendant vinyl groups in the main chain with a minimal number of internal olefins, there is still room for investigation regarding the substituent effects on the number of both internal olefins and vinyl groups.
Thus, we preliminary investigated a wide variety of catalysts including metallocene and post-metallocene catalysts to find catalysts that are capable of generating copolymers with predominant pendant vinyl groups in ethylene/butadiene copolymerization [12]. As a result, we found a qualitative correlation between the catalyst structure and the insertion mode of butadiene. As a general rule, catalysts incorporating α-olefins with 1,2-insertion favor the 1,2-insertion of butadiene. On the other hand, catalysts incorporating α-olefins wit 2,1-insertion prefer the 1,4-insertion of butadiene. Namely, the catalysts inserting α-olefins by the 1,2-insertion tend to produce olefin/butadiene copolymers having pendant vinyl groups. This tendency is, in general, pronounced when steric regulation becomes significant. From our preliminary results obtained through the screening study, C2-symmetric zirconocene, rac-[Me2Si(2-methyl-4-phenylindenyl)2]ZrCl2 [10,11] was an especially efficient catalyst to incorporate butadiene by 1,2-insertion to obtain the desired polymer structure of an ethylene/butadiene copolymer.
Interestingly, this phenomenon is supported by molecular modeling calculations for an analogous C2-symmetric zirconocene, rac-[H2C(3-R-1-indenyl)2]ZrCl2 catalyst system (R = H, Me, Et, iPr, tBu) in that η2 primary coordination of butadiene is energetically favorable compared to η2 secondary coordination and the η4-cis coordination of butadiene [8]. In addition, the relative amount of constitutional monomer units derived from butadiene are affected by the bulkiness of the substituent. To that end, we have anticipated that tuning the steric property of the rac-[R2Si(Indenyl’)2]ZrCl2 (Indenyl’ = generic substituted indenyl) would lead to a further increase in the pendant vinyl group while keeping the internal olefins at a negligible level. Herein we describe studies on a series of rac-[SiMe2(Indenyl’)2]ZrCl2 catalysts with diverse steric properties to clarify factors that control the insertion mode of butadiene, and investigated methods to prepare copolymers that contain predominant pendant vinyl groups and are substantially free of internal olefins (derived from 1,4- or 1,3-addition units).

2. Results and Discussion

2.1. Ethylene/Butadiene Copolymerization by rac-[SiMe2(Indenyl’)2]ZrCl2

Influence of Substituent on Insertion Mode of Butadiene in Ethylene/Butadiene Copolymerization

