Next Article in Journal
Development of PVDF Membrane Nanocomposites via Various Functionalization Approaches for Environmental Applications
Next Article in Special Issue
Investigations of the Ligand Electronic Effects on α-Diimine Nickel(II) Catalyzed Ethylene Polymerization
Previous Article in Journal
Seaweed Polysaccharide-Based Nanoparticles: Preparation and Applications for Drug Delivery
Previous Article in Special Issue
Immortal Ring-Opening Polymerization of rac-Lactide Using Polymeric Alcohol as Initiator to Prepare Graft Copolymer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Active and Isospecific Styrene Polymerization Catalyzed by Zirconium Complexes Bearing Aryl-substituted [OSSO]-Type Bis(phenolate) Ligands

Department of Chemistry, Graduate School of Science and Engineering, Saitama University, 255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan
*
Authors to whom correspondence should be addressed.
Polymers 2016, 8(2), 31; https://doi.org/10.3390/polym8020031
Submission received: 17 December 2015 / Revised: 13 January 2016 / Accepted: 20 January 2016 / Published: 26 January 2016
(This article belongs to the Special Issue Metal-Mediated Polymer Synthesis)

Abstract

:
[OSSO]-type dibenzyl zirconium(IV) complexes 9 and 10 possessing aryl substituents ortho to the phenoxide moieties (ortho substituents, phenyl and 2,6-dimethylphenyl (Dmp)) were synthesized and characterized. Upon activation with dMAO (dried methylaluminoxane), complex 9 was found to promote highly isospecific styrene polymerizations ([mm] = 97.5%–99%) with high molecular weights Mw up to 181,000 g·mmol−1. When the Dmp-substituted pre-catalyst 10/dMAO system was used, the highest activity, over 7700 g·mmol(10)−1·h−1, was recorded involving the formation of precisely isospecific polystyrenes of [mm] more than 99%.

Graphical Abstract

1. Introduction

Stereo-controlled polystyrene such as syndiotactic (sPS) and isotactic polystyrenes (iPS) is one of the most versatile polymeric materials, owing to its high melting point, high crystallinity, and excellent resistance to heat and chemicals [1,2]. Since the syndiospecific polymerization of styrene catalyzed by homogeneous CpTiCl3/MAO system was discovered by Ishihara and coworkers at Idemitsu Kosan (Tokyo, Japan) [3,4], numerous efforts have been devoted to employ efficient metal-based catalysts for the production of sPS [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19]. In sharp contrast, isotactic-enriched polystyrenes are still prepared using heterogeneous catalysis [20,21,22,23,24,25,26] or anionic polymerization [27,28,29,30]. While many homogeneous nickel catalysts for the synthesis of iPS have been reported [31,32,33], there are only a few examples of metallocene and post-metallocene catalysts that produce completely isotactic polystyrene [34,35]. Okuda et al. reported that MAO-activated titanium complex 1a supported by 1,4-dithiabutane-bridged [OSSO]-type bis(phenolate) ligand catalyzes the styrene polymerization with good activity (330 g·mmol(1a)−1·h−1) to yield ultra-high molecular weight isotactic polystyrene (Mw = 5,300,000 g·mol−1) [36,37]. Very recently, they also demonstrated that thermally stable robust zirconium complex 2 possessing a dicumyl-substituted [OSSO]-type bis(phenolate) ligand can efficiently polymerize styrene in living fashion giving highly isotactic poly(styrene) ([mm] > 99%) with quite high activity up to 3158 g·mmol(2)−1·h−1 [38]. Capacchione and Proto et al. also presented the living isospecific polymerization of styrene and 1,3-dienes promoted by using [OSSO]-type titanium complex 1b and MAO to form isotactic-poly(styrene)-block-poly(1,3-diene) copolymes [39,40]. ansa-Bridged bis(indenyl) allyl yttrium and neodymium complexes 3 developed by Carpentier et al. also acted as single-site catalysts for the completely isospecific polymerization of styrene with relatively high activity (3a: 1066–1637 g·mmol(3a)−1·h−1; 3b: 392–1094 g·mmol(3b)−1·h−1) [41,42].
Recently, we have succeeded in the development of an [OSSO]-type bis(phenolate) ligand (4) based on a trans-1,2-cyclooctanediyl platform and the preparation of several early-transition metal and aluminum complexes [43,44,45,46,47,48]. We have also found that zirconium(IV) and hafnium(IV) complexes 5 [49,50] and 6 [51] incorporating ligand 4 with activator could promote precisely isospecific polymerizations of α-olefins such as 1-hexene, 4-methyl-1-pentene, and propylene involving excellent activity (Scheme 1). However, the polymerization of styrene using complexes 5 or 6 and activator did not occur, even at high temperatures, probably due to steric hindrance by bulky tBu groups at the ortho positions on the phenolate moieties in 4. These results encouraged us to develop a new type of [OSSO]-type bis(phenolate) ligands, which have a suitable coordination environment to achieve the isospecific styrene polymerization. Herein, we present the synthesis and structural characterization of [OSSO]-type dibenzyl zirconium(IV) complexes possessing aryl substituents ortho to the phenoxide moieties, as well as their ability to catalyze isospecific styrene polymerization.
Scheme 1. Related post-metallocene complexes 13 for styrene polymerization, [OSSO]-type ligand 4 and its complexes 5 and 6.
Scheme 1. Related post-metallocene complexes 13 for styrene polymerization, [OSSO]-type ligand 4 and its complexes 5 and 6.
Polymers 08 00031 g003

