Elsevier

Polymer

Volume 54, Issue 25, 27 November 2013, Pages 6716-6724
Polymer

Fully-reversible and semi-reversible coordinative chain transfer polymerizations of 1,3-butadiene with neodymium-based catalytic systems

https://doi.org/10.1016/j.polymer.2013.10.031Get rights and content

Abstract

Coordinative chain transfer polymerization (CCTP) of 1,3-butadiene was assessed by employing several traditional Ziegler–Natta type Nd-based catalytic systems. Both the types of alkylaluminum as CTA and chloride donor as third component significantly affected the chain transfer characteristic of the CCTP systems. Among the catalytic systems examined, Nd(OiPr)3/Al(iBu)2H/Me2SiCl2 and Nd(OiPr)3/Al(iBu)2H/Al2Et3Cl3 systems exhibited the highest catalytic efficiency, yielding 6–10 polymer chains per Nd atom in the presence of 20 equiv. CTA. Kinetic examination revealed that Nd(OiPr)3/Al(iBu)2H/Me2SiCl2 and Nd(OiPr)3/Al(iBu)2H/Al2Et3Cl3 catalytic systems proceeded with fully- and semi-reversible chain transfer reactions, respectively. Quantitative formation of polymers was observed in each step of the 1,3-butadiene seeding polymerization, indicating the living mode of the two catalytic systems. Moreover, the triblock copolymers, PBD-b-PIP-b-PBD and PBD-b-PIP-b-PCL, were successfully synthesized with Nd(OiPr)3/Al(iBu)2H/Me2SiCl2 catalytic system.

Introduction

Ziegler–Natta type catalytic systems based on neodymium salts represent an important class of catalysts and have been utilized industrially for 1,3-congjugated dienes polymerization in a large capacity due to their high catalytic activity and excellent stereoselectivity [1], [2], [3]. However, those catalysts are still suffering from the poor controllability over molecular weight, resulting in the polymers with rather broad molecular weight distribution (MWD) [4], [5]. It is reported that high cis-1,4 polydienes coupled with narrow MWD are desirable for high abrasion resistance, low heat buildup and high tensile properties [6], [7]. Living polymerizations [8], [9], [10], including anionic polymerization [11], [12], living coordination polymerization [13], [14], [15], [16], can feasibly realize controlled polymerization of 1,3-conjugated dienes. However, these living polymerization processes produce only one polymer chain per active metal centre, so searching for adaptive approaches to synthesize polydienes with controlled polymerization behaviors more expediently and economically is still imperative.

As an efficient alternative to living polymerization, the new concept of coordinative chain transfer polymerization (CCTP) has gathered an upsurge in research interest recently [17], [18], [19], [20]. Besides sharing some similar features with living polymerization, CCTP preserves various unprecedented advantages such as atom economy, controlling of statistical copolymers composition [21], [22], [23], [24], [25], [26], and design of new polymeric materials via chain shuttling polymerization [18], [27], [28], [29], [30]. CCTP involves the use of a single transition metal-based catalyst and a chain transfer agent (CTA). The heart of CCTP is the highly efficient, rapid and reversible chain transfer reaction between active transition metal-based propagating centers and CTA which is usually in the form of main group metal alkyl such as ZnR2, MgR2, and AlR3 [31], [32]. In comparison with classical living polymerization, chain transfer to CTA allows the growth of several polymer chains per active species which has grabbed much attention from industrial point of view. Arriola et al. has derived a mathematical model and put forward three kinds of simulations in olefin polymerizations, that is, fully reversible chain transfer, semi-reversible chain transfer and irreversible chain transfer [32], [33]. Experiment examples for irreversible chain transfer [34] and reversible chain transfer polymerization [35], [36] suitable for the simulations have been accumulated, however, that for semi-reversible chain transfer polymerization has not been disclosed yet.

Tremendous advances of CCTP have been made in the field of olefin polymerization in the past decade, and innovative features such as well-controlling over the microstructure and architecture of polymers have emerged [37], [38]. The concept of CCTP was introduced to the field of conjugated dienes polymerization only recently and mainly focused on isoprene polymerization. For instance, Cp*La(BH4)2(THF)2/EtMg(nBu) [25], [39], [40], [41], rac-[{Me2C(Ind)2}Y(1,3-(SiMe3)2–C3H3)]/ZnEt2 (Al(iBu)3 or Mg(nBu)2) [26], and bis(allyl)[(C5Me4–C6H4-o-NMe2)Gd (η3-C3H5)2]/AlR3 catalytic systems [42] can serve as efficient catalysts for CCTP of isoprene. In addition to these sophisticated lanthanide complexes based catalyst systems, traditional Ziegler–Natta type neodymium (Nd)-based catalytic systems are developed for the CCTP of isoprene by our group [43], [44]. However, the CCTP of 1,3-butadiene remained less explored. Trans-1,4 stereoselective chain transfer polymerization of butadiene was reported by using Nd(O-2,6-t-Bu2-4-Me-Ph)3(THF)/Mg(n-Hex)2 system [45]. Cis-1,4 stereoselective chain transfer polymerization of butadiene was achieved by using Nd(vers)3 based catalyst systems, however the polymerization rate was highly decreased in transfer conditions [46], [47]. Recently, chain shutting copolymerization of styrene, isoprene, and 1,3-butadiene was reported by using two different scandium catalysts with Al(iBu)3 as chain shuttling agent [48]. Hence, it is necessary to explore new catalytic systems for CCTP of 1,3-butadiene to get a deep understanding of the polymerization mechanism.

