atomistic model of reptation at polymer interfaces · 2019. 9. 9. · properties of the polymer...

1
Atomistic model of reptation at polymer interfaces D.G. Luchinsky 1,2 , H. Hafiychuk 1 , M. Barabash 2 , V. Hafiychuk 1 , K. Chaki 3 , H. Nitta 3 , T. Ozawa 3 , K. R. Wheeler 4 , P.V.E. McClintock 2 1 SGT Inc., ARC, Moffett Field, California, USA, 2 Department of Physics, Lancaster University, UK, 3 Materials Science Section, Engineering Technology Division, JSOL Corp, JP, 4 Ames Research Center, Moffett Field, California, USA Abstract We study a molecular dynamics model of a polymer-polymer interface for a polyetherim- ide/polycarbonate (PEI/PC) blend, including its thermodynamic properties, its chain reptation, and its corresponding welding characteristics. The strength of the sample is analyzed by measuring strain-stress curves in simulations of uni-axial elon- gation. The work is motivated by potential applica- tions to 3D manufacturing in space. Introduction Understanding properties of polymer-polymer interfaces is a long- standing problem of fundamental and technological importance [1]. Polymer chains behavior such as reptation and entanglement deter- mine the welding dynamics and strength in 3D printing [2]. The atomistic structure of polymers substantially influences these prop- erties. However, earlier research was mostly focused either on bulk properties [3] or on coarse-grained models of interfaces [4]. Figure 1:Two amorphous cells (gray and cyan) of PEI/PC polymer blend are combined in one chemical sample across an atomically flat interface. Here we develop fully atomistic model of the polymer-polymer inter- face in PEI/PC blends. We asses MD capabilities in resolving poly- mer dynamics at polymer-polymer interfaces with atomic resolution and analyze thermal and mechanical properties of the blends in the presence of planar interface. We investigate the chains diffusion at the interface and the corresponding relaxation of the interface entanglement and non-bonding energy to the bulk state. Model We have prepared amorphous cells of these blends using the soft- ware package J-OCTA [5]. Cells were relaxed using the following the procedure [5]: (i) equilibration in the NVE at a given temperature T 0 ; (ii) compression in the NPT Andersen-Nose-Hoover thermostat at T 0 and P = 100 MPa; (iii) additional equilibration in the NVE equilibration at T 0 ; (iv) relaxation in the NPT ensemble at T 0 and P 0 ; and (v) elimination of the translational component of the ve- locity in the NVT Nose-Hoover thermostat at T 0 and zero pressure. Each stage was computed during 100 ps with time step 1 fs. The temperature and the pressure during equilibration were T 0 = 600 K and P 0 =1 atm (except for the 2-nd step, where P = 100 MPa.) The atomistically flat surfaces shown in Fig. 1 were prepared by moving polymer chains inside the cells and using LJ-wall. The parameters of the LJ-wall at the interface were: (i) cut-off 10 Å; (ii) sigma 5 Å; (iii) epsilon 10 kJ/mol ; and (iv) density 20 g/cm 3 . Validation To validate the model we we compared MD predictions with the fol- lowing experimentally measured properties: (i) density ρ; (ii) glass transition temperature T g ; (iii) bulk modulus B ; (iv) coefficient of thermal expansion α P ; (v) IR spectra and density of states S (v ); and (vi) specific heat capacity c p . Figure 2:Bulk modulus B estimated using fluctuations (red open diamonds) in comparison with experimental data (teal circles). In the inset: density ρ vs T obtained in MD calculations as compared to the experimental data (teal circles). Arrows show the glass transition temperatures T g obtained in experiment and simulations at the intersections of the straight lines corresponding to the linear fit of ρ(T ). Experimental data are given for ULTEM 1000 [6]. The 1-st example shows (see Fig. 2) the results of the bulk modulus (B ) simulations in comparison with experimental data.The MD re- sults were obtained using formula B = V σ 2 V /k B T , where V is the volume, σ 2 V is its variance, and k B is the Boltzmann constant. Ex- perimental data for T<T g were estimated using measured Young’s modulus E and the equation B = E/3(1 2ν ). In the 2-nd example, the density ρ from the MD simulations is compared to the experimental data as a function of temperature (see the inset of Fig. 2). The linear fit of the ρ(T ) below and above the glass transition allows one to estimate T g . Diffusion of chains To simulate welding dynamics the samples were allowed to equili- brate in the NPT ensemble that keeps pressure, temperature, and the total number of particles fixed (Note that the temperature at the interface of two filaments is nearly fixed during 300 ns used in MD simulations). At time instances 60 and 240 ns the sample was quenched to the room temperature. Additional thermal cycling be- tween 300 and 600 K was used to analyze temperature dependence Figure 3:Atom density profiles of the polymers on both sides of the interface: (i) initial (black solid lines); (ii) at 0.5 ns (blue dashed lines); (iii) at 40 ns (red dashed lines). The inset shows example of a single chain