2. Experimental Section

2.1. General

All manipulations of air- and/or moisture-sensitive compounds were performed either using standard Schlenk-line techniques or in UNICO 650F (Ibaraki, Japan) or Glovebox Japan E300 gloveboxes (Tokyo, Japan) under an inert atmosphere of argon. Hexane and toluene were purchased from Kanto Chemical (Tokyo, Japan) and were dried over a potassium mirror prior to use. C6D6 was dried over a potassium mirror, and it was degassed by a freeze–thaw cycle prior to use. 1H and 13C NMR spectra were recorded on a Bruker DPX-400 (400 and 101 MHz, respectively, (Billerica, MA, USA) using CDCl3 or C6D6 as the solvent at room temperature. High-resolution mass spectrometry (HRMS) data were recorded by using a Hitachi-Hitec NanoFrontier eLD (Tokyo, Japan). All melting points were determined on a Mel-Temp capillary tube apparatus (Saitama, Japan) and are uncorrected. The molecular weights (Mw) and molecular weight distributions (Mw/Mn) of the polymers were evaluated by gel permeation chromatography (GPC) on a SCL-10AVP/LC-10ATVP/DGU-14A/ CTO-10ACVP/RID-10A apparatus (Shimadzu Corporation, Kyoto, Japan) using a GPC KF-804L (Shodex Corporation, Tokyo, Japan) column. The analyses were performed at room temperature using THF as the solvent and standard polystyrene as the reference. Differential scanning calorimetry (DSC) analyses were performed on a Seiko Instruments DSC 6200 apparatus (Tokyo, Japan) at a rate of 5 °C/min, under continuous flow of N2 (0.5 mL/min), using aluminum capsules. [OSSO]-type ancillary ligands 7 and 8 [53] and dMAO [54] were prepared by the literature procedures.

2.2. Preparation of Dibenzyl Zirconium(IV) Complex 9

A solution of 7 (298 mg, 0.430 mmol) in toluene (10 mL) was added to a solution of Zr(CH2Ph)4 [55] (196 mg, 0.430 mmol) in toluene (10 mL) at room temperature. The mixture was stirred for 1 h at room temperature, and the solvent was removed under reduced pressure. The residue was washed with hexane and dried in vacuo to give dibenzyl zirconium(IV) complex 9 (346 mg) in 83% yield as yellow crystals. 9: Mp 274–275 °C (dec.).
1H NMR (400 MHz) δ 0.65 (br s, 2H), 0.91 (br s, 2H), 1.07 (br s, 6H), 1.29–1.41 (m, 6H), 1.38 (d, J = 9 Hz, 2H), 2.01 (d, J = 9 Hz, 2H), 2.39 (br s, 2H), 3.15 (d, J = 14 Hz, 2H), 3.35 (d, J = 14 Hz, 2H), 6.50 (d, J = 7 Hz, 4H), 6.88 (d, J = 2 Hz, 2H), 6.96–7.21 (m, 14 H), 7.34–7.40 (m, 8H), 7.54 (d, J = 2 Hz, 2H), 7.67 (d, J = 7 Hz, 4H).
13C{1H} NMR (101 MHz) δ 25.4 (CH2), 26.1 (CH2), 28.7 (CH2), 34.5 (CH2), 48.5 (CH), 59.3 (CH2), 123.1 (C), 123.3 (CH), 127.0 (2CH), 127.6 (CH), 128.6 (CH), 128.7 (CH), 129.1 (2CH), 129.2 (2CH), 129.3 (2CH), 129.9 (2CH), 130.5 (CH), 130.8 (2CH), 132.5 (C), 133.3 (C), 140.1 (C), 141.1 (C), 144.4 (C), 158.2 (C).

2.3. Preparation of Dibenzyl Zirconium(IV) Complex 10

A solution of 8 (380 mg, 0.608 mmol) in toluene (10 mL) was added to a solution of Zr(CH2Ph)4 [55] (277 mg, 0.608 mmol) in toluene (5 mL) at room temperature. The mixture was stirred for 1 h at room temperature, and the solvent was removed under reduced pressure. The residue was washed with hexane (2 mL) and dried to give dibenzyl zirconium(IV) complex 10 (499 mg, 92%) as yellow crystals. 10: Mp 240–241 °C (dec.).
1H NMR (400 MHz, C6D6) δ 0.76 (m, 2H), 0.82 (d, J = 8 Hz, 2H), 1.00 (m, 2H), 1.17–1.30 (m, 6H), 1.45–1.55 (m, 4H), 1.65 (d, J = 8 Hz, 2H), 2.05 (s, 6H), 2.10 (s, 1H), 2.26 (s, 6H), 2.38 (s, 6H), 2.41 (br s, 1H), 3.04 (d, J = 15 Hz, 2H), 3.11 (d, J = 15 Hz, 2H), 6.33 (br s, 2H), 6.47 (d, J = 7 Hz, 4H), 6.71 (br s, 2H), 6.91 (t, J = 7 Hz, 2H), 7.07 (t, J = 7 Hz, 4H), 7.11–7.15 (m, 4H), 7.25 (d, J = 7 Hz, 2H).
13C{1H} NMR (101 MHz, C6D6) δ 20.6 (CH3), 21.3 (CH3), 21.4 (CH3), 21.9 (CH2), 25.4 (CH2), 26.2 (CH2), 34.3 (CH2), 48.1 (CH), 58.0 (CH2), 122.0 (C), 122.8 (CH), 127.6 (CH), 127.8 (CH), 127.9 (CH), 129.2 (2CH), 129.3 (C), 129.6 (2CH), 130.2 (CH), 130.8 (C), 131.4 (CH), 136.2 (C), 137.4 (C), 139.9 (C), 144.6 (C), 156.1 (C).

