@article{zhai_koh_vogt_simon_2024, title={Kinetics of nanoconfined benzyl methacrylate radical polymerization}, volume={62}, ISSN={2642-4150 2642-4169}, url={http://dx.doi.org/10.1002/pol.20230332}, DOI={10.1002/pol.20230332}, abstractNote={AbstractThe effect of nanoconfinement on the kinetics of benzyl methacrylate radical polymerization is investigated using differential scanning calorimetry. Controlled pore glass (CPG), ordered mesoporous carbons, and mesoporous silica are used as confinement media with pore sizes from 2 to 8 nm. The initial polymerization rate in CPG and mesoporous silica increases relative to the bulk and increases linearly with reciprocal pore size; whereas, the rate in the carbon mesopores decreases linearly with reciprocal pore size; the changes are consistent with the rate being related to the ratio of the pore surface area to pore volume. Induction times are longer for nanoconfined polymerizations, and in the case of CPG and carbon mesopores, autoacceleration occurs earlier, presumably due to the limited diffusivity and lower termination rates for the confined polymer chains. The molecular weight of the polymer synthesized in the nanopores is generally higher than that obtained in the bulk except at the lowest temperatures investigated. The equilibrium conversion under nanoconfinement decreases with decreasing temperature and with confinement size, exhibiting what appears to be a floor temperature at low temperatures.}, number={9}, journal={Journal of Polymer Science}, publisher={Wiley}, author={Zhai, Chunhao and Koh, Yung P. and Vogt, Bryan D. and Simon, Sindee L.}, year={2024}, month={Feb}, pages={1922–1933} } @article{denton_koh_simon_mckenna_2024, title={On the glass transition temperature of TNT}, volume={736}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/j.tca.2024.179733}, DOI={10.1016/j.tca.2024.179733}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Denton, Aric A. and Koh, Yung P. and Simon, Sindee L. and McKenna, Gregory B.}, year={2024}, month={Jun}, pages={179733} } @article{pallaka_simon_2024, title={The glass transition and enthalpy recovery of polystyrene nanorods using Flash differential scanning calorimetry}, volume={160}, ISSN={0021-9606 1089-7690}, url={http://dx.doi.org/10.1063/5.0190076}, DOI={10.1063/5.0190076}, abstractNote={The glass transition (Tg) behavior and enthalpy recovery of polystyrene nanorods within an anodic aluminum oxide (AAO) template (supported nanorods) and after removal from AAO (unsupported nanorods) is studied using Flash differential scanning calorimetry. Tg is found to be depressed relative to the bulk by 20 ± 2 K for 20 nm-diameter unsupported polystyrene (PS) nanorods at the slowest cooling rate and by 9 ± 1 K for 55 nm-diameter rods. On the other hand, bulk-like behavior is observed in the case of unsupported 350 nm-diameter nanorods and for all supported rods in AAO. The size-dependent Tg behavior of the PS unsupported nanorods compares well with results for ultrathin films when scaled using the volume/surface ratio. Enthalpy recovery was also studied for the 20 and 350 nm unsupported nanorods with evolution toward equilibrium found to be linear with logarithmic time. The rate of enthalpy recovery for the 350 nm rods was similar to that for the bulk, whereas the rate of recovery was enhanced for the 20 nm rods for down-jump sizes larger than 17 K. A relaxation map summarizes the behavior of the nanorods relative to the bulk and relative to that for the 20 nm-thick ultrathin film. Interestingly, the fragility of the 20 nm-diameter nanorod and the 20 nm ultrathin film are identical within the error of measurements, and when plotted vs departure from Tg (i.e., T - Tg), the relaxation maps of the two samples are identical in spite of the fact that the Tg is depressed 8 K more in the nanorod sample.}, number={12}, journal={The Journal of Chemical Physics}, publisher={AIP Publishing}, author={Pallaka, Madhusudhan R. and Simon, Sindee L.