Investigations into the origins of plastic flow and strain hardening in amorphous glassy polymers

https://doi.org/10.1016/j.ijplas.2013.11.007Get rights and content

Highlights

  • Molecular simulations were used to investigate the origins of plasticity.

  • Unstable expansion with free volume growth, on yield, causes strain softening.

  • Residual plasticity is due to irreversible conformational changes during deformation.

  • Hardening is due to rise in closely contacting chain parts during compression.

  • Torsion state flips govern small strain trend and non-bonded forces govern hardening.

Abstract

In this work, through well designed molecular dynamics (MD) simulations on carefully equilibrated macromolecular ensembles of a generic model of an amorphous glassy polymer, we attempt to understand the origins of plasticity in these materials. In particular, we are interested in the phenomena of post yield strain softening and strain hardening that are typical of these materials. The intrinsic uniaxial stress–strain behaviour is characterized by an elastic part, yield followed by strain-softening and hardening at large strains. Existing constitutive theories of glassy polymers rely on the so called rubber elastic analogy where strain hardening is assumed to be the result of the affine orientation and subsequent stretching of an underlying entanglement network. Using MD simulations and a continuum thermo-mechanical framework, we isolate the plastic free energy associated with the energetic back-stress that governs hardening. We show that the hardening owes its origin to the non-bonded part of the forcefield and is related to the increase in the number of close binary contacts between parts of macromolecular chains that are pushed into close proximity due to compressive deformation. The strain softening is a result of a burst of free volume evolution that accompanies a sudden unstable volume expansion occurring right after yield as the polymer undergoes a deformation induced disordering. This is accompanied by a sharp increase in the room available for changes in dihedral conformations. These conformational changes are mainly responsible for accommodating the plastic deformation as many of them are unable to flip back to their original conformations even after unloading. Finally, by fitting the uniaxial responses obtained from the MD simulations with a standard constitutive model, we show that the barriers between the torsional states govern the small strain behaviour while the strength of the non-bonded interactions govern hardening.

Introduction

With increasing strain, the constant strain rate uniaxial compressive (or intrinsic) stress strain response of most amorphous glassy polymers (e.g., Hasan and Boyce, 1993 for polystyrene, PS; G’Sell et al., 1992 for polycarbonate, PC; Tordjeman et al., 1997 for polymethyl methacrylate, PMMA) exhibits a linear elastic region, prominent strain softening after yield, almost perfect plasticity and finally steep hardening. The extent of softening (also called the ‘yield drop’) at small strains as well as the hardening at large strains depend on the macromolecular architecture. The extent of the yield drop as well as the hardening is technologically important as the former leads to localisation of strains into thin bands while the latter stabilises them (Lai and Van der Giessen, 1997). For instance, intrinsic response of PC shows a small yield drop followed by a steep hardening while PS shows a sharper yield drop followed by a milder hardening (van Melick et al., 2003) leading to significantly different deformation and fracture behaviour in these two materials. Moreover, yield drop and hardening also depend on the applied strain rate, hydrostatic stress and test temperature. The rate, temperature and pressure induced subtleties in the yield drop and hardening responses have been well characterized for almost all important polymers. However, the mechanistic origins of plasticity in these materials in general and the reasons for the yield drop and hardening in particular, are not yet clear. We start by discussing the major points of view that presently exist.

Attempts at providing mechanistic explanations for the elasto-plastic behaviour of these materials date back to Ree and Eyring, 1955, Robertson, 1966. The former, based on the idea that yielding is achieved by surmounting the activation energy barrier for the sliding of a segment of a macromolecular chain over others, is able to describe the variation of the yield stress with temperature for a number of polymers (Bauwens-Crowet et al., 1969). The Robertson (1966) model, on the other hand, connects yielding to activation of conformational jumps on the main chain in parts of the macromolecules. In-situ Fourier transform infrared experiments on thin films of PS have shown that (Theodorou and Suter, 1985, Xu et al., 1989) there indeed is a sharp increase in the number fraction of certain conformers around the yield point. More recent investigations have pointed out that yielding in a range of amorphous polymers seem to be related to the secondary β relaxations. In fact, many of the molecular motions that are responsible for the β relaxations in commercial polymers have been identified (Rana et al., 2005). It has been postulated (van Breemen et al., 2012) that the yield drop results from an interplay between molecular motions related to both the α and β transitions and correspondingly involves two different activation energies.