2.4. General Procedure for Styrene Polymerization

A 50 mL Schlenk-flask was charged sequentially with catalytic precursor 9 or 10 (2.0 μmol), dMAO as an activator (0.50 mmol), and toluene (5 mL) at 25 °C. After stirring for 5 min at the temperature, styrene (3.0 g, 28.8 mmol) was added to the reaction mixture. The mixture was stirred for 60, 10, or 5 min at a desired temperature. The reaction was quenched by addition of methanol and HCl aq. The mixture was extracted with CH2Cl2 and the organic layer was washed with water and dried over anhydrous Na2SO4. The solvent was removed in vacuo at 70 °C during overnight to leave poly(styrene).

2.5. X-ray Crystallographic Analysis

Yellow single crystals of 9 were grown by slow evaporation of its saturated hexane solution at −20 °C. The intensity data were collected at 100 K for 9 on a Bruker SMART APEX II ULTRA (Billerica, MA, USA) equipped with a CCD area detector with graphite-monochromated MoKa radiation (l = 0.71073 Å). The structure was solved by direct methods and refined by full-matrix least-squares procedures on F2 for all reflections (SHELX-97) [56]. Hydrogen atoms of 9 were located by assuming ideal geometry and were included in the structure calculations without further refinement of the parameters. Crystallographic data and details of refinement for 9: C60H56O2S2Zr, 2(C7H8), Mw = 1148.66, orthorhombic, space group P212121, a = 11.8318(11) Å, b = 21.988(2) Å, c = 22.933(2) Å, V = 5966.1(10) Å3, Z = 4, Dcalc = 1.279 g cm−3, R1 (I > 2σI) = 0.0434, wR2 (all data) = 0.1053 for 11081 reflections, 495 restraints, and 833 parameters, GOF = 1.017.

3. Results and Discussion

3.1. Synthesis of Dibenzyl Zirconium(IV) Complexes 9 and 10

According to a previous report [53], new [OSSO]-type ancillary ligands 7 and 8 with ortho, para-diphenylphenol or ortho-2,6-dimethylphenyl (Dmp), para-methylphenol substituents, respectively, were prepared in the total yields of 45% or 39%, respectively. Treatment of [OSSO]-type ligands 7 or 8 with Zr(CH2Ph)4 in toluene at room temperature gave the corresponding dibenzyl zirconium(IV) complexes 9 or 10 as air- and moisture-sensitive pale yellow crystals in 83% or 92% yields, respectively (Scheme 2). Similarly to the NMR observation of the related [OSSO]-type dibenzyl zirconium(IV) complex 5 [49,50], all NMR data of 9 and 10 showed the magnetical equivalency of two phenolate moieties as well as the two benzyl ligands, indicating that complexes 9 and 10 take a C2-symmetric, helical structure on the NMR time scale. For example, in the 1H NMR, AB patterns due to the S-benzyl protons appeared at δ 3.15 and 3.34 with J = 14 Hz for 9 and at δ 3.04 and 3.11 with J = 15 Hz for 10. The sulfur-bonded methine protons in the cyclooctane ring were observed at δ 2.39 for 9 and δ 2.41 for 10 as a broad singlet.
Scheme 2. Synthesis of aryl-substituted [OSSO]-type dibenzyl zirconium(IV) complexes 9 and 10.
Scheme 2. Synthesis of aryl-substituted [OSSO]-type dibenzyl zirconium(IV) complexes 9 and 10.
Polymers 08 00031 g004
The molecular structure of diphenyl derivative 9 was confirmed by X-ray crystallography, as shown in Figure 1. In the crystalline state, the zirconium center possesses a distorted octahedral geometry with cis-arranged two sulfur atoms and two benzyl groups, adopting a cis-α configuration as similar as the previously reported [OSSO]-type zirconium(IV) complexes [49,50,57,58,59,60]. One benzyl ligand of 9 is coordinated to the zirconium center by η1-mode with the Zr–Cbenzyl–Cipso angle of 124.3(2)°, whereas the other has an acute Zr–Cbenzyl–Cipso angle of 84.4(2)° consistent with the η2-coordination fashion as similarly as in the case of tBu-substituted complex 5a; the Zr-Cipso distance (2.592(3) Å) is extremely shortened compared with that in 5a (2.820(3) Å) [48], indicating that this η2-benzyl group in 9 is strongly bound to the metal than that in 5a due to the less steric hindrance around the zirconium center in 9 as expected. The Zr–S bond lengths in 9 (2.8073(7), 2.8148(11) Å) are comparable to those in 5a (2.8107(8), 2.7682(8) Å) [49,50] and [Zr{2,2′-(OC6H2-4,6-Br2)2 CH2SCH2CH2SCH2}(CH2Ph)2] (2.7934(7), 2.7932(6) Å) [61].
Figure 1. ORTEP drawing of dibenzyl zirconium(IV) complex 9 (50% thermal ellipsoids, hydrogen atoms and solvated toluene molecules were omitted for clarity). Selected bond lengths Å and bond angles °: Zr1–O1 = 1.998 (2), Zr1–O2 = 2.005 (2), Zr1–C47 = 2.305 (3), Zr1–C48 = 2.592 (3), Zr1–C54 = 2.321 (4), Zr1–S1 = 2.8148 (11), Zr1–S2 = 2.8079 (9), O1–Zr1–O2 = 160.25 (9), S1–Zr1–S2 = 70.74 (3), C47–Zr1–C54 = 121.54 (13), Zr1–C47–C48 = 84.4 (2), Zr1–C54–C55 = 124.3 (3), S1–C1–C2–S2 = 67.3 (3).
Figure 1. ORTEP drawing of dibenzyl zirconium(IV) complex 9 (50% thermal ellipsoids, hydrogen atoms and solvated toluene molecules were omitted for clarity). Selected bond lengths Å and bond angles °: Zr1–O1 = 1.998 (2), Zr1–O2 = 2.005 (2), Zr1–C47 = 2.305 (3), Zr1–C48 = 2.592 (3), Zr1–C54 = 2.321 (4), Zr1–S1 = 2.8148 (11), Zr1–S2 = 2.8079 (9), O1–Zr1–O2 = 160.25 (9), S1–Zr1–S2 = 70.74 (3), C47–Zr1–C54 = 121.54 (13), Zr1–C47–C48 = 84.4 (2), Zr1–C54–C55 = 124.3 (3), S1–C1–C2–S2 = 67.3 (3).
Polymers 08 00031 g001

