Next Article in Journal
Preparation, Characterization and Mechanical Properties of Bio-Based Polyurethane Adhesives from Isocyanate-Functionalized Cellulose Acetate and Castor Oil for Bonding Wood
Next Article in Special Issue
Stereoregular Brush Polymers and Graft Copolymers by Chiral Zirconocene-Mediated Coordination Polymerization of P3HT Macromers
Previous Article in Journal
Homoserine Lactone as a Structural Key Element for the Synthesis of Multifunctional Polymers
Previous Article in Special Issue
Direct Synthesis of Branched Carboxylic Acid Functionalized Poly(1-octene) by α-Diimine Palladium Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Living Polymerization of Propylene with ansa-Dimethylsilylene(fluorenyl)(cumylamido) Titanium Complexes

1
State Key Laboratory for Modification of Chemical Fibers and Polymer Materials, College of Materials Science and Engineering, Donghua University, Shanghai 201620, China
2
State Key Laboratory of Polyolefins and Catalysis, Shanghai Research Institute of Chemical Industry, Shanghai 200062, China
3
Department of Applied Chemistry, Graduate School of Engineering, Hiroshima University, 1-4-1 Kagamiyama, Higashi-Hiroshima 739-8527, Japan
*
Authors to whom correspondence should be addressed.
Polymers 2017, 9(4), 131; https://doi.org/10.3390/polym9040131
Submission received: 9 March 2017 / Revised: 31 March 2017 / Accepted: 3 April 2017 / Published: 5 April 2017
(This article belongs to the Special Issue Metal Complexes-Mediated Catalysis in Polymerization)

Abstract

:
A series of ansa-silylene(fluorenyl)(amido) titanium complexes (1a1c, 2a, and 2b) bearing various substituents on the amido and fluorenyl ligands are synthesized and characterized by elemental analysis, 1H NMR, and single crystal X-ray analysis. The coordination mode of the fluorenyl ligand to the titanium metal is η3 manner in each complex. The propylene polymerization is conducted with these complexes at 0 and 25 °C in a semi batch-type method, respectively. The catalytic activity of 1a1c bearing cumyl-amido ligand is much higher than that of 2a and 2b bearing naphthyl group in amido ligand. High molecular weight polypropylenes are obtained with narrow molecular weight distribution, suggesting a living nature of these catalytic systems at 0 °C. The polymers produced are statistically atactic, regardless of the structure of the complex and the polymerization temperature.

Graphical Abstract

1. Introduction

Half-sandwich group 4 metal complexes (usually named constrained geometry catalysts, CGC) having an ansa-monocyclopentadienylamido ligand (CpA) have captured significant attention due to their various copolymerization abilities and stereospecificity as olefin polymerization catalysts [1,2,3,4]. The replacement of the Cp ligand by the fluorenyl ligand was effective for the improvement of polymerization activity, copolymerization ability, syndiospecificity, and living polymerization nature to produce novel random and block copolymers with controlled microstructure [5,6,7,8,9,10]. The highly syndiospecific polymerization of α-olefin was also achieved with Cs-symmetric Ti or Zr complexes by the introduction of alkyl groups on the fluorenyl ligand [11,12,13,14,15]. Although C1-symmetric metallocene catalysts normally give highly isospecific polypropylene (PP) [16,17], C1-symmetric CGC catalysts exert a less positive influence on the isospecific polymerization of propylene. For example, Fink et al. showed that C1-symmetric CpA complex by the use of naphthylamido ligand produces isospecific PP with isotactic triad (mm) of 0.63 [18]. Therefore, it is important to investigate strategies for understanding the relationship between the complex structure and the stereospecificity.
We have realized syndiospecific living homo- and copolymerization of α-olefins with norbornene using Cs-symmetric ansa-(fluorenyl)(amido)dimethyltitanium complexes with high activity [19,20,21,22]. We also synthesized C1-symmetric complex and the other compounds, and found that the change of the structure of amido ligand did not affect livingness of catalytic system; however, the highest mm value was 0.61 [23,24].
In this paper, we synthesized a series of new (fluorenyl)(amido) titanium complexes (1 and 2 in Figure 1) and investigated the substitute effects of the bulkiness on the amido and fluorenyl ligands in propylene polymerization.