Alternate to the Eyring approach, another class of models seek to explain yield and strain softening behaviour of polymers. Interestingly, the yield drop in amorphous polymers is a feature of the stress strain response that is also seen in a metallic glasses (Argon, 1979). In metallic glasses, the carriers of plasticity are believed to be the so-called ‘shear transformation zones’ (STZ) which are clusters of atoms that undergo intense rearrangement under the application of stress. These regions have atomic environments with liquid-like character (Cohen and Grest (1981)) and in case of most metallic glasses, are about 30nm3 in size (Argon, 2013). The idea has been extended by Argon and co-workers to the small strain response of amorphous polymers where, the volume of the shear transformation zone is significantly larger. Using the idea that plasticity in polymers proceeds by the proliferation of liquid like STZs, Argon (2013) was able to model the yield and the yield drop for annealed and quenched PS. The underlying parameters in the model for the compressive stress variation with plastic strain involves a measure of the STZ size which is about 4 orders of magnitude larger than in case of metallic glasses (Argon and Bessonov, 1997). Also, the liquid like zones make up around 50 % of the volume of the polymer when it reaches the perfectly plastic state. Monte Carlo simulations by Mott et al., 1993, Hutnik et al., 1993 lend credence to this idea. They have shown that deformation of small macromolecular ensembles proceeds through sudden bursts of large shear activity involving a small number of segments in a chain. Experimental studies of small strain deformation of a number of linear polymers by Shenogin et al. (2004) provide evidence of the nucleation, growth and coalescence of STZs around the yield point. The proliferation of STZs is related to the increase in free volume that provides room for the drastic transformations to take place. The size and distribution of free volume within a glassy polymer sample has been determined through positron annihilation lifetime spectroscopy (Kobayashi et al., 1994, Cangialosi et al., 2005). A model for rate dependent yield and strain softening can also be formulated from the knowledge of the distribution of free volume within the polymer coupled with the assumption that it is distributed in the same manner as the local strain (Spathis, 2008).

A number of mechanistic explanations for the hardening of glassy polymers at large strains have also been proposed. Many of these theories are based on the fact that polymer chains are held together by unbreakable topological constraints called entanglements which are ‘frozen in’ when the polymer is cooled from the melt state. The hardening is believed to be a result of the segments between entanglements being pulled taut by the imposed stretch, which results in the increase in free energy due to reduction in configurational entropy. Since the idea is similar to that employed in rubber elasticity where the cross-links play the role of the unbreakable entanglements, we will henceforth refer to this as the ‘rubber elastic analogy’ to hardening. The rubber elastic analogy (e.g. Haward and Thackray, 1968, Boyce et al., 1988, Wu and Van der Giessen, 1993, Ames et al., 2009) derives rigorously from the fact that the segments sample the Langevin or Gaussian distribution of end-to-end distances as they deform and has been successful in fitting the hardening part of the uniaxial response. However, completely phenomenological descriptions of hardening (e.g. Gent, 1996) are also known to provide equally good fits. Such phenomenological models for strain hardening (as opposed to models that are based on the rubber elastic analogy) have been successfully used in recent constitutive frameworks (Anand et al., 2009, Ames et al., 2009). Also, Li and Buckley, 2010 have suggested that hardening in polymers can be modelled through stretch induced anisotropic viscoplasticity in small flow units dispersed within the polymer volume.