3.2. Styrene Polymerization

To elucidate the potential as an isospecific polymerization catalyst, we examined the coordinate polymerization of styrene using precursors 9 and 10. The results at varied temperatures are compiled in Table 1. The polymerization of styrene (3.0 g, 28.8 mmol) with 9 (2.0 μmol) upon activation with 250 equiv. of dMAO (500 μmol) in toluene at 0 °C proceeded slowly to yield a crystalline polystyrene (0.099 g) (Run 1). The microstructure of the polystyrene was determined by 13C NMR spectroscopy, which showed six sharp signals to mean an excellent isotacticity over 99% of the [mm] triad (Figure 2). Since no stereo error was observed around the lowest resonance at δ 146.3 assigned to the phenyl ipso carbon, the detail microstructure is attributed to the [mm] heptad. Similarly to the case for ansa-type complexes 3 reported by Carpentier [41,42], an enantiomorphic site control mechanism is operating for the stereospecific propagation in our system, probably due to the racemic coordination of the [OSSO]-type ligand framework. This high isotactic microstructure can be corroborated from the melting temperature (Tm = 218.6 °C) determined by DSC analysis. The GPC analysis of the resulting polymer revealed a large molecular weight (Mw = 137,000 g·mol−1) and a monomodal distribution with a very broad polydispersity (PDI = 8.6). When the polymerization temperature was raised to 25 or 40 °C, slightly decreasing isotacticity of 96.8% or 90.7% together with narrower molecular weight distributions of 3.8 or 2.9 were observed, respectively (Runs 2 and 3). At 70 °C, 9/dMAO system showed the highest activity of 618 g·mmol(9)−1·h−1, and the narrow PDI value of 2.2, while the isotacticity of the obtained polymer decreased somewhat to 87.5% and the molecular weight Mw was reduced drastically to 9500 g·mol−1, suggesting that the active species was unstable during the chain propagation step at high temperatures (Run 4). In the varying temperature experiments, the broader polydispersity and the higher molecular weight Mw were observed at lower temperatures as characteristic trends. These results would be explained by a relatively slow generation of active species and/or a slow propagation of polymer chains in the styrene polymerization process.
We then turned our attention to the use of Dmp-substituted complex 10. The polymerization of styrene (3.0 g, 28.8 mmol) employing the system consisting of 10 (2.0 μmol) and dMAO (500 μmol) in toluene at 0 °C for 10 min could also afford completely isotactic polystyrene of 0.259 g ([mm] > 99%), which corresponds to the activity of 777 g·mmol(10)−1·h−1 (Run 5). Despite the more overcrowded environment at zirconium center in 10, this activity is approximately 15 times higher than that of 9 at the same temperature (50 g·mmol(9)−1·h−1). GPC analysis exhibited that the obtained polymer was monomodal and had a high molecular weight (Mw = 257,000 g·mol−1) and a slightly large polydispersity of 3.1. Similar results were observed in the polymerizations carried out at 25 and 40 °C to produce excellent isotactic polymers ([mm] > 99%) with higher molecular weights Mw up to 380,000 g·mol−1 and narrower molecular weight distributions of 1.8 and 2.1, respectively, being consistent with a single site behavior (Runs 6 and 7). The corresponding activities recorded in the range of 2200–4100 g·mmol(10)−1·h−1 increased with elevating temperatures; at 70 °C, 10/dMAO system achieved the highest polymerization activity of 7700 g·mmol(10)−1·h−1 forming a high molecular weight polystyrene (Mw = 195,000 g·mol−1, Run 8). Despite the higher polymerization temperature, the resulting polymer had a complete isotacticity ([mm] > 99%) and a monomodal molecular weight distribution (Mw/Mn = 1.8). In the DSC measurement, the melting temperatures (Tm = 221.4–225.8 °C) of polymers produced by 10/dMAO system reflected their highly isotactic microstructures. Thus, our catalyst system exhibited remarkably higher activity, even at low temperatures (777–7700 g·mmol(10)−1·h−1 at 0–70 °C) than those of reported [OSSO]-type titanium complex 1a (330 g·mmol(1)−1·h−1 at 40 °C) [8] and ansa-type yttrium and neodymium complexes 3a (1066–1637 g·mmol(cat)−1·h−1 at 80–120 °C) and 3b (392–1094 g·mmol(cat)−1·h−1 at 60–100 °C) [11].
Table 1. Styrene polymerization with dibenzyl zirconium(IV) complexes 9 and 10 upon activation with dMAO.
Table 1. Styrene polymerization with dibenzyl zirconium(IV) complexes 9 and 10 upon activation with dMAO.
RunCat.Temp.(°C)Time (min)Activity (g·mmol−1·h−1)Mw (g·mol−1)PDI b(mm) c (%)Tm d (°C)
1906050137,0008.6>99218.6
292560139156,0003.896.8209.8
394060182181,0002.990.7-
49706061895002.287.5-
510010777257,0003.1>99225.3
61025102200380,0002.1>99225.8
7104054100338,0001.8>99222.6
8107057700195,0001.8>99221.4
a Conditions: 9 and 10 2.0 μmol, [dMAO]/[Cat.] = 250, styrene 3.0 g (28.8 mmol), toluene 5 mL; b PDI = Mw/Mn, determined by GPC (PS standard); c Determined by 13C{1H} NMR spectrum d Determined by DSC.
Figure 2. 13C{1H} NMR spectrum of polystyrene obtained by the 9/dMAO system at 0 °C (Table 1, Run 1).
Figure 2. 13C{1H} NMR spectrum of polystyrene obtained by the 9/dMAO system at 0 °C (Table 1, Run 1).
Polymers 08 00031 g002