}, year={2024}, month={Mar} } @article{koh_fondren_denton_simon_mckenna_2022, title={Amorphization and Crystallization of Hexanitroazobenzene (HNAB) Using Conventional DSC and Flash DSC}, volume={47}, ISSN={0721-3115 1521-4087}, url={http://dx.doi.org/10.1002/prep.202100366}, DOI={10.1002/prep.202100366}, abstractNote={AbstractThe present work presents results from an investigation of the glass transition and crystallization behaviors of HNAB tested over more than five orders of magnitude of cooling rate from 0.005 °C/s to 600 °C/s (0.3 to 36000 °C/min) by a combination of conventional and Flash differential scanning calorimetry (DSC). The work quantifies the influence of the thermal amorphization route on the properties of this high explosive. Cooling rates faster than 100 °C/s (6000 °C/min) result in amorphous HNAB as expected from prior work, but we also find that amorphization of the HNAB occurs at cooling rates slower than 0.008 °C/s (0.5 °C/min). The behavior of the amorphous HNAB made by slow cooling is compared with that of amorphous HNAB made by fast cooling, as well as with that made by solvent casting in terms of glass transition temperature, apparent activation energy of glass transition, and dynamic fragility parameter m. Besides, the non‐isothermal crystallization response as a function of cooling rate is also reported. The thermal stability and decomposition energy of amorphous HNAB are compared with those of the crystalline counterpart, being similar heats of decomposition of 3295 and 3392 J/g, respectively; suggesting that the amorphous HNAB will have similar thermal stability and chemical energy to the crystalline form.}, number={10}, journal={Propellants, Explosives, Pyrotechnics}, publisher={Wiley}, author={Koh, Yung P. and Fondren, Zachary T. and Denton, Aric A. and Simon, Sindee L. and McKenna, Gregory B.}, year={2022}, month={Aug} } @article{lopez_koh_zapata-hincapie_simon_2022, title={Composition-dependent glass transition temperature in mixtures: Evaluation of configurational entropy models}, volume={5}, ISSN={["1548-2634"]}, url={https://doi.org/10.1002/pen.26018}, DOI={10.1002/pen.26018}, abstractNote={AbstractThe composition dependence of the glass transition temperature (Tg) in mixtures remains an important unsolved problem. Here, it is revisited using three model systems: a series of oligomeric and polymeric cyanurates, blends of oligomeric and polymeric α‐methyl styrene, and molecular mixtures of itraconazole and posaconazole. We evaluate several entropy‐based models to determine the theoreticalTgas a function of molecular composition and compare the results against the experimental data. The assumption that the configurational entropy is invariant at theTgis tested, where the change in configurational entropy is assumed to be given by the integral of ΔCpdlnT, where ΔCpis the temperature‐dependent change in the heat capacity atTg. We find that, although the temperature‐dependent heat capacities in both liquid and glassy states are nearly independent of composition for several of the systems studied (i.e., they are nearly ideal mixtures), the composition dependence ofTgis not well described by simply adding the changes in the mass‐weighted configurational entropy of the components on going from theTgin the pure state to that of the blend. The implication is that either configurational entropy is not invariant atTgor that it cannot be obtained from the integral of ΔCpdlnT.}, journal={POLYMER ENGINEERING AND SCIENCE}, author={Lopez, Evelyn and Koh, Yung P. and Zapata-Hincapie, John A. and Simon, Sindee L.