2. Experimental Section

2.1. Materials

All operations were performed under nitrogen gas using standard Schlenk techniques, and all solvents were purified by PS-MD-5 (Innovative Technology, Oldham, UK) solvent purification system. Research grade propylene was purified by passing it through a dehydration column of ZHD-20 (Zhonghuida, Dalian, China) and a deoxidation column of ZHD-20A (Zhonghuida, Dalian, China) before used. Modified methylaluminoxane (MMAO) was donated by Tosoh-Finechem Co. (Tokyo, Japan). The ligands and complexes were prepared according to the literature procedure [14,15,24]. We have previously reported the synthesis of complex 1a [24]. In this work, we report the molecular structure and crystallographic data of complex 1a.

2.2. Synthesis of Complexes

2.2.1. Synthesis of (2,7-di-tBu)fluorenyl-cumylamido Titanium Complex (1b)

MeLi (1.6 M in ether, 15 mL, 24 mmol) was added dropwise at 0 °C into a solution of (2,7-di-tBu)fluorenyl-cumylamido ligand (2.82 g, 6 mmol) in 60 mL of diethylether. The resultant orange solution was stirred at r.t. for 4 h, then was added to a solution of TiCl4 (0.66 mL, 6 mmol) in 30 mL hexane at room temperature in the stirring condition for 2 h. The solvent was removed, and the residue was extracted with hexane (150 mL). The hexane solution was concentrated to 50 mL and cooled at −30 °C for some days to get 1b as red crystals (0.92 g, 28%).
1H NMR (CDCl3): 8.04(d, 2H, Flu), 7.63(s, 2H, Flu), 7.58(d, 2H, Flu), 7.41(d, 2H, Cph), 7.30(t, H, Cph), 7.22(d, 1H, Cph), 1.89(s, 6H, CCH3), 1.39(s, 18H, Flu-t-Bu), 0.42(s, 6H, SiCH3), –0.39(s, 6H, TiCH3). Anal. Calc. for C34H47NSiTi: C, 74.83; H, 8.68; N, 2.57. Found: C, 74.64; H, 8.85; N, 2.67.

2.2.2. Synthesis of (3,6-di-tBu)fluorenyl-cumylamido Titanium Complex (1c)

Complex 1c was synthesized in a similar way to that for 1b, and yellow powders were obtained in 28% yield.
1H NMR (CDCl3): 8.07(s, 2H, Flu), 7.58(d, 2H, Flu), 7.46(d, 2H, Flu), 7.42(d, 2H, Cph), 7.30(t, 2H, Cph), 7.23(d, 1H, Cph), 1.88(s, 6H, CCH3), 1.48(s, 18H, Flu-t-Bu), 0.35(s, 6H, SiCH3), 0.38(s, 6H, TiCH3). Anal. Calc. for C34H47NSiTi: C, 74.83; H, 8.68; N, 2.57. Found: C, 74.60; H, 8.84; N, 2.65.

2.2.3. Synthesis of Fluorenyl-(1-methyl-(1-naphthyl)ethyl)amido Titanium Complex (2a)

Complex 2a was synthesized in a similar way to that for 1b, and red crystals were obtained in 12% yield.
1H NMR (CDCl3): 8.58(d, 1H, Naph), 8.06(d, 2H, Flu), 7.87(d, 1H, Naph), 7.79(d, 1H, Naph), 7.63(d, 2H, Flu), 7.29–7.56(m, 4H, Naph; 4H, Flu), 2.15(s, 6H, CCH3), 0.03(s, 6H, SiCH3), −0.19(s, 6H, TiCH3). Anal. Calc. for C30H33NSiTi: C, 74.52; H, 6.88; N, 2.90. Found: C, 74.77; H, 7.02; N, 2.31.

2.2.4. Synthesis of (2,7-di-tBu)fluorenyl-(1-methyl-(1-naphthyl)ethyl)amido Titanium Complex (2b)

Complex 2b was synthesized in a similar way to that for 1b, and red crystals were obtained in 11% yield.
1H NMR (CDCl3): 8.70(d, 1H, Naph), 8.04(d, 2H, Flu), 7.86(d, 1H, Naph), 7.78(d, 1H, Naph), 7.58(t, 2H, Flu; 3H, Naph), 7.44(m, 2H, Flu), 7.30(t, 1H, Naph), 2.13(s, 6H, CCH3), 1.39(s, 18H, Flu-t-Bu), 0.09(s, 6H, SiCH3), −0.23(s, 6H, TiCH3). Anal. Calc. for C38H49NSiTi: C, 76.61; H, 8.29; N, 2.35. Found: C, 76.39; H, 8.38; N, 2.19.

