Elsevier

Polymer

Volume 51, Issue 5, 2 March 2010, Pages 1088-1099
Polymer

Effects of molecular weight on poly(ω-pentadecalactone) mechanical and thermal properties

https://doi.org/10.1016/j.polymer.2010.01.007Get rights and content

Abstract

A series of poly(ω-pentadecalactone) (PPDL) samples, synthesized by lipase catalysis, were prepared by systematic variation of reaction time and water content. These samples possessed weight-average molecular weights (Mw), determined by multi-angle laser light scattering (MALLS), from 2.5 × 104 to 48.1 × 104. Cold-drawing tensile tests at room temperature of PPDL samples with Mw between 4.5 × 104 and 8.1 × 104 showed a brittle-to-ductile transition. For PPDL with Mw of 8.1 × 104, inter-fibrillar slippage dominates during deformation until fracture. Increasing Mw above 18.9 × 104 resulted in enhanced entanglement network strength and strain-hardening. The high Mw samples also exhibited tough properties with elongation at break about 650% and tensile strength about 60.8 MPa, comparable to linear high density polyethylene (HDPE). Relationships among molecular weight, Young's modulus, stress, strain at yield, melting and crystallization enthalpy (by differential scanning calorimetry, DSC) and crystallinity (from wide-angle X-ray diffraction, WAXD) were correlated for PPDL samples. Similarities and differences of linear HDPE and PPDL molecular weight dependence on their mechanical and thermal properties were also compared.

Introduction

Polyethylene is the most widely used commodity polymer. It is found in many consumer products, such as milk jugs, detergent bottles, margarine tubs, garbage containers, water pipes, just to name a few. Poly(ω-pentadecalactone) (PPDL) is a new type of thermoplastic that can be synthesized by lipase catalysis [1], [2], [3]. The chemical structure of PPDL, with 14 methylene groups and an in-chain ester linkage in each repeating unit, is very similar to that of linear high density polyethylene (HDPE) (Scheme 1). Polyethylene (PE) cannot be easily decomposed into small molecules after usage. To achieve extensive degradation of the PE carbon backbone, treatment of PE with strong oxidized agents such as nitric acid [4], ozone [5] and permanganic acid [6] or pyrolysis at high reaction temperatures (above 370 °C) [7], [8] is required. Therefore, “white pollution” [9], [10] from un-recycled PE plastics is a mounting problem that mankind must confront. An important advantage of PPDL over PE is that the former has ester groups in the backbone that are susceptive to chain breakage. Consequently, gentle enzymatic hydrolysis can, in principle, be used to decompose PPDL back into monomer building blocks. Currently, on-going studies are in progress to find a suitable enzyme for PPDL biological recycling both in our laboratory and elsewhere [11].

Chemical catalysts such as potassium alkoxides [12], diethylzinc [13], [14] and yttrium isopropoxide [15] can be used for conversion of macrolactones to polyesters. For ω-pentadecalactone (PDL), the use of immobilized lipase catalysis has been proven to be superior to chemical catalyzed routes for making polymers, resulting in more rapid polymerization kinetics as well as yielding polyesters of relatively higher molecular weight. Immobilized lipase-catalyzed polymerization of macrolactones was first published by Uyama et al. [16]. Our laboratory demonstrated that, using Novozym 435 that consists of Candida antarctica lipase B (CALB) physically immobilized on a macroporous support, the polymerization process of PDL gives PPDL with number-average molecular weight (Mn) up to 8.6 × 104 in yields exceeding 90% (route is illustrated in Scheme 1) [17]. Thus, enzyme-catalyzed preparation of PPDL with high molecular weight has provided suitable materials for evaluation of their physico-mechanical properties. For example, mechanical properties of PPDL with Mn 6.5 × 104 (polydispersity, Mw/Mn, 2.0) [18] and thiol-functionalized PPDL telechelics [19] have been studied.

Previous crystallographic work [20] indicated that the a and b parameters of PPDL unit cells are slightly larger than those of PE's and the until cell parameter along the fiber axis is much larger than that of PE's. DMA tests showed that PPDL with Mn 6.5 × 104 had a glass transition at −27 °C. Also, the observed high storage modulus was attributed to high crystallinity as determined by DSC and WAXS [18]. PPDL-based copolymers with other monomers including trimethylene carbonate [2], p-dioxanone [3] and ω-caprolactone [21] were synthesized and studied by thermal and X-ray analysis. Copolymers were found to be highly crystalline random copolymers over the entire composition range. This behavior was attributed to co-crystallization of comonomer units in a common lattice or isomorphic substitution of comonomer units.