}, year={2022}, month={May} } @article{tian_koh_orski_simon_2022, title={Dodecyl Methacrylate Polymerization under Nanoconfinement: Reactivity and Resulting Properties}, volume={55}, ISSN={0024-9297 1520-5835}, url={http://dx.doi.org/10.1021/acs.macromol.1c01724}, DOI={10.1021/acs.macromol.1c01724}, abstractNote={The effect of nanoconfinement on the free radical polymerization of dodecyl methacrylate (DMA) with di-tert-butyl peroxide (DtBP) initiator is investigated over a wide temperature range from 110 to 190 °C using differential scanning calorimetry. The reaction shows a distinct induction time, which decreases as temperature increases, with an activation energy that is the same, albeit, of opposite sign, as that for dissociation of the initiator. The rate of reaction increases with increasing temperature and is higher in nanopores than in bulk conditions, with an Arrhenius temperature dependence at temperatures lower than 160 °C and an activation energy that is approximately 10% lower in the nanoconfined cases than for bulk. The higher reaction rate and lower activation energies in the nanopores are presumably due to specific interactions between the monomer and the native silanol groups on the pore surface. The enhancement of the reaction rate is found to be inversely related to the length of the alkyl group and the water contact angle comparerd data for several poly(n-alkyl methacrylate) studied previously. For bulk and nanoconfined DMA polymerizations, the molar mass increases as temperature decreases with a cross-linked product obtained at temperatures below 170 °C. The gel fraction increases as temperature decreases and is nearly 80% at 110 °C. In the nanopores, the molar mass is smaller compared to that in bulk conditions at high temperatures. The results can be described by a simplified recursive model.}, number={19}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Tian, Qian and Koh, Yung P. and Orski, Sara V. and Simon, Sindee L.}, year={2022}, month={Sep}, pages={8723–8730} } @article{pallaka_bari_simon_2022, title={Origin of the broad endothermic peak observed at low temperatures for polystyrene and metals in Flash differential scanning calorimetry}, volume={62}, ISSN={["1548-2634"]}, url={https://doi.org/10.1002/pen.26102}, DOI={10.1002/pen.26102}, abstractNote={AbstractThe glass transition behavior of polystyrene as a function of different cooling rates with scanning to two different end temperatures, 30°C and −80°C, was investigated for four different substrate conditions using Flash differential scanning calorimetry, a fast scanning nanocalorimetry technique. In addition, structural recovery of polystyrene was performed at 20°C for aging times from 0.01 s to 8 h with scanning to −80°C for the same samples. A broad endotherm appears to grow at low temperatures (T << Tg) as cooling rate decreases and aging time increases, which is influenced by the substrate underlying the film, as well as by the end temperature condition in the scanning experiment. On the other hand, the endothermic overshoot associated with Tg is not influenced by substrate or scan end temperature. In addition, indium and vapor‐deposited gold, both crystalline materials, show the growth of a very similar broad endotherm at low temperatures as cooling rate decreases and aging time increases indicating that the low‐temperature endotherm is an artifact and not a relaxation associated with the material under investigation. Several potential explanations are put forward.}, number={9}, journal={POLYMER ENGINEERING AND SCIENCE}, publisher={Wiley}, author={Pallaka, Madhusudhan R. and Bari, Rozana and Simon, Sindee L.}, year={2022}, month={Sep}, pages={3059–3069} } @article{sakib_koh_simon_2022, title={The absolute heat capacity of polymer grafted nanoparticles using fast scanning calorimetry}, volume={7}, ISSN={["1548-2634"]}, url={https://doi.org/10.1002/pen.26078}, DOI={10.1002/pen.26078}, abstractNote={AbstractThe absolute heat capacity of a matrix‐free polystyrene grafted silica nanocomposite system is investigated using ultrafast Flash differential scanning calorimetry (Flash DSC) and compared to results for a neat polystyrene sample of similar molecular weight. The heat capacity of the polymer in the nanocomposite is 8% lower than that of the neat polystyrene. In addition, the step change in heat capacity at the glass transition Tg is 15% lower, indicating a 2 nm‐thick immobile layer of polymer around the nanoparticles. There is neither an increase in the nanocomposite heat capacity towards that of the neat material nor any additional ∆Cp steps, indicating that the glassy layer around the nanoparticles remains immobile up to 300°C. In addition to the analysis of the absolute heat capacity, the dynamic temperature gradient in the Flash DSC sample increases linearly with increasing rate of measurement, whereas the noise in the heat capacity data increases linearly with the reciprocal of the scanning rate; the optimum rate for measurements lies between 300 and 1000 K/s for the samples studied here and will depend on the sample thickness.