Elsevier

Polymer

Volume 194, 24 April 2020, 122374
Polymer

Synthesis of zinc complexes bearing pyridine derivatives and their application of ε-caprolactone and L-Lactide polymerization

https://doi.org/10.1016/j.polymer.2020.122374Get rights and content

Highlights

  • The polypyridne-zinc complexes were synthesized.

  • 1,10-Phenanthroline derivatives enhanced the catalytic activity of Zn complexes.

  • Hydrogens of polypyridine co-activate the carbonyl group of monomers.

Abstract

A series of pyridine and polypyridine-zinc complexes were synthesized, and their application in ε-caprolactone and l-lactide polymerization was studied. In ε-caprolactone polymerization, zinc complexes bearing 1,10-phenanthroline derivatives exhibited a higher polymerization rate than did diethyl zinc (2–10 times). In l-lactide polymerization, zinc complexes bearing 1,10-phenanthroline derivatives exhibited a considerably higher polymerization rate than did diethyl zinc (2–220 times). According to density functional theory results, the ortho-hydrogens of polypyridine ligand can co-activate the carbonyl group of monomers with zinc metal center to increase the catalytic activity of zinc catalysts.

Introduction

Biodegradable polymers, such as polylactide (PLA) and poly-ε-caprolactone (PCL), are used as replacements for petrochemical plastics to mitigate the plastic pollution caused after discarding non-degradable petrochemicals. In addition, PLA [[1], [2], [3], [4], [5], [6], [7], [8], [9], [10]] and PCL [7,9,[11], [12], [13], [14], [15], [16], [17], [18], [19]] and are also regarded as biomaterials with various applications because of their biodegradability, biocompatibility, and permeability [20,21]. Ring-opening polymerization (ROP) is useful method in which metal complexes [[22], [23], [24], [25], [26], [27], [28], [29]] are used as catalysts. This method exhibits higher controllability for producing polymer than using traditional polycondensation [30,31].

Zinc, the 24th most abundant element in Earth's crust, has normal oxidation state of +2 which yield to an excellent Lewis acid and is suitable to be used as a catalyst [[32], [33], [34], [35], [36], [37], [38], [39], [40], [41], [42], [43], [44]]. In Zn catalyst researches, the use of binary catalyst systems for ROP is quite common in the literature [[45], [46], [47], [48]]. Specifically, combining a metal alkyl complex with an alcohol (iso-propanol or benzyl alcohol) in an alcoholysis reaction is a common approach for raising initiating ligand nucleophilicity and as a result ROP activity. However, others types of Zn catalysts were also reported. Herres-Pawlis [[33], [34], [35], [36], [37], [38], [39], [40]] have reported several Zn complexes with neutral nitrogen-donor ligands and two monoanionic ligands for ε-caprolactone (CL) and l-lactide (LA) polymerization. However, these Zn complexes with two monoanionic ligands revealed very low reactivity in ROP at condition of 130–150 °C in bulk and reaction for 20–48 h as shown in Fig. 1.

The low reactivity of these Zn complexes are most likely due to halides or acetates are not good nucleophiles, and, therefore, lower their initiating polymerization ability. The ROP process is a two-step process (Fig. 2). The first step is the initiation (Fig. 2, (a)–(b)), and the second step is propagation (Fig. 2, (c)–(d)) of the second monomer attacked by the metal-alkoxy group formed in the initiation step. Although the monomer can be coordinated to a metal center, a low nucleophilic initiator cannot open the ring easily (Fig. 2, (b)). In addition, if the initiation rate is lower than the propagation rate, the molar mass of polymers cannot be controlled using the ratio of monomer/initiator, and their dispersity (Đ) values are broad as observed by Herres-Pawlis (Fig. 1).

To solve this problem, Jeong [[41], [42], [43], [44]] introduced methyl lithium as an alkyl-transferred reagent in the ROP process and assumed that the dimethyl Zn complex was produced in situ. The methyl groups functioned as initiators and were stronger nucleophiles than halides or acetates. The polymerization results in Fig. 3 revealed that the LA polymerization can be conducted at 25 °C, and the polymerization time is 12 h with 97%–99% conversion. In addition, the narrow polydispersity index (Đ) values (1.23–1.27) confirmed the high controllability of these Zn complex and LiMe. Though the addition of LiMe greatly increases the ROP performance, LiMe is such a strong base that it may deprotonate the ligands and/or attack the imino group to produce unexpected byproducts. To reduce the damages caused by LiMe, LiOCHMe2 [[49], [50], [51]] was incorporated instead of LiMe. As shown in Fig. 3b, the polymerization time has been significantly reduced.