2.3. Polymerization Procedure

Polymerization was performed in a 100 mL glass reactor equipped with a magnetic stirrer and carried out using semi-batch method. The reactor was charged with prescribed amounts of MMAO/2,6-di-tert-butyl-4-methyl phenol (BHT) and solvent (30 mL). After the solution of cocatalyst was saturated with gaseous propylene under atmospheric pressure, polymerization was started by the addition of 1 mL solution of the Ti complex (20 μmol) in reactor, and the consumption rate of propylene was monitored by a mass flow meter. Polymerization was conducted for a certain time and terminated with acidic methanol. The polymers obtained were adequately washed with methanol and dried under vacuum condition at 80 °C for 6 h to be a constant weight.

2.4. Analytical Procedure

Single crystal of complex was obtained from a solution of hexane at −30 °C. The single crystals were mounted under nitrogen atmosphere at low temperature, and data collection was made on a Bruker APEX2 diffractometer using graphite monochromated with Mo Kα radiation (λ = 0.71073 Å). The SMART program package was used to determine the unit cell parameters. The absorption correction was applied using SADABS program [25]. All structures were solved by direct methods and refined on F2 by full-matrix least-squares techniques with anisotropic thermal parameters for non-hydrogen atoms. Hydrogen atoms were placed at calculated positions and were included in the structure calculation. Calculations were carried out using the SHELXS-97, SHELXL-2014, or Olex2 program [26,27,28,29,30,31,32]. Crystallographic data are summarized in Table 1, and CIF files and check are provided in the Supporting Information.
Molecular weights and molecular weight distributions of polymers were measured by a polymer laboratory PL GPC-220 chromatograph (Agilen, Santa Clara, CA, USA) equipped with one PL1110-1120 column and two PL MIXED-B 7.5 mm × 300 mm columns at 150 °C using 1,2,4-trichlorobenzene as solvent and calibrated by polystyrene standards. The 1H NMR spectra of complexes and 13C NMR spectra of PPs were measured on a Bruker Asend™ 600 spectrometer (Bruker, Karlsruhe, Germany). Differential scanning calorimeter (DSC) analyses were performed on TA Q2000 instrument (Waters, New Castle, DE, USA), and the DSC curves of the samples were recorded under a nitrogen atmosphere at a heating rate of 10 °C/min from 40 to 200 °C.

3. Results and Discussion

3.1. Molecular Structure of Complexes

The complexes were synthesized in good yield using one-pot reaction of the corresponding ligand with a two-fold excess of MeLi and TiCl4 in hexane. The molecular structure of complexes determined by single crystal X-ray analysis are shown in Figure 2, and the selected bond lengths and angles of complexes are shown in Table 2. The bond lengths between the titanium and the fluorenyl carbons (C(1), C(2), C(3), C(4), and C(5)) are almost the same as that of previously reported Cs-symmetric (fluorenyl)(amido) titanium complexes [24], which suggests that the fluorenyl ligand is coordinating to the titanium metal with η3-form, irrespective of the substituent structure.
In each complex, no symmetry plane or axis was observed, indicating the C1-symmetric nature in solid state. On the other hand, the 1H NMR spectra of all the complexes showed that the methyl groups bonded to the Si and Ti atoms are equivalent, which can be explained by the fast rotation of the admido ligand on the NMR time-scale at room temperature.