A number of issues have been pointed out in connection with the rubber elastic analogy. Two important parameters that enter the model for hardening are N and CR which are the average number of ‘links’ making up a segment between two adjacent entanglements and the hardening modulus. The latter is given as CR=nkBT, where n is the volume density of links, kB the Boltzmann constant and T the temperature. Recent MD simulations on macromolecular systems with various degrees of atomistic detail (Hoy and Robbins, 2007, Hoy and Robbins, 2008, Mahajan and Basu, 2010) have pointed out that, if a link is identified with a monomer, N (which now becomes the ‘entanglement length’ or the number of monomeric units between adjacent entanglements) turns out to be lower than that obtained from measurements of plateau modulus of the corresponding melt. In other words, the number of links between entanglements N is not the same as the entanglement length Ne measured for the corresponding melt. The hardening modulus required to obtain a good fit is also much lower than what is obtained from realistic entanglement densities. Moreover, the modulus CR is proportional to temperature whereas experiments (e.g. van Melick et al., 2003) show that they decrease with temperature. Recent simulations and experiments have pointed to a correlation between the hardening modulus and the yield stress (Govaert et al., 2008, Robbins and Hoy, 2009) with varying strain rate and temperature, giving rise to the possibility that the underlying micro-mechanics for both are not different. Lastly, in cross-linked rubber, intermolecular interactions are very low and the segments are able to sample all possible orientations in phase space. In a glass at T<Tg, it is uncertain how thermal activation can provide sufficient impetus to the segments to sample all possible configurations. Moreover, simulations show that end to end vectors of segments between entanglements do not flow affinely with the imposed macroscopic deformation, contrary to the basic assumption in rubber elasticity (Mahajan and Basu, 2010).

Counter arguments to the above criticisms of the rubber elastic analogy should also be mentioned at this point. Firstly, it has been argued that a link should not be identified with a single monomer. Rather, in the glassy state, a link is constituted of a large number of monomers (somewhat akin to a Kuhn segment). This means that N is indeed expected to be smaller than the entanglement length. The definition of a link is then unclear and so is the exact interpretation of N. Secondly, following Raha and Bowden (1972), the entanglement length is expected to increase with temperature in a way that the total number of links nN remains the same (see, Basu and Van der Giessen, 2002). Thus, if the increase in temperature is overwhelmed by the decrease in n, the modulus CR may decrease with temperature. Thermo-mechanical simulations of uniaxial deformation have shown that this is possible (see, Basu and Van der Giessen, 2002). Atomistic simulations however, have not thrown up evidence of significant changes in entanglement length with temperature. Loo et al., 2000, Lee and Ediger, 2010, using different experimental techniques, have shown that there is a significant rise in mobility of amorphous chains induced by plastic deformation. It can be argued (Argon, 2013) that the increased mobility aids the free sampling of the phase space by the segments. However, while the experiments have measured mobility close to the glass transition temperature Tg, deformation induced enhancement of mobility deep into the glassy regime seems unlikely.

Atomistic simulations of macromolecular ensembles have, in general, played a significant role in elucidating various aspects of plastic behaviour in glassy amorphous polymers. Methods for creating large macromolecular ensembles with acceptable short range and long range equilibration are available (Auhl et al., 2003, Mahajan and Basu, 2010). Simulations on amorphous polymers have been performed with detailed atomistic models (e.g. Lyulin and Michels, 2006), united atom models with linear chains of polyethylene (PE)-like beads (e.g. Brown and Clarke, 1991, Mahajan and Basu, 2010) as well as coarse grained bead spring models (e.g. Rottler and Robbins, 2003, Hoy and Robbins, 2007). Simulations with all levels of detail have shed light on various aspects of deformation (Capaldi et al., 2002, Vorselaars et al., 2009, Mahajan and Basu, 2010), fracture (Rottler and Robbins, 2003, Kulmi and Basu, 2006) ageing/rejuvenation (Warren and Rottler, 2008, Mahajan et al., 2010) and glass transition (Negi and Basu, 2006) in these materials. In this paper, we focus on the deformation of polymers in the glassy state and in particular, attempt to elucidate processes that govern the micro-mechanics of deformation at different levels of strain. We have conducted a series of MD simulations on carefully equilibrated amorphous macromolecular samples of PE-like chains in order to understand the interplay between the underlying forcefield and the uniaxial compressive response. From a practical point of view, it is useful to understand this interaction as this provides a route to designing polymers with targeted mechanical properties. It has been shown by Majumder et al. (2010) that detailed atomistic models of linear polymers can be coarse grained into PE-like chains with a suitably calibrated forcefield.