In this contribution, we examined the ability of several typical Ziegler–Natta Nd-based catalytic systems for the CCTP of 1,3-butadiene. The effects of various chain transfer agents based on alkylaluminum and chloride donors on the chain transfer ability of catalytic systems were firstly assessed. The kinetic experiments were designed and carried out to examine the polymerization process with different catalytic systems. We report notably herein fully-reversible and semi-reversible coordinative chain transfer polymerizations of 1,3-butadiene, which have never been reported as far as we are aware. Moreover, the living metal-polybutadienyl species could further initiate block copolymerization, affording PBD-b-PIP-b-PBD and PBD-b-PIP-b-PCL triblock copolymers.

Section snippets

Materials

Nd(OiPr)3 was purchased from J&K Chemical Co. and used as a toluene solution (0.10 mol/L). This compound is actually an oxo-isopropoxide cluster of the type [Nd55-O)(μ2-OiPr)43-OiPr)4(OiPr)5] [49] and generally denoted as Nd(OiPr)3 [50], [51]. We added this sentence into the experiment part. Neodymium versatate [Nd(vers)3] was synthesized according to the literature [52] and diluted to 0.1 mol/L by toluene. Alkylaluminum [AlEt3, AlEt2H, Al(iBu)3, Al(iBu)2H, and Al(Oct)3] were commercial

Screening out catalytic systems for 1,3-butadiene polymerization

It is known alkylaluminum and halide donor are essential for 1,3-conjugated dienes polymerizations by using Nd-based Ziegler–Natta catalyst to achieve high activity and cis-1,4 selectivity [55]. It has been proved that the nature of alkylaluminum remarkably influences the chain transfer rate and polymerization rate in the polymerization process [46], consequently, affording polymers with different molecular weight and MWD. While how the halide donor affects polymerization process is not clear

Conclusion

Coordination chain transfer polymerization (CCTP) of butadiene was evaluated by using traditional Nd-based catalytic systems. The catalytic systems, Nd(OiPr)3/Al(iBu)2H/Me2SiCl2 and Nd(OiPr)3/Al(iBu)2H/Al2Et3Cl3, proved to be highly efficient for CCTP of butadiene, yielding 6–10 polymer chains per Nd atom in the presence of 20 equiv. CTA. The kinetic results indicate that Nd(OiPr)3/Al(iBu)2H/Me2SiCl2 system polymerized in fully reversible chain transfer manner, while Nd(OiPr)3/Al(iBu)2H/Al2Et3Cl

Acknowledgments

The authors appreciate financial supports from the National Science and Technology Infrastructure Program (2007BAE14 B01-06), The Fund for National Natural Scientific Foundation of China (No. 20974016 and No. 51203147), The Jilin Provincial Research Fund for Basic Research, China (No. 20130102007JC).

References (63)

  • C.Y. Ren et al.

    Polymer

    (2007)
  • G.H. Kwag et al.

    Polymer

    (2005)
  • G. Ricci et al.

    Coord Chem Rev

    (2010)
  • D.R. Gong et al.

    Polymer

    (2009)
  • F. Wang et al.

    Polymer

    (2012)
  • J. Gromada et al.

    J Organomet Chem

    (2003)
  • W.M. Dong et al.

    Polymer

    (2003)
  • G. Ricci et al.

    J Organomet Chem

    (2005)
  • L. Friebe et al.

    Adv Polym Sci

    (2006)
  • A. Fischbach et al.

    Adv Polym Sci

    (2006)
  • Z.Q. Shen

    Inorg Chim Acta

    (1987)
  • L. Porri et al.

    Macromol Symp

    (1998)
  • Y. Hu et al.

    Macromol Chem Phys

    (2013)
  • T.E. Patten et al.

    Science

    (1996)
  • O.W. Webster

    Science

    (1991)
  • M. Szwarc

    Nature

    (1956)
  • J.J. Benvenuta-Tapia et al.

    J Appl Polym Sci

    (2010)
  • J.A. Lopez et al.

    Macromol React Eng

    (2009)
  • B. Wang et al.

    Macromolecules

    (2008)
  • G.W. Coates et al.

    Angew Chem Int Ed

    (2002)
  • A. Valente et al.

    Chem Rev

    (2013)
  • P. Zinck

    Polym Int

    (2012)
  • R. Kempe

    Chem Eur J

    (2007)
  • J.F. Pelletier et al.

    Angew Chem Int Ed

    (1996)
  • S.B. Amin et al.

    Angew Chem Int Ed

    (2008)
  • L.R. Sita

    Angew Chem Int Ed

    (2009)
  • A. Valente et al.

    Macromol Rapid Commun

    (2009)
  • A. Valente et al.

    J Polym Sci Part A Polym Chem

    (2011)
  • A. Valente et al.

    J Polym Sci Part A: Polym Chem

    (2011)
  • L. Annunziata et al.

    Macromolecules

    (2011)
  • D.J. Arriola et al.

    Science

    (2006)
  • Cited by (0)

    View full text