3.2. Propylene Polymerization

Propylene polymerizations were conducted with 1a1c, 2a, and 2b activated by MMAO in a semi batch-type operation under an atmospheric pressure of propylene in toluene at 0 and 25 °C (Table 3). Cumyl-amido complexes 1a1c showed much higher activity (1900–2500 kg PP/(mol Ti·h)) than naphthyl-amido complexes 2a and 2b (20–120 kg PP/(mol Ti·h)), of which value increased according to the increase of polymerization temperature and was independent of the introduction of t-butyl group on the fluorenyl ligand. Significantly lower activity in 2a and 2b can be attributed to the steric hindrance of the amido group, where the steric hindrance of naphthyl group (102°/103°) is bigger than the cumyl group (86°/90°) as estimated by Tolman cone angle (Table 4) [33].
1a1c gave high molecular weight PP (Mn > 230,000) for 5 min at 0 °C with narrow molecular weight distribution, and the number of polymer chains (N) were about 65–70% of the titanium complex used. These results suggest that the propylene polymerization proceeded in a living manner. The molecular weights of PPs were decreased by raising the polymerization temperature, with broader molecular weight distribution accompanied by the increase of N value, indicating that the chain transfer reaction frequently occurred by the improvement of polymerization temperature. The molecular weights of PPs obtained by 2a and 2b were lower, but the molecular weight distributions were extremely narrow. The N values were about 25% of the titanium used, which indicates poor catalytic efficiency by the introduction of the bulky naphthyl group.
All the PPs obtained were amorphous measured by the DSC analysis. To investigate the substituent effect of the amido group on the stereospecificity, we conducted 13C NMR analysis of PPs obtained. The steric pentad distributions are summarized in Table 5. All the polymers obtained in each catalyst system were statistically atactic. The result indicates that the introduction of sterically demanded N-alkyl and t-butyl groups on the amido and the fluorenyl ligand, respectively, had little influence on the stereospecificity.

4. Conclusions

(Fluorenyl)(amido) titanium complexes 1a1c, 2a, and 2b were synthesized and characterized by 1H NMR, elemental analysis, and single crystal X-ray analysis. All synthetic catalysts are Cs-symmetric in solution, irrespective of the bulkiness of the amido ligand. The cumyl-amido complexes 1a1c showed much higher activity than the naphthyl-amido complexes 2a and 2b because of the steric hindrance of the amido group. The modification of the structure did not change the livingness of the catalytic system, whereas the change of ligand structure was not efficient for the isospecificity of CGC catalyst.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4360/9/4/131/s1.

Acknowledgments

This work was supported by National Natural Science Foundation of China (Grant No. 21174026), the program for New Century Excellent Talents in University, the Program for Prefessor of Special Appointment (Eastern Scholar) at Shanghai Institutions of Higher Learning, “Shu Guang” project supported by Shanghai Municipal Education Commission and Shanghai Education Development Foundation, and the Fundamental Research Funds for the Central Universities. The authors thank Tosoh-Finechem Co. for generous donating MMAO.