We evaluated eight C2-symmetric zirconocene complexes, rac-[SiMe2(Indenyl’)2]ZrCl2 (Figure 1); rac-dimethylsilylbis(1-indenyl)zirconium dichloride (1), rac-dimethylsilyl-bis[1-(2-methyl-4,5 -benzoindenyl)] zirconium dichloride (2), rac-dimethylsilyl-bis[l-(2-methyl-4-phenylindenyl)] zirconium dichloride (3), rac-dimethy1si1y1-bis(2-ethyl-4-phenylindenyl)zirconium dichloride (4), rac-dimethylsilyl-bis[l-(2-n-propyl-4-(1-naphthyl)indenyl)]zirconium dichloride (5), rac-dimethylsilyl -[1-(2-ethyl-4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta[a]naphthalenyl)indenyl))][1-(2-n-propyl -4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta[a]naphthalenyl)indenyl))]zirconium dichloride (6), rac-dimethylsilyl-bis[1-(2-ethyl-4-(9-phenanthryl)indenyl)]zirconium dichloride (7), and rac-dimethylsilyl-bis[l-(2-n-propyl-4-(9-phenanthryl)indenyl)]zirconium dichloride (8) for ethylene or propylene/butadiene copolymerization in the presence of hydrogen under identical conditions. The resultant copolymers were analyzed by 13C-NMR, and the butadiene content and relative abundance of various enchained monomer units were determined. The copolymers were also analyzed by 1H-NMR to determine the quantity of the double bonds (vinyl groups and internal olefins) more accurately, because by 13C-NMR analysis, the polymer is exposed to harsh conditions such as 140 °C for 12 h, so that the consumption of double bonds, especially the consumption of internal olefins during an analysis is anticipated.
Figure 1. Ansa-zirconocene catalysts employed in this study.
Figure 1. Ansa-zirconocene catalysts employed in this study.
Catalysts 05 02001 g001
Ethylene/butadiene copolymerization results are summarized in Table 1. In the case of complex 1 /iBu3Al/Ph3CB(C6F5)4, although activity was significantly high, butadiene was hardly incorporated into the resultant polymer, and the majority of butadiene units formed a five-membered ring (Table 1, run 1; Figure 2). However, a strong correlation between the microstructure of the copolymer and the bulkiness of the substituent on rac-[SiMe2(Indenyl’)2]ZrCl2 was observed; as bulkiness of the substituent increases, a relative abundance of rings relative to all the structural units derived from butadiene decreased and conversely, that of the vinyl groups increased (Table 1, run 1–8; Figure 2). Surprisingly, complex 8, which has the most bulky substituent generated 18 times as much vinyl groups compared to uncrowded complex 1. These results suggest that, steric hindrance on rac-[SiMe2(Indenyl’)2]ZrCl2 is effective to generate pendant vinyl groups by hampering the formation of sterically demanding cyclopropane and the cyclopentane rings. The amount of internal olefins was negligible for all cases regardless of the type of substituent, even according to sensitive 1H-NMR analysis (Figure 3 and Figure 4).
Figure 2. Butadiene content and relative abundance of various enchained monomer units (mol % calculated by 13C-NMR) in ethylene/butadiene copolymerization by complexes 18/iBu3Al/Ph3CB(C6F5)4. Polymerization conditions refer to entries 18 in Table 1.
Figure 2. Butadiene content and relative abundance of various enchained monomer units (mol % calculated by 13C-NMR) in ethylene/butadiene copolymerization by complexes 18/iBu3Al/Ph3CB(C6F5)4. Polymerization conditions refer to entries 18 in Table 1.
Catalysts 05 02001 g002
Table 1. Ethylene/butadiene copolymerization results for complexes 18 a.
Table 1. Ethylene/butadiene copolymerization results for complexes 18 a.
EntryCatA bMw c (103)Mw/MnBD in Polym. d (mol %)Mole Fraction eQuantity per 1000 Carbon fTg (°C)Tm (°C)
1,2-addition unitsInternal olefins gVinyl groups hInternal olefins g
fv/fcycPro/fcycPen
11371.291.92.0211/0/89010.2-113
2244.9156.63.245/29/66010.2-100
3314.189.22.4728/29/43050.1−3371/117
446.237.72.4944/21/35070.3−3952/114
556.871.12.0846/21/330110.3−3255
665.350.52.1953/15/320130.3−3345
772.477.93.8947/26/280150.1−3252/113
883.677.92.91261/25/140180.3−3333/105
9 i84.4114.02.81049/27/230180.1−3137
10 j84.5132.52.91272/14/140410.0−23-
11 k81.935.73.11079/5/160490.2−14-
12 l81.339.33.91695/0/50761.3−10-
a Conditions: hexane 250 mL, butadiene 25 g, ethylene 0.51 MPa (partial pressure at 60 °C), H2 372 mL (volume under atmospheric pressure at 20 °C), Zr 0.002 mmol, iBu3Al 0.5 mmol, Ph3CB(C6F5)4 0.012 mmol, polymerization time 20 min, 60 °C, total pressure 0.78 MPa; b Activity based on polymer yield; unit: kg-polymer/(mmol-Zr hr); c Weight average molecular weight determined by GPC using polystyrene calibration; d Butadiene content in the copolymer determined by 13C-NMR; e Relative abundance of various enchained monomer units relative to all the structural units derived from butadiene determined by 13C-NMR; f determined by 1H-NMR; g The total amount of f1,4 and f1,3; h The total amount of pendant vinyl groups and terminal vinyl groups; i hexane 1000 mL, butadiene 100 g, ethylene 0.55 MPa (partial pressure at 60 °C), H2 620 mL (volume under atmospheric pressure at 20 °C), Zr 0.004 mmol, iBu3Al 2.0 mmol, Ph3CB(C6F5)4 0.024 mmol; j hexane 1000 mL, butadiene 100 g, ethylene 0.55 MPa (partial pressure at 60 °C), H2 620 mL (volume under atmospheric pressure at 20 °C), Zr 0.004 mmol, DMAO 11.2 mmol as Al; k hexane 1000 mL, butadiene 300 g, ethylene 0.42 MPa (partial pressure at 60 °C), H2 620 mL (volume under atmospheric pressure at 20 °C), Zr 0.004 mmol, DMAO 11.2 mmol as Al; l hexane 2000 mL, butadiene 190 g, ethylene 0.17 MPa (partial pressure at 70 °C), H2 990 mL (volume under atmospheric pressure at 20 °C), Zr 0.008 mmol, DMAO 22.4 mmol as Al, 70 °C.
Figure 3. Quantity of vinyl groups and internal olefins (per 1000 carbon calculated by 1H-NMR) in ethylene/butadiene copolymerization by complexes 18/iBu3Al/Ph3CB(C6F5)4. Polymerization conditions refer to entries 18 in Table 1.
Figure 3. Quantity of vinyl groups and internal olefins (per 1000 carbon calculated by 1H-NMR) in ethylene/butadiene copolymerization by complexes 18/iBu3Al/Ph3CB(C6F5)4. Polymerization conditions refer to entries 18 in Table 1.
Catalysts 05 02001 g003
Figure 4. 1H-NMR spectra of the ethylene/butadiene copolymers obtained from entries 1, 8, and 12 of Table 1.
Figure 4. 1H-NMR spectra of the ethylene/butadiene copolymers obtained from entries 1, 8, and 12 of Table 1.
Catalysts 05 02001 g004
Interestingly, DMAO is more efficient than iBu3Al/Ph3CB(C6F5)4 for generating vinyl groups (Table 1, entries 9 and 10). As part of an effort to increase the number of vinyl groups, the butadiene/ethylene ratio in the feed was increased at DMAO activation conditions. As a result, we succeeded in synthesizing ethylene/butadiene copolymers which are dominated by vinyl groups (fv = 79 mol %; Table 1, entry 11). Further increase in butadiene/ethylene feed ratio [13] led to formation of ethylene/butadiene copolymers whose structural units derived from butadiene are mostly vinyl groups (fv = 95 mol %; Table 1, entry 12). Notably, the quantity of vinyl groups per 1000 carbon reached 76 for this condition (Table 1, entry 12; Figure 4). Although the formation of internal olefins accompanied this condition (1.3 per 1000 Carbon; Table 1, entry 12), it is a quantity far too low compared to that of the vinyl groups. The remaining internal olefins will be eliminated by introducing a higher amount of hydrogen [10,11].