reptation at the interface. The good agreement of the developed model with experimental data paves the way to semi- quantitative predictions of the interface properties of the polymer blends considered. of thermodynamic variables. The samples obtained after quenching and thermal cycling with different thickness of the welded layer were used for simulations of the strain-stress curves and shear viscosity. The profiles of the atomic densities on the two sides of the interface are shown in Fig. 3. It can be seen from the figure that the two samples are initially well separated, with the half-width of the gap at half bulk density 5Å. After 0.5 ns the density at the interface has nearly reached its bulk value. After another 40 ns the density profiles stay almost the same with only the tails of the distributions extending to the other side by nearly 20 Å. Shear viscosity Figure 4:Shear viscosity as a function of shear rate for large system after 60 ns (blue square), 240 ns (brown circles) of healing and after thermal cycling (gray asterisk). The shear deformation was applied in XY plane parallel to the interface plane of the sample. The results of simulation of shear deformation performed at welding times 60 ns and 240 ns are shown in Fig. 4. The dependence of η on the shear rate demonstrates characteristic shear thinning behavior, which can be clearly seen in log-log scale. Next, we observe that the shear viscosity calculated at 240 ns is shifted towards larger values as expected on theoretical grounds. However, this shift is consistent but is only weakly resolved (i.e. error bars) for high shear rates used in MD simulations. Strain-stress curves The strain-stress curves obtained in the MD simulations for three different welding times are compared with the experimental curve in Fig. 5. The curves deviate from the experimental data due to the high elongation rate used in the MD simulations. However, it can be seen from the figure that both Young’s modulus and the Yield strength increase as the thickness of welded layer increasing. We note that the calculations were performed under assumption of nominal Poisson’s ratio ν = 0.36. This assumption becomes increasingly inaccurate as the breaking of the sample is initiated at the interface. For this reason the strain-stress curves obtained in MD are shown by the dashed lines for the values of strain larger than 0.1. Figure 5:MD simulations of the strain-stress curves as a function of welding time: (i) 60 ns (diamonds); (ii) 240 ns squares; and (iii) after additional welding dur- ing temperature cycling (1200 ns). The experimental data obtained by ULTEM 1000 [6] are shown by teal shaded circles. Conclusion We developed a fully atomistic molecular dynamics model of the PEI/PC amorphous polymer blends with flat interface. The sam- ple was welded for over 300 ns and quickly quenched to 300 K after 60 and 240 ns. The model was validated by comparison of the MD predictions with experimental data for the density, glass transition temperature, bulk modulus, and thermal expansion coefficient be- low T g (where experimental data were available). Two characteristic time scales of welding were observed: (i) fast diffusion (t . 20 ns) when chain’s ends diffuse by filling in vacancies on both sides of the interface; and (ii) slow diffusion (t . 300 ns) when chain’s ends slowly diffuse through the bulk material. The analysis of the strain-stress curves as a function of time reveals the effect of the thickness of the welded layer on the strength of the interface. It was shown that both Young’s modulus and the Yield strength increase as the thickness of welded layer increasing. Over- all, the good agreement of the developed model with experimental data paves the way to semi-quantitative predictions of the interface properties of the polymer blends considered. References [1] S. Prager and M. Tirrell, “The healing process at polymer-polymer interfaces”, The Journal of Chemical Physics, 75(10):5194–5198, nov 1981. [2] Q. Sun, G.M. Rizvi, C.T. Bellehumeur, and P. Gu, “Effect of processing conditions on the bonding quality of FDM polymer filaments”, Rapid Prototyping Journal, 14(2):72–80, 2008. [3] S. G. Falkovich,S. V. Lyulin, V. M. Nazarychev, S. V. Larin, A. A. Gurtovenko, N. V. Lukasheva, and A. V. Lyulin, “Influence of the electrostatic interactions on thermophysical properties of polyimides: Molecular-dynamics simulations”, Journal of Polymer Science Part B: Polymer Physics, 52(9):640–646, feb 2014. [4] Ting Ge, Mark O. Robbins, Dvora Perahia, and Gary S. Grest, “Healing of polymer interfaces: Interfacial dynamics, entanglements, and strength”, Phys. Rev. E, 90:012602, Jul 2014. [5] JSOL corp., User’s Manual, J-OCTA Overview, J-OCTA 4.0, 2018. [6] Heat capacity, https://www.protolabs.com/media/1014801/ultem-1000-im.pdf, 2018. Acknowledgements We thank Gabriel Jost for the help in running simulations on Amazon Web Ser- vices. We thank Tracie Prayer for providing experimental data and guidance. The research was supported by NASA STMD/GCD/LSM grant in Space Man- ufacturing and by a Leverhulme Trust Research Project Grant RPG-2017-134.