4. Conclusions

We have established a controlled isospecific polymerization of styrene employing dibenzyl zirconium(IV) complexes (9 and 10) supported by new aryl-substituted [OSSO]-type bis(phenolate) ligands based on trans-cyclooctanediyl platform. Use of the Dmp-substituted pre-catalyst 10 in the presence of dMAO as an activator was critical for the formation of precisely isospecific polystyrenes with high activities. Notably, the isotactic polymers produced by 10/dMAO system had significantly larger molecular weights and narrower polydispersity indexes of nearly 2.0.

Supplementary Materials

Supplementary materials can be accessed at www.mdpi.com/2073-4360/8/2/31/s1. Figure S1. 1H-NMR spectrum of dibenzyl zirconium(IV) complex 9; Figure S2. 1H-NMR spectrum of dibenzyl zirconium(IV) complex 10; Figure S3. 13C{1H}-NMR spectrum of polystyrene obtained by the 10/dMAO system at 0 °C (Table 1, Run 5); Figure S4. 13C{1H}-NMR spectrum of polystyrene obtained by the 10/dMAO system at 25 °C (Table 1, Run 6); Figure S5. 13C{1H}-NMR spectrum of polystyrene obtained by the 10/dMAO system at 40 °C (Table 1, Run 7); Figure S6. 13C{1H}-NMR spectrum of polystyrene obtained by the 10/dMAO system at 70 °C (Table 1, Run 8); Figure S7. DSC chart of polystyrene obtained by the 10/dMAO system at 0 °C (Table 1, Run 5); Figure S8. DSC chart of polystyrene obtained by the 10/dMAO system at 25 °C (Table 1, Run 6); Figure S9. DSC chart of polystyrene obtained by the 10/dMAO system at 40 °C (Table 1, Run 7); Figure S10. DSC chart of polystyrene obtained by the 10/dMAO system at 70 °C (Table 1, Run 8); Scheme S1. Preparation of dibenzyl zirconium(IV) complex 9; Scheme S2. Preparation of dibenzyl zirconium(IV) complex 10.

Acknowledgments

This work was supported by Grants-in-Aid (No. 23750034 to Norio Nakata) from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT), Japan. Norio Nakata is thankful for financial support from the Japan Petroleum Institute and Tokuyama Science Foundation. Tomoyuki Toda acknowledges a JSPS fellowship for young scientists.