2.2. Propylene/Butadiene Copolymerization by Complexes 3 and 8

Following the ethylene/butadiene copolymerization, propylene/butadiene copolymerization was conducted using complexes 3 and 8 combined in activation with MMAO so as to compare with the previous study [10,11]. As with the results obtained by the ethylene/butadiene copolymerization, a similar substituent effect on the vinyl content was observed. Therefore, in this case again, we succeeded in synthesizing a propylene/butadiene copolymer possessing predominant vinyl groups (Table 2; Figure 5; Figure 6). Notably, the vinyl content obtained by complex 8 overwhelmed that by the previously reported complex 3 [10,11].
Table 2. Propylene/butadiene copolymerization results using complexes 18 a.
Table 2. Propylene/butadiene copolymerization results using complexes 18 a.
EntryCatA bMw c (103)Mw /MnBD in Polym. d (mol %)Mole Fraction eQuantity per 1000 Carbon fTg (°C)Tm (°C)
1,2-addition unitsInternal olefins gfH-1,4Vinyl groups hInternal olefins g
fv/fcycPro/fcycPen
1316.699.62.4538/2/28032606122
2825.4257.03.0960/5/25010170−15.3/3.9102
3 i80.737.62.63758/13/2603863.7–24-
a Conditions: hexane 500 mL, butadiene 20 g, propylene 0.63 MPa (partial pressure at 60 °C), H2 124 mL (volume under atmospheric pressure at 20 °C), Zr 0.002 mmol, MMAO 5.6 mmol as Al, Polymerization time 20 min, 60 °C, total pressure 0.78 MPa; b Activity based on polymer yield; unit: kg-polymer/(mmol-Zr hr); c Weight average molecular weight determined by GPC using polystyrene calibration; d Butadiene content in the copolymer determined by 13C-NMR; e Relative abundance of various enchained monomer units relative to all the structural units derived from butadiene determined by 13C-NMR; f determined by 1H-NMR; g The total amount of f1,4 and f1,3; h The total amount of pendant vinyl groups and terminal vinyl groups; i butadiene 200 g, propylene 0.26 MPa (partial pressure at 60 °C).
Likewise, increasing the butadiene concentration in the feed led to the formation of propylene/butadiene copolymers abundant with vinyl groups (quantity of vinyl groups = 86 per 1000 carbon; Table 2, entry 3; Figure 5). Although an extremely high butadiene/propylene feed ratio entails the formation of a small amount of internal olefins (quantity of internal olefins = 3.7 per 1000 carbon; Table 2, entry 3; Figure 5), the quantity is still far too low compared to that of the vinyl group. The remaining internal olefins will be eliminated by introducing a higher amount of hydrogen, as mentioned above [10,11].
Figure 5. 1H-NMR spectra of the propylene/butadiene copolymers obtained from the runs 1, 2, and 3 of Table 2.
Figure 5. 1H-NMR spectra of the propylene/butadiene copolymers obtained from the runs 1, 2, and 3 of Table 2.
Catalysts 05 02001 g005
Figure 6. 13C-NMR spectra of the propylene/butadiene copolymers obtained from the runs 1, 2, and 3 of Table 2.
Figure 6. 13C-NMR spectra of the propylene/butadiene copolymers obtained from the runs 1, 2, and 3 of Table 2.
Catalysts 05 02001 g006

3. Experimental Section

3.1. General

All manipulations were performed using dry box techniques under a purified N2 atmosphere or Schlenk techniques under N2 atmosphere, or on a high-vacuum line, unless otherwise indicated.