Based on aforementioned reports, we believe dialkyl coordinated zinc complexes will outperform those in produced analogies, and, therefore, designed and synthesized a series of diethyl Zn complexes bearing various pyridine, bipyridine, and 1,10-phenanthroline derivatives as depicted in Fig. 4. These Zn complexes were used as catalysts with benzyl alcohol as an initiator to study the effect of ligands in LA and CL polymerization. In addition to the polymerization results, density functional theory-based calculations enable us to have deeper insight to the mechanism of the ROP reactions.

Section snippets

Synthesis and characterization of Zn complexes

All ligands reacted with a stoichiometric amount of diethyl zinc in toluene to produce Zn compounds (Fig. 4). Their formula and structures were confirmed through 1H and 13C Nuclear magnetic resonance (NMR spectra), elemental analysis, and X-ray crystal analysis. For LPyZnEt2, the 1H NMR spectrum (Fig. S39A) revealed that ortho-Hs of pyridine were at 8.25 ppm, and the ratio of pyridine to ethyl groups was 1:2. Compared with 1H NMR spectra of LPy2ZnEt2 (Fig. S39B, ortho-Hs at 8.39, the ratio of

Free energies of intermediates and transition states

The free energies of all intermediates and transition states studied in this DFT research are plotted in Fig. 14. The right-hand side reflects the energies of CL polymerization; the left-hand side illustrates the energies of LA polymerization. The reason why some intermediates (LA-II-A, CL-III and CL-IV-B) are not in the valleys of the plot is that some trivial transient structures between two intermediates are not shown here.

The plot indicated that after accepting CL and LA, the free energies

Conclusions

A series of Zn complexes bearing pyridine, bipyridine, and 1,10-phenanthroline derivatives were synthesized, and their application of CL and LA was studied. In CL polymerization, bipyridine, and 1,10-phenanthroline derivatives enhanced the catalytic activity of Zn complexes compared with diethyl zinc, with the exception of LOMedpy and Lo-Mephen, but reduced the controllability for producing PCLs. In LA polymerization, all ligands enhanced the catalytic activity of Zn complexes compared with

Experimental section

Standard Schlenk techniques and a N2-filled glovebox were used throughout the isolation and handling of all the compounds. Solvents, ε-caprolactone, and deuterated solvents were purified prior to use. Deuterated chloroform, Deuterated benzene, diethyl zinc, ε-caprolactone, and l-lactide were purchased from Acros. Benzyl alcohol, pyridine, 2,2′-bipyridine, 6,6′-dimethyl-2,2′-bipyridine, 4,4′-di-tert-butyl-2,2′-bipyridine, 2,6-di(2-pyridyl)pyridine, 4,4′-di-methyl-2,2′-bipyridine,

CRediT authorship contribution statement

Guan-Jhen Chen: Data curation, Formal analysis. Shi-Xian Zeng: Data curation, Formal analysis. Cheng-Hsun Lee: Data curation. Yu-Lun Chang: Software. Chun-Juei Chang: Validation. Shangwu Ding: Writing - review & editing. Hsuan-Ying Chen: Conceptualization, Writing - original draft. Kuo-Hui Wu: Writing - review & editing. I-Jy Chang: Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgment

This study is supported by Kaohsiung Medical University (KMU-DK109004), NSYSU-KMU JOINT RESEARCH PROJECT, (#NSYSUKMU 107-P010), and the Ministry of Science and Technology (Grant MOST 107-2113-M-037-001 and 107-2113-M-003-005). We thank Center for Research Resources and Development at Kaohsiung Medical University for the instrumentation and equipment support.

References (78)

  • S. Nayab et al.

    Synthesis and X-ray crystal structure of dichloro[S-1-phenyl-N-(S-pyrrolidin-2-ylmethyl)ethanamine]zinc(II) and its catalytic application to rac-lactide polymerization

    Polyhedron

    (2011)
  • S. Nayab et al.

    Zinc complexes bearing N,N′-bidentate entiopure ligands: synthesis, structure and catalytic activity toward ring opening polymerisation of rac-lactide

    Polyhedron

    (2012)
  • S. Nayab et al.

    Synthesis, characterization and X-ray structures of enantiopure-diamine zinc(II) complexes as pre-catalysts for stereoselective production of poly(rac-lactide)

    Polyhedron

    (2013)
  • K.S. Kwon et al.

    Synthesis and structural characterisation of zinc complexes bearing furanylmethyl and thiophenylmethyl derivatives of (R,R)-1,2-diaminocyclohexanes for stereoselective polymerisation of poly(rac-lactide)

    Polyhedron

    (2014)
  • K.S. Kwon et al.

    Synthesis, characterisation and X-ray structures of zinc(II) complexes bearing camphor-based iminopyridines as pre-catalysts for heterotactic-enriched polylactide from rac-lactide

    Polyhedron

    (2015)
  • M.-S. Kang et al.