}, journal={POLYMER ENGINEERING AND SCIENCE}, publisher={Wiley}, author={Sakib, Nazam and Koh, Yung P. and Simon, Sindee L.}, year={2022}, month={Jul} } @article{zhao_grassia_simon_2021, title={Mobility of Pressure-Densified and Pressure-Expanded Polystyrene Glasses: Dilatometry and a Test of KAHR Model}, volume={54}, ISSN={["1520-5835"]}, url={https://doi.org/10.1021/acs.macromol.1c00983}, DOI={10.1021/acs.macromol.1c00983}, abstractNote={The structural recovery of pressure-densified (PDG) and, for the first time, pressure-expanded (PEG) glasses are experimentally investigated using pressurizable dilatometry. Both glasses show early devitrification on heating, indicating that these glasses have more mobility, compared to the conventional isobarically formed glass. The Kovacs–Aklonis–Hutchinson–Ramos (KAHR) model of structural recovery is able to reasonably predict the behavior of the pressure-expanded glass, but the KAHR model fails with the pressure-densified glass. The results suggest two limitations of the model: (i) the structural recovery is assumed to depend on the instantaneous liquid state and (ii) the same relaxation kinetics are assumed for the temperature and pressure perturbations. Modification of the KAHR model, allowing the departure from equilibrium, δ, to initially depend on the liquid state that the glass came from and to evolve toward the state that the glass is going to, improves the ability of the model to predict the early devitrification for the pressure-densified glass. Another modification of the KAHR model, allowing the temperature and pressure perturbations to relax independently of one another, results in effectively capturing the increased thermal expansion coefficient of glass lines during heating, as well as a "memory"-like aging behavior, for the pressure-densified glass.}, number={18}, journal={MACROMOLECULES}, publisher={American Chemical Society (ACS)}, author={Zhao, Xiao and Grassia, Luigi and Simon, Sindee L.}, year={2021}, month={Sep}, pages={8352–8364} } @article{zhao_cheng_koh_kelly_mckenna_simon_2021, title={Prediction of the Synergistic Glass Transition Temperature of Coamorphous Molecular Glasses Using Activity Coefficient Models}, volume={18}, ISSN={["1543-8392"]}, url={https://doi.org/10.1021/acs.molpharmaceut.1c00353}, DOI={10.1021/acs.molpharmaceut.1c00353}, abstractNote={The glass transition temperature (Tg) of a binary miscible mixture of molecular glasses, termed a coamorphous glass, is often synergistically increased over that expected for an athermal mixture due to the strong interactions between the two components. This synergistic interaction is particularly important for the formulation of coamorphous pharmaceuticals since the molecular interactions and resulting Tg strongly impact stability against crystallization, dissolution kinetics, and bioavailability. Current models that describe the composition dependence of Tg for binary systems, including the Gordon-Taylor, Fox, Kwei, and Braun-Kovacs equations, fail to describe the behavior of coamorphous pharmaceuticals using parameters consistent with experimental ΔCP and Δα. Here, we develop a robust thermodynamic approach extending the Couchman and Karasz method through the use of activity coefficient models, including the two-parameter Margules, non-random-two-liquid (NRTL), and three-suffix Redlich-Kister models. We find that the models, using experimental values of ΔCP and fitting parameters related to the binary interactions, successfully describe observed synergistic elevations and inflections in the Tg versus composition response of coamorphous pharmaceuticals. Moreover, the predictions from the NRTL model are improved when the association-NRTL version of that model is used. Results are reported and discussed for four different coamorphous systems: indomethacin-glibenclamide, indomethacin-arginine, acetaminophen-indomethacin, and fenretinide-cholic acid.