The current study aimed to investigate the effect of PPDL molecular weight on its mechanical, thermal and rheological properties. The chosen synthetic methods enabled the preparation of PPDL with Mw values up to 48.1 × 104 and PDI values close to 2.0. Films were prepared by press-molding at 130 °C and tensile testing was performed on these samples. Based on the shape of stress–strain curves, a brittle-to-ductile transition along with maximum elongation at break was observed. Since the chemical structure of PPDL is similar to polyethylene (PE), its thermal and mechanical properties were compared with those of a commercially obtained PE sample [22]. In addition, PPDL films with different molecular weights were analyzed by differential scanning calorimetry (DSC), wide-angle X-ray diffraction (WAXD), small-angle X-ray scattering (SAXS) and dynamic mechanical analysis (DMA) to reveal the molecular weight dependence of mechanical, thermal, crystalline and rheological properties. It was found that Young's modulus and stress at break exhibit distinct crystallinity dependence; in addition, differences, of elongation at break and true stress at break, between PPDL and linear PE, are discussed.

Section snippets

Materials

Samples of ω-pentadecalactone (PDL, 98%) and anhydrous p-xylene (>99%) were purchased from Aldrich Chemical Co. and were used as received. Chloroform was purchased from PHARMCO-AAPER Inc. (>99.9%). Anhydrous toluene (98%), purchased from Aldrich Chemical Co., was dried over sodium and then was distilled under nitrogen. Novozym 435 (specific activity 10,000 PLU/g) was a gift from Novozymes (Bagsvaerd, Denmark) and consists of Candida antarctica Lipase B (CALB) physically adsorbed within the

PPDL synthesis

Previous work by our laboratory demonstrated that polymerization of PDL for 2 h at 70 °C in dry toluene (PDL to toluene 1:2 wt/vol) with 10 w/w-% monomer-to-catalyst gave PPDL (without fractionation) with Mn 7900 g/mol (GPC relative to polystyrene) [17]. Preparation of PPDL with higher molecular weights requires increasing diffusion constraints that slow chain propagation and decreasing reaction water content to start fewer chains [21], [24], [25], [26]. To prepare a series of PPDL samples (1–3,

Discussion

This section considers the effects of Mw on PPDL mechanical, thermal and crystallization properties and compares these to those of linear PE.

Conclusions

Synthesis of large-scale PPDL samples was performed by lipase catalysis. Variation of PPDL Mw from 2.5 to 48.1 × 104 was achieved by manipulating reaction variables including reaction water content, method of mixing and reaction time. Tensile testing, DSC, X-ray, DMA and GPC were used to investigate the effect of molecular weight on mechanical, thermal and crystalline material properties. Cold-drawing tensile tests at room temperature revealed a brittle-to-ductile transition for PPDL samples with

Acknowledgements

The authors thank the National Science Foundation (NSF) and industrial members of NSF-I/UCRC for Biocatalysis and Bioprocessing of Macromolecules at NYU/Polytechnic University for financial support, intellectual input, and encouragement during the course of this research. BH also thanks support from NSF (DMR-0906512). We are also grateful to Mr. Seong Chan Park and Dr. Miriam Rafailovich at SUNY at Stony Brook for providing access to their Instron Tensile testing and DMA equipment.

References (48)

  • J.G. Zhang et al.

    Polym Degrad Stab

    (2006)
  • G. Capaccio et al.

    Polymer

    (1975)
  • A. Kumar et al.

    Biomacromolecules

    (2000)
  • W. Wu et al.

    Polymer

    (1992)
  • P.I. Vincent

    Polymer

    (1960)
  • L. Mazzocchetti et al.

    Macromolecules

    (2009)
  • G. Ceccorulli et al.

    Biomacromolecules

    (2005)
  • Z.Z. Jiang et al.

    Biomacromolecules

    (2007)
  • A. Keller et al.

    J Polym Sci Part A-2

    (1971)
  • G.N. Patel et al.

    J Polym Sci Polym Phys Ed

    (1975)
  • S. Haridoss et al.

    J Appl Phys

    (1984)
  • A. Sezgin et al.

    Ind Eng Chem Res

    (1998)
  • C.Q. Fang et al.

    Polym Comp

    (2008)
  • I. Meulen et al.

    Biomacromolecules

    (2008)
  • Z. Jedlinski et al.

    Macromol Chem Phys

    (1996)
  • K.L. Makovetsky

    Vysokomol Soedin Ser B

    (1999)
  • A. Yevstropov et al.

    Vysokomol Soedin Ser A

    (1983)
  • Z. Zhong et al.

    Macromol Chem Phys

    (2000)
  • H. Uyama et al.

    Bull Chem Soc Jpn

    (1995)
  • A. Kumar et al.

    Macromolecules

    (2000)
  • M.L. Focarete et al.

    J Polym Sci Part B Polym Phys

    (2001)
  • N. Simpson et al.

    Macromolecules

    (2008)
  • M. Gazzano et al.

    J Polym Sci Part B Polym Phys

    (2003)
  • Cited by (68)

    View all citing articles on Scopus
    View full text