3.2. Materials

Hexane employed as a polymerization solvent (Kanto Chemical Co., Inc., Tokyo, Japan) was dried over Al2O3 and degassed by the bubbling with nitrogen gas. Toluene used for dissolving pre-catalysts was purified by passage through columns of activated alumina and a BASF R3-11 oxygen scavenger (BASF, Ludwigshafenm, Germany). Ethylene and propylene were obtained from Sumitomo Seika Co., Tokyo, Japan. Triisobutylalminium (iBu3Al) and Methyl-isobutylalumoxane (modified methylalumoxane, MMAO) were purchased from Tosoh Finechem Corp., Yamaguchi, Japan and used as received. Dried methylaluminoxane (DMAO) was prepared by concentrating MAO (1.2 M in toluene, Albemarle, Baton Rouge, LA, USA) in vacuo to remove the remaining trimethylaluminum, providing a solid white powder; the white powder was dissolved in dried toluene prior to use. Triphenylcarbenium tetrakis(pentafluorophenyl)borate ([Ph3C]+[B(C6F5)4]) was purchased from Asahi Glass Corp., Tokyo, Japan and used as received. rac-dimethylsilylbis(1-indenyl)zirconium dichloride (1) was purchased from Strem Chemicals Inc., Newburyport, MA, USA. Ansa-zirconocene catalysts 28 employed in this study (Figure 1) were prepared according to the methods described in our previous papers: rac-dimethylsilyl -bis[1-(2-methyl-4,5-benzoindenyl)]zirconium dichloride (2) [14], rac-dimethylsilyl-bis[l-(2-methyl -4-phenylindenyl)]zirconium dichloride (3) [14], rac-diniethy1si1y1-bis(2-ethyl-4-phenylindenyl) zirconium dichloride (4) [15], rac-dimethylsilyl-bis[l-(2-n-propyl-4-(1-naphthyl)indenyl)]zirconium dichloride (5) [15], rac-dimethylsilyl-[1-(2-ethyl-4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta[a] naphthalenyl)indenyl))][1-(2-n-propyl-4-(5-(2,2-dimethyl-2,3-dihydro-1H-cyclopenta[a]naphthalenyl) indenyl))]zirconium dichloride (6) [16], rac-dimethylsilyl-bis[1-(2-ethyl-4-(9-phenanthryl)indenyl)] zirconium dichloride (7) [15], rac-dimethylsilyl-bis[l-(2-n-propyl-4-(9-phenanthryl)indenyl)]zirconium dichloride (8) [15]. [η1: η5-tert-butyl(dimethylfluorenylsilyl)-amido]dimethyltitanium ([t-BuNSiMe2Flu] TiMe2) was synthesized according to the methods described in the literature [17].

3.3. Polymerization Procedure

3.3.1. Ethylene/Butadiene Copolymerization (Table 1)

A prescribed amount of hexane was introduced into a SUS autoclave (1000 mL; entry 1–8, 4000 mL; entry 9–12) equipped with two propeller-like stirrers under nitrogen at an atmospheric pressure of 25 °C. A toluene solution of DMAO or iBu3Al was loaded. Then a prescribed amount of butadiene was charged. Subsequently, the mixture was heated to a prescribed polymerization temperature, and a prescribed amount of hydrogen (the volume under atmospheric pressure at 20 °C) was added, then ethylene was introduced into the reactor up to the polymerization pressure (0.78MPa). For runs in which iBu3Al/Ph3CB(C6F5)4 was employed as a co-catalyst, Zr complex/toluene, iPh3CB(C6F5)4/toluene, and toluene wash (2 mL) were sequentially injected into the reactor to start the reaction. For runs in which the DMAO was employed as a co-catalyst, Zr complex/toluene and toluene wash (2 mL) were injected sequentially into the reactor to start the reaction. The pressure was kept constant at 0.78 MPa by feeding ethylene on demand. After the reaction was performed at a prescribed polymerization temperature for 20 min, the polymerization was terminated by adding a small amount of methanol. The resulting mixture was poured into a large excess of methanol containing hydrochloric acid to precipitate the polymer. The polymer was separated by filtration, and admixed with a stabilizing agent, 0.1 wt % of antioxidant (SUMILIZER® GS(F): 2-[1-(2-Hydroxy-3,5-di-tert-pentylphenyl)ethyl]-4,6-di-tert-pentylphenyl acrylate, sumitomo chemical co. ltd, Tokyo, Japan), then dried under reduced pressure at 20 °C for 24 h.

3.3.2. Propylene/Butadiene Copolymerization (Table 2)

A prescribed amount of hexane was introduced into a SUS autoclave (1000 mL) equipped with two propeller-like stirrers under nitrogen at an atmospheric pressure of 25 °C. A toluene solution of MMAO was loaded, then a prescribed amount of butadiene was charged. Subsequently, the mixture was heated to a prescribed polymerization temperature, and a prescribed amount of hydrogen (the volume under atmospheric pressure at 20 °C) was added. Then, propylene was introduced into the reactor up to the polymerization pressure (0.78 MPa). Zr complex/toluene and toluene wash (2 mL) were injected sequentially into the reactor to start the reaction. The pressure was kept constant at 0.78 MPa by feeding propylene on demand. After the reaction was performed at 60 °C for 20 min, the polymerization was terminated by adding a small amount of methanol. The work-up procedure for the propylene/butadiene copolymer was identical to that for the ethylene/butadiene copolymerization.