Author Contributions

All authors tried their best to contribute effectively to perform and analyze this experimental work. They all participated to the writing of the present manuscript. Huajin Wang performed the overall experimental work. Xinwei Wang, Yanjie Sun, and Hailong Cheng participated in analysis of structural data. The settings up of the experimental protocols as well as the interpretation of the obtained results were performed under the supervision of Zhengguo Cai and Takeshi Shiono.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stevent, J.C.; Timmers, F.J.; Wilson, D.R.; Schmidt, G.F.; Nickias, P.N.; Rosen, R.K.; Knight, G.W.; Lai, S.Y. Constrain Geometry Addition Polymerization Catalysts, Processer for Their Preparation, Precursors Therefor, Methods of use, and Novel Polymers Formed Therewith. Patent EP0416815, 1991. [Google Scholar]
  2. Canich, J.A.M. Monocyclopentadienyl Titanium Metal Compounds for Ethylene-α-Olefin-Copolymer Production Catalysts. U.S. Patent 5264405, 15 October 1993. [Google Scholar]
  3. Mcknight, A.L.; Waymouth, R.M. Group 4 ansa-cyclopentadienyl-amido catalysts for olefin polymerization. Chem. Rev. 1998, 98, 2587–2598. [Google Scholar] [CrossRef] [PubMed]
  4. Braunschweig, H.; Breitling, F.M. Constrained geometry complexes-synthesis and applications. Coord. Chem. Rev. 2006, 250, 2691–2720. [Google Scholar] [CrossRef]
  5. Nakayama, Y.; Sogo, Y.; Cai, Z. Copolymrization of ethylene with 1,1-disubstituted olefins catalyzed by ansa-(fluorenyl)(cyclododecylamido) dimethyltitanium complexes. J. Polym. Sci. Part A 2013, 51, 1223–1229. [Google Scholar] [CrossRef]
  6. Xu, G. Copolymerization of ethylene with styrene catalyzed by the [η15-tert-butyl(dimethylfluorenylsilyl)amido] metthyltitanium “Cation”. Macromolecules 1998, 31, 2395–2402. [Google Scholar] [CrossRef]
  7. Kirillov, E.; Razavi, A.; Carpentier, J. Syndiotactic-enriched propylene-styrene copolymer using fluorenyl-based half-titanocene catalysts. J. Mol. Catal. Part A 2006, 249, 230–235. [Google Scholar] [CrossRef]
  8. Shiono, T.; Sugimoto, M.; Hasan, T. Random copolymerization of norbornene with higher 1-alkene with ansa-fluorenylamidodimethyltitanium catalyst. Macromolecules 2008, 41, 8292–8294. [Google Scholar] [CrossRef]
  9. Shiono, T.; Suginoto, M.; Hasan, T.; Cai, Z. Facile synthesis of hydroxy-functionalized cycloolefin copolymer using ω-alkenylaluminium as a comonomer. Macromol. Chem. Phys. 2013, 214, 2239–2244. [Google Scholar] [CrossRef]
  10. Tanaka, R.; Ikeda, T.; Nakayama, Y.; Shiono, T. Pseudo-living copolymerization of norbornene and ω-alkynelborane—Synthesis of monodisperse functionalized cycloolefin copolymer. Polymer 2015, 56, 218–222. [Google Scholar] [CrossRef]
  11. Razavi, A.; Thewalt, U. Preparation and crystal structures of the complexes (η5-C5H3TMS-CMe25-C13H8)MCl2(M = Hf, Zr or Ti): Mechanistic aspects of the catalytic formation of a isotactic-syndiotactic stereoblock-type polypropylene. J. Organomet. Chem. 2001, 621, 267–276. [Google Scholar] [CrossRef]
  12. Busico, V.; Cipullo, R.; Cutillo, F.; Talarico, G.; Razavi, A. Syndiotactic poly(propylene) from [Me2Si(3,6-di-tert-butyl-9-fluorenyl) (N-tert-butyl)]TiCl2-based catalysts: Chain-end or enantiotopic-sites stereocontrol? Macromol. Chem. Phys. 2003, 204, 1269–1274. [Google Scholar] [CrossRef]
  13. Irwin, L.J.; Miller, S.A. Unprecedented syndioselectivity and syndiotactic polyolefin melting temperature: Polypropylene and poly(4-methyl-1-pentene) from a highly active, sterically expanded η1-fluorenyl-η1-amido zirconium complex. J. Am. Chem. Soc. 2005, 127, 9972–9973. [Google Scholar] [CrossRef] [PubMed]
  14. Nishii, K.; Hagihara, H.; Ikeda, T.; Akita, M.; Shiono, T. Stereospecific polymerization of propylene with group 4 ansa-fluorenylamidodimethyl complexes. J. Organomet. Chem. 2006, 691, 193–201. [Google Scholar] [CrossRef]