Author Contributions

Tomoyuki Toda, Yusuke Saito, and Takanori Watanabe contributed to the synthesis, polymerization, and data analysis; Norio Nakata performed X-ray crystallography and wrote the manuscript; and Norio Nakata and Akihiko Ishii proposed idea on the design of the experiment, reviewed and approved the final manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tomotsu, N.; Ishihara, N.; Newman, T.H.; Malanga, M.T. Syndiospecific polymerization of styrene. J. Mol. Catal. A 1998, 128, 167–190. [Google Scholar] [CrossRef]
  2. Malanga, M. Syndiotactic polystyrene materials. Adv. Mater. 2000, 12, 1869–1872. [Google Scholar] [CrossRef]
  3. Ishihara, N.; Kuramoto, M.; Uoi, M.; (to Idemitsu Kosan Co. Ltd.). Japanese Patent 62187708, 1985.
  4. Ishihara, N.; Seimiya, T.; Kuramoto, M.; Uoi, M. Crystalline syndiotactic polystyrene. Macromolecules 1986, 19, 2464–2465. [Google Scholar] [CrossRef]
  5. Luo, Y.; Baldamus, J.; Hou, Z. Scandium half-metallocene-catalyzed syndiospecific styrene polymerization and styrene-ethylene copolymerization: Unprecedented incorporation of syndiotactic styrene-styrene sequences in styrene–ethylene copolymers. J. Am. Chem. Soc. 2004, 126, 13910–13911. [Google Scholar] [CrossRef] [PubMed]
  6. Kirillov, E.; Lehmann, C.W.; Razavi, A.; Carpentier, J.-F. Highly syndiospecific polymerization of styrene catalyzed by allyl lanthanide complexes. J. Am. Chem. Soc. 2004, 126, 12240–12241. [Google Scholar] [CrossRef] [PubMed]
  7. Rodrigues, A.-S.; Kirillov, E.; Lehmann, C.W.; Roisnel, T.; Vuillemin, B.; Razavi, A.; Carpentier, J.-F. Allyl ansa-lanthanidocenes: Single-component, single-site catalysts for controlled syndiospecific styrene and styrene–ethylene (co)polymerization. Chem. Eur. J. 2007, 13, 5548–5565. [Google Scholar] [CrossRef] [PubMed]
  8. Rodrigues, A.-S.; Kirillov, E.; Vuillemin, B.; Razavi, A.; Carpentier, J.-F. Binary ansa-lanthanidocenes/dialkylmagnesium systems versus single-component catalyst: Controlled synthesis of end-capped syndiotactic oligostyrenes. J. Mol. Catal. A 2007, 273, 87–91. [Google Scholar] [CrossRef]
  9. Jaroschik, F.; Shima, T.; Li, X.; Mori, K.; Ricard, L.; Le Goff, X.-F.; Nief, F.; Hou, Z. Synthesis, characterization, and reactivity of mono(phospholyl) lanthanoid(III) bis(dimethylaminobenzyl) complexes. Organometallics 2007, 26, 5654–5660. [Google Scholar] [CrossRef]
  10. Nishiura, M.; Mashiko, T.; Hou, Z. Synthesis and styrene polymerisation catalysis of η5- and η1-pyrrolyl-ligated cationic rare earth metal aminobenzyl complexes. Chem. Commun. 2008, 2019–2021. [Google Scholar] [CrossRef] [PubMed]
  11. Fang, X.; Li, X.; Hou, Z.; Assoud, J.; Zhao, R. 1,2-Azaborolyl-ligated half-sandwich complexes of scandium(III) and lutetium(III): Synthesis, structures, and syndiotactic polymerization of styrene. Organometallics 2009, 28, 517–522. [Google Scholar] [CrossRef]
  12. Xu, X.; Chen, Y.; Sun, J. Indenyl abstraction versus alkyl abstraction of [(Indenyl)ScR2(thf)] by [Ph3C] [B(C6F5)4]: Aspecific and syndiospecific styrene polymerization. Chem. Eur. J. 2009, 15, 846–850. [Google Scholar] [CrossRef] [PubMed]
  13. Bonnet, F.; Violante, C.D.C.; Roussel, P.; Mortreux, A.; Visseaux, M. Unprecedented dual behaviour of a half-sandwich scandium-based initiator for both highly selective isoprene and styrene polymerisation. Chem. Commun. 2009, 3380–3382. [Google Scholar] [CrossRef] [PubMed]
  14. Perrin, L.; Sarazin, Y.; Kirillov, E.; Carpentier, J.-F.; Maron, L. On the initiation mechanism of syndiospecific styrene polymerization catalyzed by single-component ansa-lanthanidocenes. Chem. Eur. J. 2009, 15, 3773–3783. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, D.; Luo, Y.; Gao, W.; Cui, D. Stereoselective polymerization of styrene with cationic scandium precursors bearing quinolyl aniline ligands. Organometallics 2010, 29, 1916–1923. [Google Scholar] [CrossRef]
  16. Li, X.; Nishiura, M.; Hu, L.; Mori, K.; Hou, Z. Alternating and random copolymerization of isoprene and ethylene catalyzed by cationic half-sandwich scandium alkyls. J. Am. Chem. Soc. 2009, 131, 13870–13882. [Google Scholar] [CrossRef] [PubMed]
  17. Guo, F.; Nishiura, M.; Koshino, H.; Hou, Z. Scandium-catalyzed cyclocopolymerization of 1,5-hexadiene with styrene and ethylene: Efficient synthesis of cyclopolyolefins containing syndiotactic styrene–styrene sequences and methylene-1,3-cyclopentane units. Macromolecules 2011, 44, 6335–6344. [Google Scholar] [CrossRef]
  18. Pan, L.; Zhang, K.; Nishiura, M.; Hou, Z. Syndiospecific living copolymerization of styrene with ε-caprolactone by scandium catalysts. Macromolecules 2010, 43, 9591–9593. [Google Scholar] [CrossRef]
  19. Pan, Y.; Rong, W.; Jian, Z.; Cui, D. Ligands dominate highly syndioselective polymerization of styrene by using constrained-geometry-configuration rare-earth metal precursors. Macromolecules 2012, 45, 1248–1253. [Google Scholar] [CrossRef]
  20. Natta, G.; Pino, P.; Corradini, P.; Danusso, F.; Mantica, E.; Mazzanti, G.; Moraglio, G. Crystalline high polymers of α-olefins. J. Am. Chem. Soc. 1955, 77, 1708–1710. [Google Scholar] [CrossRef]
  21. Natta, G.; Corradini, P. Kristallstruktur des isotaktischen polystyrols. Makromol. Chem. 1955, 16, 77–80. [Google Scholar] [CrossRef]