3.4. Polymer Characterization

3.4.1. Microstructure Analysis of Ethylene/Butadiene Copolymer by 13C-NMR

13C-NMR spectra of the ethylene/butadiene copolymers were recorded on an ECA500 spectrometer (125 MHz) from Japan Electron Optics Laboratory Co. Ltd. (JEOL, Tokyo, Japan), using 1,1,2,2-tetrachloroethane-d2 as a solvent at 140 °C. Chemical shifts were referenced to the residual solvent peak (δ13C = 72.4 ppm). The spectrum was analyzed with reference to the chemical shifts of the signals assigned to the vinyl groups, the cyclopropane skeleton, the cyclopentane skeleton, the 1,4-addition units and the 1,3-addition units in the ethylene/butadiene copolymers which were described in previous papers [8,18]. The determination of the butadiene content (1) and the vinyl groups content (2); the cyclopropane skeleton content (3); the cyclopentane skeleton content (4); the 1,4-addition units content (5) and the 1,3-addition units content (6) relative to all the butadiene-derived structural units was performed in accordance with the following Equations (1) to (6).
Proportion of structural units derived from butadiene relative to all the monomer units in the copolymer is given by Equation (1).
BD content (mol %) = 100  ×  (A + B + C + D + E)/{(A + B + C + D + E) + [(1000 − (A + B + C + D + E)  ×  4)]/2}
Fraction of vinyl groups (fv), cyclopropane rings (fcycPro), cyclopentane rings (fcycPen), 1,4-addition units (f1,4), and 1,3-addition units (f1,3) relative to all the structural units derived from butadiene is given by Equations (2)–(6), respectively.
f v (mol %) = 100 × A/(A + B + C + D + E)
f cycPro (mol %) = 100 × B/(A + B + C + D + E)
f cycPen (mol %) = 100 × C/(A + B + C + D + E)
f 1 , 4 (mol %) = 100 × D/(A + B + C + D + E)
f 1 , 3 (mol %) = 100 × E/(A + B + C + D + E)
The letters A to E indicate the respective integrated values per one carbon of the structural units derived from the structures illustrated below, based on the total integrated value, 1000, of all the peaks in the 13C-NMR spectrum of the ethylene/butadiene copolymers.
A:
(Integrated value per one carbon of the vinyl groups) = integrated value of the peak at 113.8 ppm ( R 1 CH ( HC = C _ H 2 ) R 2 )
B:
(Integrated value per one carbon of the cyclopropane skeleton) = [(integrated value of the peak at 17.2 ppm (cis CH in Chemical Formula (α))) + (integrated value of the peak at 19.2 ppm (trance CH in Chemical Formula (α)))]/2
Catalysts 05 02001 i001
C:
(Integrated value per one carbon of the cyclopentane skeleton) = [(integrated value of the peak at 43.1 ppm (cis CH in Chemical Formula (β))) + (integrated value of the peak at 46.3 ppm (trans CH in Chemical Formula (β)))]/2
Catalysts 05 02001 i002
D:
(Integrated value per one carbon of the 1,4-addition units) = [integrated value of the peaks at 131–130 ppm ( R 1 H C _ = C _ H R 2 ) ] / 2
E:
(Integrated value per one carbon of the 1,3-addition units) = [integrated value of the peak at 137.3 ppm ( R 1 HC = C _ H CH ( Me ) R 2 ) ]

3.4.2. Microstructure Analysis of Ethylene/Butadiene Copolymers by 1H-NMR

1H-NMR spectra of the ethylene/butadiene copolymers were recorded on a JEOL270 spectrometer (270 MHz) from Japan Electron Optics Laboratory Co. Ltd. (JEOL, Tokyo, Japan), using o-dichlorobenzene-d4 as a solvent at 120 °C. Chemical shifts were referenced to the residual solvent peak (δ1H = 7.15 ppm). Quantity of the vinyl groups Equation (7) and the internal olefins Equation (8) in ethylene/butadiene copolymers were quantified as per 1000 Carbon in accordance with the following Equations (7) and (8), respectively.
Quantity of vinyl groups per 1000 Carbon  = F + G
Quantity of internal olefins per 1000 Carbon  = H
The letters F to H indicate the respective integrated values per one proton of the structural units derived from the structures illustrated below, based on the total integrated value, 2000, of all the peaks in the 1H-NMR spectrum of the ethylene/butadiene copolymer.
F:
(Integrated value per one proton of the pendant vinyl groups) = integrated value of the peak at 5.4–5.6 ppm ( R 1 CH ( H _ C = C H 2 ) R 2 )
G:
(Integrated value per one proton of the terminal vinyl groups) = integrated value of the peak at 5.6–5.8 ppm ( R 1 ( H _ C = C H 2 ) )
H:
(Integrated value per one proton of the internal olefins) = [integrated value of the peaks at 5.2–5.4 ppm ( R 1 H _ C = C H _ R 2 ) ] / 2