  15. Miller, S.A.; Bercaw, J.E. Highly Stereoregular syndiotactic polypropylen formation with metallocene catalysts via influence of distal ligand substituents. Organometallics 2004, 23, 1777–1789. [Google Scholar] [CrossRef]
  16. Bader, M.; Marquet, N.; Kirillov, E.; Roisnel, T.; Razavi, A.; Lhost, O.; Carpentier, J.F. Old and new C1-symmetric group 4 metallocenes {(R1R2C)-(R2’R3’R6’R7-Flu)(3-R3-5-R4-C5H2)}ZrCl2: From highly isotactic polypropylenes to vinyl end capped isotactic-enriched oligomers. Organometallics 2012, 31, 8375–8387. [Google Scholar] [CrossRef]
  17. Esteb, J.J.; Chien, J.C.C.W.; Rausch, M.D. Novel C1 symmetric zirconocenes containing substituted indenyl moieties for the stereoregular polymerization of propylene. J. Organomet. Chem. 2003, 688, 153–160. [Google Scholar] [CrossRef]
  18. Kleinschmidt, R.; Griebenow, Y.; Fink, G. Stereospecific propylene polymerization using half-sandwich metallocene/MAO systems: A mechanistic insight. J. Mol. Catal. A Chem. 2000, 157, 83–90. [Google Scholar] [CrossRef]
  19. Cai, Z.; Nakayama, Y.; Shiono, T. Living random copolymerization of propylene and norbornene with ansa-fluorenylamidodimethyltitanium complex: Synthesis of novel syndiotactic polypropylene-b-poly(propylene-ran-norbornene). Macromolecules 2006, 39, 2031–2033. [Google Scholar] [CrossRef]
  20. Cai, Z.; Harada, R.; Nakayama, Y.; Shiono, T. Highly active living random copolymerization of norbornene and 1-alkene with ansa-fluorenylamidodimethyltitanium derivative: Substituent effects on fluorenyl ligand. Macromolecules 2010, 43, 4527–4531. [Google Scholar] [CrossRef]
  21. Cai, Z.; Ikeda, T.; Akita, M.; Shiono, T. Substituent effects of tert-butyl groups on flurenyl ligand in syndiospecific living polymerization of propylene with ansa-fluorenylamidodimethyltitanium complex. Macromolecules 2005, 38, 8135–8139. [Google Scholar] [CrossRef]
  22. Cai, Z.; Ohmagari, M.; Nakayama, Y.; Shiono, T. Highly active syndiospecific living polymerization of higher 1-alkene with ansa-fluorenylamidodimethyltitanium complex. Macromol. Rapid Commun. 2009, 30, 1812–1816. [Google Scholar] [CrossRef] [PubMed]
  23. Cai, Z.; Su, H.; Nakayama, Y.; Shiono, T.; Akita, M. Synthesis of C1 symmetrical ansa-cyclopentadienylamidotitanium complexes and their application for living polymerization of propylene. J. Organomet. Chem. 2014, 770, 136–141. [Google Scholar] [CrossRef]
  24. Tanaka, R.; Yanase, C.; Cai, Z. Structure-stereospecificity relationships of propylene polymerization using substituted ansa-silylene(fluorenyl)(amido) titanium complexes. J. Organomet. Chem. 2016, 804, 95–100. [Google Scholar] [CrossRef]
  25. Sheldrick, G.M. SADABS: An Empirical Absorption Correction Program for Area Detector Data; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  26. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  27. Sheldrick, G.M. A short history of SHELX. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  28. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  29. SAINT+, Vdrsion 6.22a, Bruker AXS Inc.: Madison, WI, USA, 2002.
  30. SAINT+, Vdrsion v7.68A, Bruker AXS Inc.: Madison, WI, USA, 2009.
  31. SHELXTL NT/2000, Version 6.1, Bruker AXS Inc.: Madison, WI, USA, 2002.
  32. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  33. Tolman, C.A.; Seidel, W.C.; Gosser, L.W. Formation of three-coordinate Nike(0) complexes by phosphorus ligand dissociation from NiL4. J. Am. Chem. Soc. 1974, 96, 53–60. [Google Scholar] [CrossRef]
Figure 1. Chemical diagrams of the ansa-(fluorenyl)(amido) titanium complexes studied.
Figure 1. Chemical diagrams of the ansa-(fluorenyl)(amido) titanium complexes studied.
Polymers 09 00131 g001
Figure 2. Structure of fluorenylamidotitanium complexes 1a, 1b, 2a, and 2b. Hydrogen atoms are omitted for clarity. Atoms are drawn at the 40% probability level.
Figure 2. Structure of fluorenylamidotitanium complexes 1a, 1b, 2a, and 2b. Hydrogen atoms are omitted for clarity. Atoms are drawn at the 40% probability level.
Polymers 09 00131 g002
Table 1. Crystallographic data and parameters for 1a, 1b, 2a, and 2b.