}, number={9}, journal={MOLECULAR PHARMACEUTICS}, publisher={American Chemical Society (ACS)}, author={Zhao, Xiao and Cheng, Sixue and Koh, Yung P. and Kelly, Brandon D. and McKenna, Gregory B. and Simon, Sindee L.}, year={2021}, month={Sep}, pages={3439–3451} } @article{zhao_simon_2020, title={A model-free analysis of configurational properties to reduce the temperature- and pressure-dependent segmental relaxation times of polymers}, url={https://doi.org/10.1063/1.5131623}, DOI={10.1063/1.5131623}, abstractNote={The segmental relaxation time data for poly(vinyl acetate), poly(vinyl chloride), and linear and star polystyrene are analyzed using a model-free method to determine how the temperature- and pressure-dependent relaxation times, τ, scale with the relative configurational thermodynamic properties. The model-free method assumes no specific mathematical form, such as reciprocal linearity, and the configurational properties are referred to an isochronal state to eliminate the bias associated with the definition of the ideal glassy state. The scaling ability of a given configurational property is strongly material-dependent with the logarithm of τ scaling better with TSc and Hc for poly(vinyl acetate), with TSc, Hc, and Uc for poly(vinyl chloride), and with TSc, Hc, and Vc for linear and star polystyrene. The choice of the isochronal reference state does not qualitatively affect the results.}, journal={The Journal of Chemical Physics}, author={Zhao, Xiao and Simon, Sindee L.}, year={2020}, month={Jan} } @article{bari_koh_mckenna_simon_2020, title={Decomposition of HMX in solid and liquid states under nanoconfinement}, volume={686}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/j.tca.2020.178542}, DOI={10.1016/j.tca.2020.178542}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Bari, Rozana and Koh, Yung P. and McKenna, Gregory B. and Simon, Sindee L.}, year={2020}, month={Apr}, pages={178542} } @article{tian_zhao_simon_2020, title={Kinetic study of alkyl methacrylate polymerization in nanoporous confinement over a broad temperature range}, volume={205}, ISSN={0032-3861}, url={http://dx.doi.org/10.1016/j.polymer.2020.122868}, DOI={10.1016/j.polymer.2020.122868}, journal={Polymer}, publisher={Elsevier BV}, author={Tian, Qian and Zhao, Haoyu and Simon, Sindee L.}, year={2020}, month={Sep}, pages={122868} } @article{zhao_simon_2020, title={Synthesis of polymers in nanoreactors: A tool for manipulating polymer properties}, volume={211}, ISSN={0032-3861}, url={http://dx.doi.org/10.1016/j.polymer.2020.123112}, DOI={10.1016/j.polymer.2020.123112}, journal={Polymer}, publisher={Elsevier BV}, author={Zhao, Haoyu and Simon, Sindee L.}, year={2020}, month={Dec}, pages={123112} } @article{sakib_koh_huang_mongcopa_le_benicewicz_krishnamoorti_simon_2020, title={Thermal and Rheological Analysis of Polystyrene-Grafted Silica Nanocomposites}, volume={53}, url={https://doi.org/10.1021/acs.macromol.9b02127}, DOI={10.1021/acs.macromol.9b02127}, abstractNote={Two matrix-free polystyrene-grafted silica nanocomposite samples with graft chain lengths of 35 and 112 kg/mol are characterized by calorimetry and rheometry, and results are compared to neat polystyrenes of comparable molecular weights. The glass transition temperature Tg of the nanocomposites is found to be approximately 1 to 2 K higher than that of the neat materials, whereas the absolute heat capacity is approximately 4–7% lower in the glassy and liquid states. The step change in heat capacity ΔCp at Tg is 15% lower for the nanocomposites, consistent with an immobilized glassy layer of approximately 2 nm. The linear viscoelastic behavior of the nanocomposite samples differs significantly compared to their neat analogs