Upload: others

Post on 20-Sep-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Atomistic model of reptation at polymer interfaces · 2019. 9. 9. · properties of the polymer blends considered. References [1] S. Prager and M. Tirrell, “The healing process

Atomistic model of reptation at polymer interfacesD.G. Luchinsky1,2, H. Hafiychuk1, M. Barabash2, V. Hafiychuk1, K. Chaki3, H. Nitta3, T. Ozawa3, K. R. Wheeler4, P.V.E. McClintock2

1SGT Inc., ARC, Moffett Field, California, USA, 2Department of Physics, Lancaster University, UK,3Materials Science Section, Engineering Technology Division, JSOL Corp, JP, 4Ames Research Center, Moffett Field, California, USA

Abstract

We study a molecular dynamics model ofa polymer-polymer interface for a polyetherim-ide/polycarbonate (PEI/PC) blend, including itsthermodynamic properties, its chain reptation, andits corresponding welding characteristics. Thestrength of the sample is analyzed by measuringstrain-stress curves in simulations of uni-axial elon-gation. The work is motivated by potential applica-tions to 3D manufacturing in space.

Introduction

Understanding properties of polymer-polymer interfaces is a long-standing problem of fundamental and technological importance [1].Polymer chains behavior such as reptation and entanglement deter-mine the welding dynamics and strength in 3D printing [2]. Theatomistic structure of polymers substantially influences these prop-erties. However, earlier research was mostly focused either on bulkproperties [3] or on coarse-grained models of interfaces [4].

Figure 1:Two amorphous cells (gray and cyan) of PEI/PC polymer blend arecombined in one chemical sample across an atomically flat interface.

Here we develop fully atomistic model of the polymer-polymer inter-face in PEI/PC blends. We asses MD capabilities in resolving poly-mer dynamics at polymer-polymer interfaces with atomic resolutionand analyze thermal and mechanical properties of the blends in thepresence of planar interface. We investigate the chains diffusionat the interface and the corresponding relaxation of the interfaceentanglement and non-bonding energy to the bulk state.