Moreover, while advances in characterization of the intrinsic uniaxial response of these materials have enabled formulation of sophisticated general purpose phenomenological constitutive models (e.g. Haward and Thackray, 1968, Boyce et al., 1988, Wu and Van der Giessen, 1993, Ames et al., 2009, Anand and Gurtin, 2003), most of these need to use a large number of fit parameters to negotiate various subtle aspects of the stress strain response. In fact, increase in the fidelity of the constitutive model is generally achieved at the cost of adding parameters to it. Many of the parameters are obtained by fitting to the experimental data and are without a strong physical significance. A better understanding of the micromechanics of plastic flow in glassy amorphous polymers seems to present a possible remedy to the situation. Enriching continuum simulations of these materials with information obtained from atomistic ones is of paramount importance because the intimate connection between macromolecular architecture and all technologically important aspects of mechanical behaviour has been unequivocally established experimentally.

Section snippets

Computational methodology

For most simulations reported, we have used a dense polymeric system consisting of linear entangled macromolecular chains, with 160 linear chains and 200 united atoms per chain. Though we do not intend to simulate a particular polymer, inter-atomic force field for polyethylene (PE) based on an united atom model is used (Fukuda and Kuwajima, 1997). However, as shown by Majumder et al. (2010) for polystyrene, any other linear polymer can be coarse grained into a model governed by similar

Determining the entanglement structure

We intend to look at the evolution of the entanglement network as a function of the imposed uniaxial strain. To this end we use the Primitive Path (PP) analysis (Sukumaran et al., 2005) to quantify the entanglement length Ne. By definition, the PP is shortest path that a macromolecular chain of a given end to end distance can take through a maze of obstacles without crossing any of them. In simulations, the PP of all chains in the ensemble are calculated by first fixing all chain ends. All

Thermomechanics of glassy polymers

In this section, we briefly dwell on the thermomechanical framework for a solid deforming under isothermal conditions. We closely follow the developments presented in Basu and Van der Giessen, 2002, Anand and Gurtin, 2003. Further details of the procedure for extracting the plastic part of the free energy from an MD simulation of a uniaxially deforming solid is elaborated in an earlier work (Mahajan and Basu, 2010).

Consider a body occupying a volume V0 in its reference and V in its current

Hardening behaviour of glassy polymers

In this section we focus on the origins of the hardening behaviour of glassy polymers at large strains. Using the formulation presented in the previous section, we first derive, from carefully designed MD simulations of uniaxial tension and compression, estimates of ψp. Through this exercise we are able to establish that the behaviour of ψp with strain is not as expected from Eq. (36) or (37). In fact, at large strains, ψp is much smaller than what is expected from the rubber elastic analogy.

The onset of plasticity

In this section, we discuss the onset of plastic flow in glassy polymers. As shown in Fig. 2(a), the uniaxial simulations capture the yield drop quite realistically. Also, as was mentioned in the Introduction, the increase in the number of liquid like STZs are believed to be responsible for the drop in stress carrying capacity of the material following yield. The STZs in case of polymers are large and probably bigger than the periodic volume that we have chosen. In spite of the small size, MD

Origins of the plastic strain

The abrupt increase in the room for changes in dihedral conformations is the reason for the onset of plasticity in glassy polymers. The imposed macroscopic deformation is accommodated mainly by dihedral flips. Additionally, at large strains, increase in the number of close binary contacts between chains lead to hardening. Changes in bond stretching and bending energies during deformation are much smaller in comparison. In compression at least, bond lengths and angles do not move into