in several ways: first, the G′ versus ω curves shift toward lower frequencies by approximately one decade due to the increase in the glass transition temperature; second, terminal flow behavior is absent; third, the rubbery plateau moduli (GN°) decreases by 7% for the 35 kg/mol grafted particles and increases by approximately two and a half-fold for the 112 kg/mol grafted particles; and fourth, the glassy modulus increases approximately 4% consistent with hydrodynamic reinforcement. On the other hand, the magnitude of the rubbery modulus is attributed to two effects, hydrodynamic reinforcement and a change in the effective entanglement density, which is governed by corona interpenetration coupled with the silica particles acting as physical entanglement points.}, number={6}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Sakib, Nazam and Koh, Yung P. and Huang, Yucheng and Mongcopa, Katrina Irene S. and Le, Amy N. and Benicewicz, Brian C. and Krishnamoorti, Ramanan and Simon, Sindee L.}, year={2020}, month={Mar}, pages={2123–2135} } @article{bari_denton_fondren_mckenna_simon_2019, title={Acceleration of decomposition of CL-20 explosive under nanoconfinement}, volume={140}, ISSN={1388-6150 1588-2926}, url={http://dx.doi.org/10.1007/s10973-019-09027-5}, DOI={10.1007/s10973-019-09027-5}, number={6}, journal={Journal of Thermal Analysis and Calorimetry}, publisher={Springer Science and Business Media LLC}, author={Bari, Rozana and Denton, Aric A. and Fondren, Zachary T. and McKenna, Gregory B. and Simon, Sindee L.}, year={2019}, month={Nov}, pages={2649–2655} } @article{lee_yeo_hu_thalangama-arachchige_kaur_quitevis_kumar_koh_simon_2019, title={Friction and Wear of Pd-Rich Amorphous Alloy (Pd43Cu27Ni10P20) with Ionic Liquid (IL) as Lubricant at High Temperatures}, url={https://doi.org/10.3390/met9111180}, DOI={10.3390/met9111180}, abstractNote={The friction and wear behavior of palladium (Pd)-rich amorphous alloy (Pd43Cu27Ni10P20) against 440C stainless steel under ionic liquids as lubricants, i.e., 1-nonyl-3-methylimidazolium bis[(trifluoromethane)sulfonyl]amide ([C9C1im][NTf2]), were investigated using a ball-on-disc reciprocating tribometer at ambient, 100 and 200 °C with different sliding speeds of 3 and 7 mm/s, whose results were compared to those from crystalline Pd samples. The measured coefficient of friction (COF) and wear were affected by both temperature and sliding speed. The COF of crystalline Pd samples dramatically increased when the temperature increased, whereas the COF of the amorphous Pd alloy samples remained low. As the sliding speed increased, the COF of both Pd samples showed decreasing trends. From the analysis of a 3D surface profilometer and scanning electron microscopy (SEM) with electron dispersive spectroscopy (EDS) data, three types of wear (i.e., delamination, adhesive, and abrasive wear) were observed on the crystalline Pd surfaces, whereas the amorphous Pd alloy surfaces produced abrasive wear only. In addition, X-ray photoelectron spectroscopy (XPS) measurements were performed to study the formation of tribofilm. It was found that the chemical reactivity at the contacting interface increased with temperature and sliding contact speed. The ionic liquids (ILs) were effective as lubricants when the applied temperature and sliding speed were 200 °C and 7 mm/s, respectively.}, journal={Metals}, author={Lee, Jaeho and Yeo, Chang-Dong and Hu, Zhonglue and Thalangama-Arachchige, Vidura D. and Kaur, Jagdeep and Quitevis, Edward L. and Kumar, Golden and Koh, Yung P. and Simon, Sindee}, year={2019}, month={Nov} } @article{zapata h_simon_grady_2019, title={Influence of diameter on the degradation profile of multiwall carbon nanotubes}, volume={138}, ISSN={1388-6150 1588-2926}, url={http://dx.doi.org/10.1007/s10973-019-08137-4}, DOI={10.1007/s10973-019-08137-4}, number={2}, journal={Journal of Thermal Analysis and Calorimetry}, publisher={Springer Science and Business Media LLC}, author={Zapata H, John A. and Simon, Sindee L. and Grady, Brian P.