Model

We have prepared amorphous cells of these blends using the soft-ware package J-OCTA [5]. Cells were relaxed using the following theprocedure [5]: (i) equilibration in the NVE at a given temperatureT0; (ii) compression in the NPT Andersen-Nose-Hoover thermostatat T0 and P = 100 MPa; (iii) additional equilibration in the NVEequilibration at T0; (iv) relaxation in the NPT ensemble at T0 andP0; and (v) elimination of the translational component of the ve-locity in the NVT Nose-Hoover thermostat at T0 and zero pressure.Each stage was computed during 100 ps with time step 1 fs. Thetemperature and the pressure during equilibration were T0 = 600K and P0 = 1 atm (except for the 2-nd step, where P = 100 MPa.)The atomistically flat surfaces shown in Fig. 1 were prepared bymoving polymer chains inside the cells and using LJ-wall. Theparameters of the LJ-wall at the interface were: (i) cut-off ∼10 Å;(ii) sigma ∼5 Å; (iii) epsilon ∼10 kJ/mol; and (iv) density ∼20g/cm3.

Validation

To validate the model we we compared MD predictions with the fol-lowing experimentally measured properties: (i) density ρ; (ii) glasstransition temperature Tg; (iii) bulk modulus B; (iv) coefficient ofthermal expansion αP ; (v) IR spectra and density of states S(v);and (vi) specific heat capacity cp.

Figure 2:Bulk modulus B estimated using fluctuations (red open diamonds) incomparison with experimental data (teal circles). In the inset: density ρ vs T

obtained in MD calculations as compared to the experimental data (teal circles).Arrows show the glass transition temperatures Tg obtained in experiment andsimulations at the intersections of the straight lines corresponding to the linearfit of ρ(T ). Experimental data are given for ULTEM 1000 [6].

The 1-st example shows (see Fig. 2) the results of the bulk modulus(B) simulations in comparison with experimental data.The MD re-sults were obtained using formula B = V ⟨σ2

V ⟩ /kBT , where V is thevolume, σ2

V is its variance, and kB is the Boltzmann constant. Ex-perimental data for T < Tg were estimated using measured Young’smodulus E and the equation B = E/3(1 − 2ν).

In the 2-nd example, the density ρ from the MD simulations iscompared to the experimental data as a function of temperature(see the inset of Fig. 2). The linear fit of the ρ(T ) below and abovethe glass transition allows one to estimate Tg.

Diffusion of chains

To simulate welding dynamics the samples were allowed to equili-brate in the NPT ensemble that keeps pressure, temperature, andthe total number of particles fixed (Note that the temperature atthe interface of two filaments is nearly fixed during 300 ns used inMD simulations). At time instances 60 and 240 ns the sample wasquenched to the room temperature. Additional thermal cycling be-tween 300 and 600 K was used to analyze temperature dependence

Figure 3:Atom density profiles of the polymers on both sides of the interface:(i) initial (black solid lines); (ii) at ∼0.5 ns (blue dashed lines); (iii) at ∼40 ns(red dashed lines). The inset shows example of a single chain reptation at theinterface.

The good agreement of the developed model with experimental data paves the way to semi-quantitative predictions of the interface properties of the polymer blends considered.

of thermodynamic variables. The samples obtained after quenchingand thermal cycling with different thickness of the welded layer wereused for simulations of the strain-stress curves and shear viscosity.The profiles of the atomic densities on the two sides of the interfaceare shown in Fig. 3. It can be seen from the figure that the twosamples are initially well separated, with the half-width of the gapat half bulk density ∼5Å. After ∼0.5 ns the density at the interfacehas nearly reached its bulk value. After another 40 ns the densityprofiles stay almost the same with only the tails of the distributionsextending to the other side by nearly 20 Å.

Shear viscosity

Figure 4:Shear viscosity as a function of shear rate for large system after 60 ns(blue square), 240 ns (brown circles) of healing and after thermal cycling (grayasterisk).