3.4.3. Microstructure Analysis of Propylene/Butadiene Copolymer by 13C-NMR

13C-NMR spectra of the propylene/butadiene copolymers were recorded on an ECA500 spectrometer (125 MHz) from Japan Electron Optics Laboratory Co. Ltd. (JEOL, Tokyo, Japan), using 1,1,2,2-tetrachloroethane-d2 as a solvent at 140 °C. Chemical shifts were referenced to the residual solvent peak (δ13C = 72.4 ppm). The spectrum was analyzed with reference to the chemical shifts of the signals assigned to the vinyl groups, the cyclopropane skeleton, the methylcyclopentane skeleton [19], the 1,4-addition units and the 1,3-addition units in the propylene/butadiene copolymers which were described in previous papers [8,10,11]. The determination of the butadiene content (Equation (9)) and the vinyl groups content (Equation (10)), the cyclopropane skeleton content (Equation (11)), the methylcyclopentane skeleton content (Equation (12)), the 1,4-addition units content (Equation (13)), the 1,3-addition units content (Equation (14)), and the hydrogenated 1,4-addition units content (Equation (15)) relative to all the butadiene-derived structural units was performed in accordance with the following Equations (9) to (15).
Proportion of structural units derived from butadiene relative to all the monomer units in the copolymer is given by Equation (9).
BD content (mol %)  = 100   ×  (I + J + K + L + M + N)/{(I + J + K + L + M + N) + [(1000–(I + J + K + L + M + N) × 4)]/3}
Fraction of vinyl groups (fv), cyclopropane rings (fcycPro), cyclopentene rings (fcycPen), 1,4-addition units (f1,4), 1,3-addition units (f1,3), and hydrogenated 1,4-addition units (fH-1,4) relative to all the structural units derived from butadiene is given by Equations (10)–(15), respectively.
f v (mol %) = l00 × I/(I + J + K + L + M + N) 
f cycPro (mol %) = l00 × J/(I + J + K + L + M + N) 
f MecycPen (mol %) = l00 × K/(I + J + K + L + M + N) 
f 1 , 4 (mol %) = l00 × L/(I + J + K + L + M + N) 
f 1 , 3 (mol %) = l00 × M/(I + J + K + L + M + N) 
f H - 1 , 4 (mol %) = l00 × N/(I + J + K + L + M + N) 
The letters I to N indicate the respective integrated values per one carbon of the structural units derived from the structures illustrated below, based on the total integrated value, 1000, of all the peaks in the 13C-NMR spectrum of the propylene/butadiene copolymers.
I:
(Integrated value per one carbon of the vinyl groups) = integrated value of the peak at 113.2 ppm ( R 1 CH ( HC = C _ H 2 ) R 2 )
J:
(Integrated value per one carbon of the cyclopropane skeleton) = [(integrated value of the peak at 12.9 ppm (CH2 in Chemical Formula (α'))
Catalysts 05 02001 i003
K:
(Integrated value per one carbon of the methylcyclopentane skeleton) = integrated value of the peak at 32.5 ppm (CH in Chemical Formula (β'))
Catalysts 05 02001 i004
L:
(Integrated value per one carbon of the 1,4-addition units) = [integrated value of the peaks at 128–133 ppm ( R 1 H C _ = C _ H R 2 ) ] / 2
M:
(Integrated value per one carbon of the 1,3-addition units) = [integrated value of the peak at 134–136 ppm ( R 1 HC = C _ H CH ( Me ) R 2 ) ]
N:
(Integrated value per one carbon of the hydrogenated 1,4-addition units) = [integrated value of the peaks at 37.1 ppm (CH2 in Chemical Formula (γ'))]/2
Catalysts 05 02001 i005

3.4.4. Microstructure Analysis of Propylene/Butadiene Copolymer by 1H-NMR

1H-NMR spectra of the propylene/butadiene copolymers were recorded on a JEOL270 spectrometer (270 MHz) from Japan Electron Optics Laboratory Co. Ltd. (JEOL, Tokyo, Japan), using o-dichlorobenzene-d4 as a solvent at 120 °C. Chemical shifts were referenced to the residual solvent peak (δ1H = 7.15 ppm). The number of vinyl groups (Equation (16)) and the internal olefins (Equation (17)) in propylene/butadiene copolymers were quantified as per 1000 Carbon in accordance with the following Equations (16) and (17), respectively.
Quantity of vinyl groups per 1000 Carbon = O
Quantity of internal olefins per 1000 Carbon = P
The letters O and P indicate the respective integrated values per one proton of the structural units derived from the structures illustrated below, based on the total integrated value, 2000, of all the peaks in the 1H-NMR spectrum of the propylene/butadiene copolymers.
O:
(Integrated value per one proton of the vinyl groups) = [integrated value of the peak at 4.7–5.0 ppm ( R 1 CH ( HC = C H _ 2 ) R 2 ) + integrated value of the peak at 5.4–5.8 ppm ( R 1 CH ( H _ C = C H 2 ) R 2 ) ] / 3
P:
(Integrated value per one proton of the internal olefins) = [integrated value of the peaks at 5.2–5.4 ppm ( R 1 H _ C = C H _ R 2 ) ] / 2