  22. Overberger, C.; Mark, H. A convenient laboratory preparation of isotactic polystyrene. J. Polym. Sci. 1959, 35, 381–389. [Google Scholar] [CrossRef]
  23. Kern, R.J.; Hurst, H.G.; Richard, W.J. Triethylaluminum–titanium tetrachloride catalysts for preparation of crystalline polystyrene. J. Polym. Sci. 1960, 45, 195–204. [Google Scholar] [CrossRef]
  24. Xu, G.; Lin, S. A novel NdCl3-modified Ziegler-Natta catalyst for the isotactic-specific polymerization of styrene. Macromol. Rapid Commun. 1994, 15, 873–877. [Google Scholar] [CrossRef]
  25. Rosário Riberio, M.; Portela, M.F.; Deffieux, A.; Cramail, H.; Rocha, M.F. Isospecific homo- and copolymerization of styrene with ethylene in the presence of VCl3, AlCl3 as catalyst. Macromol. Rapid Commun. 1996, 17, 461–469. [Google Scholar] [CrossRef]
  26. Sun, Q.; Fan, Y.; Liao, S.; Liu, J.; Wan, F.; Xu, J. Isospecific polymerization of styrene with modified Ziegler-type catalysts. Polymer 2001, 42, 4087–4090. [Google Scholar] [CrossRef]
  27. Kern, R.J. Homogeneous synthesis of isotactic polystyrene using n-butyllithium initiator. Nature 1960, 187, 410. [Google Scholar] [CrossRef]
  28. Cazzaniga, L.; Cohen, R.E. Anionic synthesis of isotactic polystyrene. Macromolecules 1989, 22, 4125–4128. [Google Scholar] [CrossRef]
  29. Makino, T.; Hogen-Esch, T.E. Anionic synthesis of highly isotactic polystyrene in hexane in the presence of lithium hydroxides. Macromolecules 1999, 32, 5712–5714. [Google Scholar] [CrossRef]
  30. Maréchal, J.-M.; Carlotti, S.; Shcheglova, L.; Deffieux, A. Stereospecific anionic polymerization of styrene initiated by R2Mg/ROMt ‘ate’ complexes. Polymer 2004, 45, 4641–4646. [Google Scholar] [CrossRef]
  31. Ascenso, J.R.; Dias, A.R.; Gomes, P.T.; Romao, C.C.; Tkatchenko, I.; Revillon, A.; Pham, Q.T. Isospecific oligo-/polymerization of styrene with soluble cationic nickel complexes. The influence of phosphorus(III) ligands. Macromolecules 1996, 29, 4172–4179. [Google Scholar] [CrossRef]
  32. Crossetti, G.L.; Bormioli, C.; Ripa, A.; Giarrusso, A.; Porri, L. Polymerization of styrene to isotactic polymer with MAO-Ni(acac)2. Examination of the factors that influence activity and stereospecificity. Macromol. Rapid Commun. 1997, 18, 801–808. [Google Scholar] [CrossRef]
  33. Po, R.; Cardi, N.; Santi, R.; Romano, A.M.; Zannoni, C.; Spera, S. Polymerization of styrene with nickel complex/methylaluminoxane catalytic systems. J. Polym. Sci., Part A Polym. Chem. 1998, 36, 2119–2126. [Google Scholar] [CrossRef]
  34. Arai, T.; Ohtsu, T.; Suzuki, S. Homo- and copolymerization of styrene by bridged zirconocene complexes with benzindenyl ligands. Polym. Prepr. 1998, 39, 220–221. [Google Scholar]
  35. Arai, T.; Suzuki, S.; Ohtsu, T. Homo- and copolymerization of styrene by bridged zirconocene complex with benz indenyl ligand. ACS Symp. Ser. 2009, 749, 66–80. [Google Scholar]
  36. Capacchione, C.; Proto, A.; Ebeling, H.; Mülhaupt, R.; Möller, K.; Spaniol, T.P.; Okuda, J. Ancillary ligand effect on single-site styrene polymerization: Isospecificity of Group 4 metal bis(phenolate) catalysts. J. Am. Chem. Soc. 2003, 125, 4964–4965. [Google Scholar] [CrossRef] [PubMed]
  37. Capacchione, C.; Manivannan, R.; Barone, M.; Beckerle, K.; Centore, R.; Oliva, L.; Proto, A.; Tuzi, A.; Spaniol, T.P.; Okuda, J. Isospecific styrene polymerization by chiral titanium complexes that contain a tetradentate [OSSO]-Type bis(phenolato) ligand. Organometallics 2005, 24, 2971–2982. [Google Scholar] [CrossRef]
  38. Hohberger, C.; Spaniol, T.P.; Okuda, J. Living polymerization by bis(phenolate) zirconium catalysts: Synthesis of isotactic polystyrene-block-polybutadiene copolymers. Macromol. Chem. Phys. 2014, 215, 2001–2006. [Google Scholar] [CrossRef]
  39. Proto, A.; Avagliano, A.; Saviello, D.; Ricciardi, R.; Capacchione, C. Living, isoselective polymerization of styrene and formation of stereoregular block copolymers via sequential monomer addition. Macromolecules 2010, 43, 5919–5921. [Google Scholar] [CrossRef]
  40. Capacchione, C.; Saviello, D.; Ricciardi, R.; Proto, A. Living, isoselective polymerization of 4-methyl-1,3-pentadiene and styrenic monomers and synthesis of highly stereoregular block copolymers via sequential monomer addition. Macromolecules 2011, 44, 7940–7947. [Google Scholar] [CrossRef]
  41. Rodrigues, A.-S.; Kirillov, E.; Roisnel, T.; Razavi, A.; Vuillemin, B.; Carpentier, J.-F. Highly isospecific styrene polymerization catalyzed by single-component bridged bis(indenyl) allyl yttrium and neodymium complexes. Angew. Chem., Int. Ed. 2007, 46, 7240–7243. [Google Scholar] [CrossRef] [PubMed]
  42. Annunziata, L.; Rodrigues, A.-S.; Kirillov, E.; Sarazin, Y.; Okuda, J.; Perrin, L.; Maron, L.; Carpentier, J.-F. Isoselective styrene polymerization catalyzed by ansa-bis(indenyl) allyl rare earth complexes. Stereochemical and mechanistic aspects. Macromolecules 2011, 44, 3312–3322. [Google Scholar] [CrossRef]
  43. Nakata, N.; Toda, T.; Ishii, A. Recent advances in the chemistry of Group 4 metal complexes incorporating [OSSO]-type bis(phenolato) ligands as post-metallocene catalysts. Polym. Chem. 2011, 2, 1597–1610. [Google Scholar] [CrossRef]