Table 1. Crystallographic data and parameters for 1a, 1b, 2a, and 2b.
Complex1a1b2a2b
deposition numbersCCDC 1539490CCDC 1539499CCDC 1540563CCDC 1539502
formulaC26H31NSiTiC34H47NSiTiC30H33NSiTiC38H49NSiTi
formula weight433.51545.71483.56595.77
crystal systemMonoclinicTriclinicMonoclinicTriclinic
space groupP1 21/c1P 1 - P1 21/c1P 1 -
a (Å)8.8160(15)10.0609(13)12.742(3)10.3445(13)
b (Å)27.430(5)11.5953(15)17.562(4)11.8424(16)
c (Å)9.9361(18)15.4910(3)11.404(2)15.4920(3)
β(deg)108.372(4)100.124(3)95.269(4)99.822(4)
V3)2280.3(7)1561.1(4)2541.1(9)1636.8(4)
Z4242
F(000)9205881024640
Dcalcd. (g·cm−3)1263116112641209
μ (mm1)0.4400.3340.4020.325
theta range for data collection2.284° to 30.611°2.002° to 30.542°1.980° to 30.772°1.965° to 25.499°
reflections collected23,08715,85625,45111,599
independent reflections7002941379206105
final R indices [I > (I)][R(int) = 0.1147][R(int) = 0.0235][R(int) = 0.0610][R(int) = 0.0569]
R1 = 0.0512,
wR2 = 0.0963
R1 = 0.0398,
wR2 = 0.0940
R1 = 0.0455,
wR2 = 0.1040
R1 = 0.0508,
wR2 = 0.1051
Table 2. Selected bond lengths (Å) and bond angles (°) for related complexes.
Table 2. Selected bond lengths (Å) and bond angles (°) for related complexes.
Parameter1a1b2a2b
Ti(1)–C(1)2.241(2)2.250(14)2.260(19)2.267(3)
Ti(1)–C(2)2.409(3)2.384(14)2.414(19)2.413(3)
Ti(1)–C(3)2.586(3)2.599(14)2.5562.576(3)
Ti(1)–C(4)2.594(3)2.615(15)2.5382.569(3)
Ti(1)–C(5)2.398(2)2.430(15)2.396(19)2.394(3)
Ti(1)–N(1)1.919(2)1.922(12)1.922(17)1.920(2)
Ti(1)–Si(1)2.836(9)2.828(6)2.8502.848(11)
N(1)–Ti(1)–C(1)78.35(9)78.56(5)78.01(7)77.74(11)
Ti(1)–N(1)–Si(1)100.88(10)100.40(6)101.30(8)101.76(12)
N(1)–Si(1)–C(1)93.66(10)94.20(6)93.75(8)93.88(13)
Table 3. Propylene polymerization with Ti complexes 1a2b activated by MMAO a.
Table 3. Propylene polymerization with Ti complexes 1a2b activated by MMAO a.
EntryCat.Temp. (°C)Time (min)Yield (g)Activity bMn c (×104)Mw/Mn cN d (μmol)
11a053.21193025.21.4913
21b053.46208023.91.3714
31c053.59215026.11.5814
41a2554.05243016.82.0524
51b2554.06243019.41.7021
61c2554.21253016.61.9325
72a0120.271695.181.185.2
82b0120.084211.731.164.9
92a25120.4741194.811.439.9
102b25120.362914.381.278.3
a Polymerization conditions: toluene = 30 mL, Ti = 20 μmol, Al/Ti = 600, propylene = 1 atm; b Activity in kg of PP/(mol of Ti·h); c Number-average molecular weight and molecular weight distribution determined by gel-permeation chromatography (GPC) using universal calibration; d Calculated from yield and Mn.
Table 4. Tolman’s cone angle of amido groups of complexes 1a, 1b, 2a, and 2b.
Table 4. Tolman’s cone angle of amido groups of complexes 1a, 1b, 2a, and 2b.
CatalystFluorenylN-alkylθ (°)
1aFluPh86
1b2,7-di-tBu-FluPh90
2aFluNaphthyl103
2b2,7-di-tBu-FluNaphthyl102
Table 5. Steric pentad distributions for samples in Table 3 (entries 4–6 and 10).
Table 5. Steric pentad distributions for samples in Table 3 (entries 4–6 and 10).
EntryStereosequence distribution a
mmmmmmmrrmmrmmrrmmrm + rmrrrmrmrrrrmrrrmrrm
40.040.050.060.220.220.130.090.140.05
50.020.030.040.180.220.110.150.190.06
60.030.050.030.130.200.140.160.210.05
100.120.100.080.290.160.050.030.080.09
a Determined by 13C NMR spectroscopy.

Share and Cite

MDPI and ACS Style

Wang, H.; Wang, X.; Sun, Y.; Cheng, H.; Shiono, T.; Cai, Z. Living Polymerization of Propylene with ansa-Dimethylsilylene(fluorenyl)(cumylamido) Titanium Complexes. Polymers 2017, 9, 131. https://doi.org/10.3390/polym9040131

AMA Style

Wang H, Wang X, Sun Y, Cheng H, Shiono T, Cai Z. Living Polymerization of Propylene with ansa-Dimethylsilylene(fluorenyl)(cumylamido) Titanium Complexes. Polymers. 2017; 9(4):131. https://doi.org/10.3390/polym9040131

Chicago/Turabian Style

Wang, Huajin, Xinwei Wang, Yanjie Sun, Hailong Cheng, Takeshi Shiono, and Zhengguo Cai. 2017. "Living Polymerization of Propylene with ansa-Dimethylsilylene(fluorenyl)(cumylamido) Titanium Complexes" Polymers 9, no. 4: 131. https://doi.org/10.3390/polym9040131

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