3.4.5. GPC Analysis

Molecular weights (Mw and Mn) and molecular weight distributions (MWDs) of the polymers were determined using a Waters GPC2000 gel permeation chromatograph (Waters, Milford, MA, USA) equipped with four TSKgel columns (two sets of TSKgel GMH6-HT and two sets of TSKgel GMH6-HTL) at 140 °C using PS calibration. o-Dichlorobenzene was employed as a solvent at a flow rate of 1.0 mL/min.

3.4.6. DSC Analysis

The glass transition temperature (Tg) and the melting point (Tm) of the polymers were determined by differential scanning calorimetry (DSC) with a Shimadzu DSC-60 instrument (Shimadzu Corporation, Kyoto, Japan). In a N2 (nitrogen) atmosphere, the polymer was heated from ordinary temperature to 200 °C at a temperature increasing rate of 50 °C/min, held at the temperature for 5 min, cooled to −100 °C at a temperature decreasing rate of 10 °C/min, and held at the temperature for 5 min. The temperature was again increased to 200 °C at a temperature increasing rate of 10 °C/min. The Tg and the Tm were obtained from an endothermic curve recorded during the second scanning.

4. Conclusions

In this study, we investigated the substituent effects of C2-symmetric zirconocene complexes 18, rac-[Me2Si(Indenyl’)2]ZrCl2, after activation, on the structures of the resultant copolymers in the copolymerization of ethylene or propylene with butadiene. As a result, we can reveal a larger steric hindrance on the complex tends to suppress the formation of cyclopropyl and cyclopentyl moieties in the copolymer. Additionally, we have found that the steric hindrance exercises a great influence on the ratio of the pendant vinyls and vinylenes (internal olefins), and complex 8 having the largest steric hindrance provides copolymers having predominant pendant vinyl groups with virtually no internal olefins. These results indicate that the steric hindrance on the zirconocene complex employed can control the insertion mode of butadiene and, moreover, mitigate the cyclization process leading to the generation of cyclopropyl and cyclopentyl moieties. The copolymers can be transformed into functionalized copolymers using procedures from the literature. The functionalized copolymers we obtained will have a wide variety of applications covering various fields.

Acknowledgments

We thank Mitsui Chemical Analysis & Consulting Service Inc., for GPC and NMR measurements. We thank A. Valentine for fruitful discussions and suggestions.