  44. Nakata, N.; Ishii, A. Precise polymerization of α-olefins using a mixed donor-type ligand containing oxygen and sulfur atoms. Koubunshi Ronbunshu 2015, 72, 285–294. [Google Scholar] [CrossRef]
  45. Toda, T.; Nakata, N.; Matsuo, T.; Ishii, A. Synthesis and structures of dialkyl zirconium complexes with an [OSSO]-type bis(phenolate) ligand bearing a trans-1,2-cyclooctanediylbis(thio) unit. J. Organomet. Chem. 2011, 696, 1258–1261. [Google Scholar] [CrossRef]
  46. Nakata, N.; Toda, T.; Ishii, A.; Matsuo, T. Titanium complexes supported by an [OSSO]-Type bis(phenolato) ligand based on a trans-cyclooctanediyl platform: Synthesis, structures, and 1-hexene polymerization. Inorg. Chem. 2012, 51, 274–281. [Google Scholar] [CrossRef] [PubMed]
  47. Toda, T.; Nakata, N.; Ishii, A.; Matsuo, T. Synthesis, structure, and 1‑hexene polymerization catalytic ability of Group 5 metal complexes incorporating an [OSSO]-type ligand. ACS Catal. 2013, 3, 1764–1767. [Google Scholar] [CrossRef]
  48. Nakata, N.; Saito, Y.; Ishii, A. Neutral methyl and cationic aluminum complexes supported by a trans-1,2-cyclooctanediyl-bridged [OSSO]-type bis(phenolato) ligand: Synthesis, structures, and use in catalysis for propylene oxide polymerization. Organometallics 2014, 33, 1840–1844. [Google Scholar] [CrossRef]
  49. Ishii, A.; Toda, T.; Nakata, N.; Matsuo, T. Zirconium complex of an [OSSO]-type diphenolate ligand bearing trans-1,2-cyclooctanediylbis(thio) core: Synthesis, structure, and isospecific 1-hexene polymerization. J. Am. Chem. Soc. 2009, 131, 13566–13567. [Google Scholar] [CrossRef] [PubMed]
  50. Toda, T.; Nakata, N.; Matsuo, T.; Ishii, A. Extremely active α-olefin polymerization and copolymerization with ethylene catalyzed by a dMAO-activated zirconium(IV) dichloro complex having an [OSSO]-type ligand. RSC Adv. 2015, 8, 88826–88831. [Google Scholar] [CrossRef]
  51. Nakata, N.; Toda, T.; Matsuo, T.; Ishii, A. Controlled isospecific polymerization of α‑olefins by hafnium complex incorporating with a trans-cyclooctanediyl-bridged [OSSO]-type bis(phenolate) ligand. Macromolecules 2013, 46, 6758–6764. [Google Scholar] [CrossRef]
  52. Nakata, N.; Saito, Y.; Watanabe, T.; Ishii, A. Completely isospecific polymerization of 1-hexene catalyzed by hafnium(IV) dichloro complex incorporating with an [OSSO]-type bis(phenolate) ligand. Top. Catal. 2014, 57, 918–922. [Google Scholar] [CrossRef]
  53. Takano, M.; Ito, K.; Ishii, A.; Nakata, N.; Kawauchi, F. Jpn. Kokai Tokkyo Koho JP 2013166735 29 August 2013.
  54. Hasan, T.; Ioku, A.; Nishii, K.; Shiono, T.; Ikeda, T. Syndiospecific living polymerization of propene with [t-BuNSiMe2Flu]TiMe2 using MAO as cocatalyst. Macromolecules 2001, 34, 3142–3145. [Google Scholar] [CrossRef]
  55. Rong, Y.; Al-Harbi, A.; Parkin, G. Highly variable Zr-CH2-Ph bond angles in tetrabenzylzirconium: Analysis of benzyl ligand coordination modes. Organometallics 2012, 31, 8208–8217. [Google Scholar] [CrossRef]
  56. Sheldrick, G.M. SHELXL-97. In Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  57. Cohen, A.; Yeori, A.; Goldberg, I.; Kol, M. Group 4 complexes of a new [OSSO]-Type dianionic ligand. Coordination chemistry and preliminary polymerization catalysis studies. Inorg. Chem. 2007, 46, 8114–8116. [Google Scholar] [CrossRef] [PubMed]
  58. Cohen, A.; Goldberg, I.; Venditto, V.; Kol, M. Oscillating non-metallocenes—From stereoblock-isotactic polypropylene to isotactic polypropylene via zirconium and hafnium dithiodiphenolate catalysts. Eur. J. Inorg. Chem. 2011, 5219–5223. [Google Scholar] [CrossRef]
  59. Buffet, J.-C.; Okuda, J. Group 4 metal initiators for the controlled stereoselective polymerization of lactide monomers. Chem. Commun. 2011, 47, 4796–4798. [Google Scholar] [CrossRef] [PubMed]
  60. Wang, X.; Thevenon, A.; Brosmer, J.L.; Yu, I.; Khan, S.I.; Mehrkhodavandi, P.; Diaconescu, P.L. Redox control of Group 4 metal ring-opening polymerization activity toward l‑lactide and ε‑caprolactone. J. Am. Chem. Soc. 2014, 136, 11264–11267. [Google Scholar] [CrossRef] [PubMed]
  61. Konkol, M.; Nabika, M.; Kohno, T.; Hino, T.; Miyatake, T. Synthesis, structure and α-olefin polymerization activity of Group 4 metal complexes with [OSSO]-type bis(phenolate) ligands. J. Organomet. Chem. 2011, 696, 1792–1802. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Nakata, N.; Toda, T.; Saito, Y.; Watanabe, T.; Ishii, A. Highly Active and Isospecific Styrene Polymerization Catalyzed by Zirconium Complexes Bearing Aryl-substituted [OSSO]-Type Bis(phenolate) Ligands. Polymers 2016, 8, 31. https://doi.org/10.3390/polym8020031

AMA Style

Nakata N, Toda T, Saito Y, Watanabe T, Ishii A. Highly Active and Isospecific Styrene Polymerization Catalyzed by Zirconium Complexes Bearing Aryl-substituted [OSSO]-Type Bis(phenolate) Ligands. Polymers. 2016; 8(2):31. https://doi.org/10.3390/polym8020031

Chicago/Turabian Style

Nakata, Norio, Tomoyuki Toda, Yusuke Saito, Takanori Watanabe, and Akihiko Ishii. 2016. "Highly Active and Isospecific Styrene Polymerization Catalyzed by Zirconium Complexes Bearing Aryl-substituted [OSSO]-Type Bis(phenolate) Ligands" Polymers 8, no. 2: 31. https://doi.org/10.3390/polym8020031

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