Lessons for constitutive modeling

Connections between parameters in a typical constitutive model and the underlying forcefield parameters begin to emerge from the above discussions. A typical compressive stress strain curve is shown in Fig. 8. A typical constitutive model uses a number of parameters to characterize the intrinsic uniaxial response. Using the model due to Wu and Van der Giessen (1993), we have indicated in Fig. 8, the different parameters in the model that govern different aspects of the curve. Since we will be

Conclusions

We have performed carefully designed MD simulations with a view to understand the origins of plasticity in glassy amorphous polymers. In particular, we have focused on the origins of hardening at large strains and the phenomenon of yield drop right after yield.

The onset of plasticity in glassy polymers is an abrupt transition from a near-equilibrium elastic state to one in which stress levels are high enough for rapid and rather drastic structural changes to occur. Under uniaxial compression,

References (63)

  • J. Lai et al.

    A numerical study of crack-tip plasticity in glassy polymers

    Mech. Mater.

    (1997)
  • H.X. Li et al.

    Necking in glassy polymers: effects of intrinsic anisotropy and structural evolution kinetics in their viscoplastic flow

    Int. J. Plast.

    (2010)
  • A.V. Lyulin et al.

    Simulation of polymer glasses: from segmental dynamics to bulk mechanics

    J. Non-Cryst. Solids

    (2006)
  • D. Mahajan et al.

    Ageing and rejuvenation in glassy amorphous polymers

    J. Mech. Phy. Solids

    (2010)
  • S.J. Plimpton

    Fast parallel algorithms for short-range molecular dynamics

    J. Comput. Phys.

    (1995)
  • S. Raha et al.

    Birefringence of plastically deformed poly(methyl methacrylate)

    Polym.

    (1972)
  • H.G.H. van Melick et al.

    On the origin of strain hardening in glassy polymers

    Polym.

    (2003)
  • P.D. Wu et al.

    On improved network models for rubber elasticity and their applications to orientation hardening in glassy polymers

    J. Mech. Phys. Solids

    (1993)
  • A.S. Argon

    A theory for low temperature plastic deformation of glassy polymers

    Philos. Mag.

    (1973)
  • A.S. Argon

    Physics of Deformation and Fracture of Polymers

    (2013)
  • A.S. Argon et al.

    Plastic deformation in polyimides with new implications on the theory of plastic flow of glassy polymers

    Philos. Mag.

    (1997)
  • R. Auhl et al.

    Equilibration of long chain polymer melts in computer simulations

    J. Chem. Phys.

    (2003)
  • C. Bauwens-Crowet et al.

    Tensile yield-stress behavior of glassy polymers

    J. Polym. Sci. A-2 Polym. Phys.

    (1969)
  • D. Brown et al.

    Molecular dynamics simulation of an amorphous polymer under tension. 1. Phenomenology

    Macromolecules

    (1991)
  • D. Cangialosi et al.

    Relaxation of free volume in polycarbonate and polystyrene studied by positron annihilation lifetime spectroscopy

    Acta. Phys. Pol. A.

    (2005)
  • F.M. Capaldi et al.

    Enhanced mobility accompanies the active deformation of a glassy amorphous polymer

    Phys. Rev. Lett.

    (2002)
  • H.M. Cohen et al.

    Dispersion of relaxation rates in dense liquids and glasses

    Phys. Rev. B

    (1981)
  • D. Frenkel et al.

    Understanding Molecular Simulation

    (2002)
  • M. Fukuda et al.

    Molecular dynamics simulation of water diffusion in atactic and amorphous isotactic polypropylene

    J. Chem. Phys.

    (1997)
  • A.N. Gent

    A new constitutive relation for rubber

    Rubber Chem. Technol.

    (1996)
  • L.E. Govaert et al.

    Does the strain hardening modulus of glassy polymers scale with the flow stress?

    J. Polym. Sci. B: Polym. Phys.

    (2008)
  • Cited by (0)

    View full text