    Synthesis and characterization of Zn(II) and Cu(II) complexes bearing (chiral substituent)(diethyl)-ethanediamine derivatives as precatalysts for rac-lactide polymerisation

    Polyhedron

    (2019)
  • H.-Y. Chen et al.

    Comparative study of lactide polymerization by zinc alkoxide complexes with a β-diketiminato ligand bearing different substituents

    J. Mol. Catal. Chem.

    (2011)
  • A. Kronast et al.

    Electron-deficient β-diiminato-zinc-ethyl complexes: synthesis, structure, and reactivity in ring-opening polymerization of lactones

    Organometallics

    (2016)
  • W. Brüser et al.

    A comparative IR/Raman, X-ray and computational study of diethylzinc pyridine complexes

    J. Organomet. Chem.

    (2016)
  • Z. Ge et al.

    Functional block copolymer assemblies responsive to tumor and intracellular microenvironments for site-specific drug delivery and enhanced imaging performance

    Chem. Soc. Rev.

    (2013)
  • T. Iwata

    Biodegradable and bio-based polymers: future prospects of eco-friendly plastics

    Angew. Chem. Int. Ed.

    (2015)
  • J. Nicolas et al.

    Design, functionalization strategies and biomedical applications of targeted biodegradable/biocompatible polymer-based nanocarriers for drug delivery

    Chem. Soc. Rev.

    (2013)
  • X. Ren et al.

    Surface modification and endothelialization of biomaterials as potential scaffolds for vascular tissue engineering applications

    Chem. Soc. Rev.

    (2015)
  • R.L. Simpson et al.

    Development of a 95/5 poly(L-lactide-co-glycolide)/hydroxylapatite and beta-tricalcium phosphate scaffold as bone replacement material via selective laser sintering

    J. Biomed. Mater. Res. B Appl. Biomater.

    (2008)
  • R. Sridhar et al.

    Electrosprayed nanoparticles and electrospun nanofibers based on natural materials: applications in tissue regeneration, drug delivery and pharmaceuticals

    Chem. Soc. Rev.

    (2015)
  • O. Castano et al.

    Angiogenesis in bone regeneration: tailored calcium release in hybrid fibrous scaffolds

    ACS Appl. Mater. Interfaces

    (2014)
  • K.S. Egorova et al.

    Biological activity of ionic liquids and their application in pharmaceutics and medicine

    Chem. Rev.

    (2017)
  • Y. Habibi

    Key advances in the chemical modification of nanocelluloses

    Chem. Soc. Rev.

    (2014)
  • K. Liu et al.

    Carbohydrate-based amphiphilic nano delivery systems for cancer therapy

    Nanoscale

    (2016)
  • C.G. Palivan et al.

    Bioinspired polymer vesicles and membranes for biological and medical applications

    Chem. Soc. Rev.

    (2016)
  • J.Y. Park et al.

    3D printed structures for delivery of biomolecules and cells: tissue repair and regeneration

    J. Mater. Chem. B

    (2016)
  • S. Tarafder et al.

    Polycaprolactone-coated 3D printed tricalcium phosphate scaffolds for bone tissue engineering: in vitro alendronate release behavior and local delivery effect on in vivo osteogenesis

    ACS Appl. Mater. Interfaces

    (2014)
  • M.G. Yeo et al.

    Preparation and characterization of 3D composite scaffolds based on rapid-prototyped PCL/β-TCP struts and electrospun PCL coated with collagen and HA for bone regeneration

    Chem. Mater.

    (2011)
  • N. Kamaly et al.

    Degradable controlled-release polymers and polymeric nanoparticles: mechanisms of controlling drug release

    Chem. Rev.

    (2016)
  • C.L. Salgado et al.

    Biocompatibility and biodegradation of polycaprolactone-sebacic acid blended gels

    J. Biomed. Mater. Res.

    (2012)
  • T. Fuoco et al.

    Aluminum alkyl complexes bearing salicylaldiminato ligands: versatile initiators in the ring-opening polymerization of cyclic esters

    Catalysts

    (2017)
  • S.M. Guillaume et al.

    Beyond stereoselectivity, switchable catalysis: some of the last frontier challenges in ring-opening polymerization of cyclic esters

    Chem. Eur J.

    (2015)
  • B.H. Huang et al.

    Metal complexes containing nitrogen-heterocycle based aryloxide or arylamido derivatives as discrete catalysts for ring-opening polymerization of cyclic esters

    Dalton Trans.

    (2016)
  • C. Redshaw

    Use of metal catalysts bearing schiff base macrocycles for the ring opening polymerization (ROP) of cyclic esters

    Catalysts

    (2017)
  • Cited by (0)

    1

    These authors contributed equally.

    View full text