The shear deformation was applied in XY plane parallel to theinterface plane of the sample. The results of simulation of sheardeformation performed at welding times 60 ns and 240 ns are shownin Fig. 4. The dependence of η on the shear rate demonstratescharacteristic shear thinning behavior, which can be clearly seen inlog-log scale. Next, we observe that the shear viscosity calculatedat 240 ns is shifted towards larger values as expected on theoreticalgrounds. However, this shift is consistent but is only weakly resolved(i.e. ∝ error bars) for high shear rates used in MD simulations.

Strain-stress curves

The strain-stress curves obtained in the MD simulations for threedifferent welding times are compared with the experimental curvein Fig. 5. The curves deviate from the experimental data due tothe high elongation rate used in the MD simulations. However, itcan be seen from the figure that both Young’s modulus and theYield strength increase as the thickness of welded layer increasing.We note that the calculations were performed under assumptionof nominal Poisson’s ratio ν = 0.36. This assumption becomesincreasingly inaccurate as the breaking of the sample is initiated atthe interface. For this reason the strain-stress curves obtained inMD are shown by the dashed lines for the values of strain largerthan 0.1.

Figure 5:MD simulations of the strain-stress curves as a function of welding time:(i) 60 ns (diamonds); (ii) 240 ns squares; and (iii) after additional welding dur-ing temperature cycling (1200 ns). The experimental data obtained by ULTEM1000 [6] are shown by teal shaded circles.

Conclusion

We developed a fully atomistic molecular dynamics model of thePEI/PC amorphous polymer blends with flat interface. The sam-ple was welded for over 300 ns and quickly quenched to 300 K after60 and 240 ns. The model was validated by comparison of the MDpredictions with experimental data for the density, glass transitiontemperature, bulk modulus, and thermal expansion coefficient be-low Tg (where experimental data were available).Two characteristic time scales of welding were observed: (i) fastdiffusion (t . 20 ns) when chain’s ends diffuse by filling in vacancieson both sides of the interface; and (ii) slow diffusion (t . 300 ns)when chain’s ends slowly diffuse through the bulk material.The analysis of the strain-stress curves as a function of time revealsthe effect of the thickness of the welded layer on the strength of theinterface. It was shown that both Young’s modulus and the Yieldstrength increase as the thickness of welded layer increasing. Over-all, the good agreement of the developed model with experimentaldata paves the way to semi-quantitative predictions of the interfaceproperties of the polymer blends considered.

References

[1] S. Prager and M. Tirrell, “The healing process at polymer-polymer interfaces”, The Journal of ChemicalPhysics, 75(10):5194–5198, nov 1981.

[2] Q. Sun, G.M. Rizvi, C.T. Bellehumeur, and P. Gu, “Effect of processing conditions on the bonding quality ofFDM polymer filaments”, Rapid Prototyping Journal, 14(2):72–80, 2008.

[3] S. G. Falkovich, S. V. Lyulin, V. M. Nazarychev, S. V. Larin, A. A. Gurtovenko, N. V. Lukasheva, and A. V.Lyulin, “Influence of the electrostatic interactions on thermophysical properties of polyimides:Molecular-dynamics simulations”, Journal of Polymer Science Part B: Polymer Physics, 52(9):640–646, feb2014.

[4] Ting Ge, Mark O. Robbins, Dvora Perahia, and Gary S. Grest, “Healing of polymer interfaces: Interfacialdynamics, entanglements, and strength”, Phys. Rev. E, 90:012602, Jul 2014.

[5] JSOL corp., User’s Manual, J-OCTA Overview, J-OCTA 4.0, 2018.[6] Heat capacity, https://www.protolabs.com/media/1014801/ultem-1000-im.pdf, 2018.

Acknowledgements

We thank Gabriel Jost for the help in running simulations on Amazon Web Ser-vices. We thank Tracie Prayer for providing experimental data and guidance.The research was supported by NASA STMD/GCD/LSM grant in Space Man-ufacturing and by a Leverhulme Trust Research Project Grant RPG-2017-134.