Author Contributions

Terunori Fujita, Makoto Mitani and Kenji Michiue conceived and designed the experiments; Kenji Michiue performed the experiments; Makoto Mitani and Kenji Michiue analyzed the data; Kenji Michiue wrote the paper; All the authors discussed the results and improved the final text of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Franssen, N.M.G.; Reek, J.N.H.; de Bruin, B. Synthesis of functional “polyolefins”: State of the art and remaining challenges. Chem. Soc. Rev. 2013, 42, 5809–5832. [Google Scholar] [CrossRef] [PubMed]
  2. Welborn, H.C., Jr. Copolymers of Ethylene and 1,3-butadiene. EP0275676, 27 July 1988. [Google Scholar]
  3. Galimberti, M.; Albizzati, E.; Abis, L.; Bacchilega, G. 13C-NMR analysis of α-olefins copolymers with 1,3-butadiene obtained with zirconocenes/methylalumoxane catalysts. Makromol. Chem. 1991, 192, 2591–2601. [Google Scholar] [CrossRef]
  4. Morizono, K.; Okada, K.; Yamaguchi, M. Unsaturated Copolymers, Processes for Preparing the Same, and Compositions Containing the Same. US6310164, 30 October 2001. [Google Scholar]
  5. Morizono, K.; Okada, K.; Yamaguchi, M. Unsaturated Copolymer, Preparation Thereof and Composition Containing Said Copolymer. Patent JPH1180269, 26 March 1999. [Google Scholar]
  6. Pragliola, S.; Milano, G.; Guerra, G.; Longo, P. Stereoselective Cyclopropanation by Cyclocopolymerization of Butadiene. J. Am. Chem. Soc. 2002, 124, 3502–3503. [Google Scholar] [CrossRef] [PubMed]
  7. Longo, P.; Pragliola, S.; Milano, G.; Guerra, G. E Stereoregular 1,1 and 1,3 Constitutional Units from 1,3-Butadiene in Copolymerizations Catalyzed by a Highly Hindered C2 Symmetric Metallocene. J. Am. Chem. Soc. 2003, 125, 4799–4803. [Google Scholar] [CrossRef] [PubMed]
  8. Longo, P.; Napoli, M.; Pragliola, S.; Costabile, C.; Milano, G.; Guerra, G. Butadiene Insertion and Constitutional Units in Ethene Copolymerizations by C2-Symmetric Metallocenes. Macromolecules 2003, 36, 9067–9074. [Google Scholar] [CrossRef]
  9. Pragliola, S.; Costabile, C.; Magrino, M.; Napoli, M.; Longo, P. Ethene/1,3-Butadiene Copolymerization in the Presence of rac-(CH2-(3-tert-butyl-1-indenyl)2)ZrCl2/MAO Catalytic System: Study of the Polymerization Mechanism by Using 13C-Labeled 1,3-Butadiene. Macromolecules 2004, 37, 238–240. [Google Scholar] [CrossRef]
  10. Ishihara, T.; Shiono, T. Hydrogenated 1,4-Insertion of Butadiene in the Copolymerization with Propylene Using an Isospecific Zirconocene Catalyst. J. Am. Chem. Soc. 2005, 127, 5774–5775. [Google Scholar] [CrossRef] [PubMed]
  11. Ishihara, T.; Ban, H.T.; Hagihara, H.; Shiono, T. Additiveeffectsofalkylaluminiumcompoundsonpr opylene-1,3-butadiene copolymerization using isospecific zirconocenecatalysts. J. Organomet. Chem. 2010, 695, 1694–1699. [Google Scholar] [CrossRef]
  12. Michiue, K.; Ishii, S.; Mitani, M.; Karino, T. Copolymer of Olefin and Conjugated Diene, and Process for Producing Same. Patent WO2010113975, 7 October 2010. [Google Scholar]
  13. Partial pressure of butadiene was increased by increasing polymerization temperature from 60 °C to 70 °C.
  14. Spaleck, W.; Kuber, F.; Winter, A.; Rohrmann, J.; Bachmann, B.; Antberg, M.; Dolle, V.; Paulus, E.F. The Influence of Aromatic Substituents on the Polymerization Behavior of Bridged Zirconocene Catalysts. Organometallics 1994, 13, 954–963. [Google Scholar] [CrossRef]
  15. Fukuoka, O.; Tashiro, K.; Kawai, K.; Saito, J.; Imuda, J.; Fujita, T.; Nitahara, M.; Ueda, T.; Kiso, Y.; Yoshida, M. New Transition Metal Compound, Olefin Polymerization Catalyst Component Comprising the Same, Olefin Polymerization Catalyst Containing this Component and Polymerization of Olefin. Patent JPH07286005, 31 October 1995. [Google Scholar]
  16. Fukuoka, O.; Kawai, K.; Nitahara, M. Catalyst for Olefin Polymerization and Polymerization Method of Olefin. Patent JPH1087716, 7 April 1998. [Google Scholar]
  17. Hagihara, H.; Shiono, T.; Ikeda, T. Living Polymerization of Propene and 1-Hexene with the [t-BuNSiMe2Flu]TiMe2/B(C6F5)3 Catalyst. Macromolecules 1998, 31, 3184–3188. [Google Scholar] [CrossRef]
  18. Choo, T.N.; Waymouth, R.M. The Dual-Site Alternating Cyclocopolymerization of 1,3-Butadiene with Ethylene. J. Am. Chem. Soc. 2003, 125, 8970–8971. [Google Scholar] [CrossRef] [PubMed]
  19. The assignment of signals in 13C-NMR spectra for the methylcyclopentane skeleton unit in the propylene/butadiene copolymer was made together with the results by Distortion-less Enhancement by Polarization Transfer (DEPT) NMR experiments.

Share and Cite

MDPI and ACS Style

Michiue, K.; Mitani, M.; Fujita, T. Synthesis of Ethylene or Propylene/1,3-Butadiene Copolymers Possessing Pendant Vinyl Groups with Virtually No Internal Olefins. Catalysts 2015, 5, 2001-2017. https://doi.org/10.3390/catal5042001

AMA Style

Michiue K, Mitani M, Fujita T. Synthesis of Ethylene or Propylene/1,3-Butadiene Copolymers Possessing Pendant Vinyl Groups with Virtually No Internal Olefins. Catalysts. 2015; 5(4):2001-2017. https://doi.org/10.3390/catal5042001

Chicago/Turabian Style

Michiue, Kenji, Makoto Mitani, and Terunori Fujita. 2015. "Synthesis of Ethylene or Propylene/1,3-Butadiene Copolymers Possessing Pendant Vinyl Groups with Virtually No Internal Olefins" Catalysts 5, no. 4: 2001-2017. https://doi.org/10.3390/catal5042001

Article Metrics

Back to TopTop