}, year={2019}, month={Mar}, pages={1351–1362} } @article{zhai_quitevis_simon_2019, title={Kinetic Study of Curing Bisphenol A Dicyanate Ester with Ionic Liquid Additive}, url={https://doi.org/10.1002/polb.24874}, DOI={10.1002/polb.24874}, abstractNote={ABSTRACT A kinetic study of the trimerization reaction of bisphenol A dicyanate ester with an aromatic imidazolium‐based ionic liquid (IL) as additive is performed using dynamic and isothermal differential scanning calorimetry. The reaction follows second‐order autocatalytic kinetics, and a slight acceleration effect is observed in the presence of the aromatic IL relative to the neat resin. The activation energy also increases with the IL additive, whereas the glass transition temperature ( T g ) is depressed, consistent with the Fox equation and a homogeneous one‐phase material. A model incorporating diffusion effects is able to describe the dynamic and isothermal curing data for both the neat resin system and that containing aromatic IL. A comparison with aliphatic‐based IL additive indicates that the reaction is more accelerated with aliphatic IL than with the aromatic IL in spite of the fact that the aliphatic additive phase separates during cure. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2019 , 57 , 1315–1324}, journal={Journal of Polymer Science Part B: Polymer Physics}, author={Zhai, Chunhao and Quitevis, Edward L. and Simon, Sindee L.}, year={2019}, month={Oct} } @article{qian_koh_pallaka_chang_lin_guzmán_grubbs_simon_mckenna_2019, title={Linear Rheology of a Series of Second-Generation Dendronized Wedge Polymers}, volume={52}, url={https://doi.org/10.1021/acs.macromol.8b02122}, DOI={10.1021/acs.macromol.8b02122}, abstractNote={A series of second-generation dendronized wedge polymers were synthesized by ring-opening metathesis polymerization, and the linear viscoelastic response over a wide range of temperatures was investigated. From 0 to 90 °C the dynamic moduli (G′(ω) and G″(ω)) were determined, and frequency–temperature superposition was used to create master curves that showed behavior from the terminal zone to the glassy regime. An apparent extremely low rubbery plateau of ∼10 kPa was observed in both the dynamic response and in the corresponding van Gurp–Palmen plot. However, further investigation shows that the apparent rubbery plateau is related to the steady-state recoverable compliance, not the onset of entanglements. In addition, these wedge polymers exhibit an extremely low glassy modulus of ∼100 MPa at 0 °C, which is shown to increase at 1 Hz to ∼700 MPa at −80 °C for the wedge polymer 2G-EHW-311. In addition, both small- and wide-angle X-ray scattering patterns were obtained for all of the polymers investigated, and these showed that the polymer molecules adopt an extended cylinder conformation. Furthermore, based on calorimetric measurements, the polymers were found to exhibit two glass transition temperatures, with a 100 K difference between the higher (Tg,hi = 26.8 ± 0.7 °C) and lower glass transition temperatures (Tg,lo = −76.1 ± 1.1 °C) for the 2G-EHW-311 material. Hence, an intermediate regime extends to well below the Tg,hi to Tg,lo, providing an explanation for the low glassy modulus of ∼100 MPa at 0 °C and its increase to ∼700 MPa when measured at Tg,hi – 100 °C and approaching the Tg,lo.}, number={5}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Qian, Zhiyuan and Koh, Yung P. and Pallaka, Madhusudhan R. and Chang, Alice B. and Lin, Tzu-Pin and Guzmán, Pablo E. and Grubbs, Robert H. and Simon, Sindee L. and McKenna, Gregory B.}, year={2019}, month={Mar}, pages={2063–2074} } @article{grassia_koh_rosa_simon_2018, title={Complete Set of Enthalpy Recovery Data Using Flash DSC: Experiment and Modeling}, volume={51}, url={https://doi.org/10.1021/acs.macromol.7b02277}, DOI={10.1021/acs.macromol.7b02277}, abstractNote={Enthalpy recovery of a single polystyrene thin film is quantified by both experiments using Flash DSC and modeling using a new modified TNM model. Experimental data include Kovacs' three signatures of structural recovery: intrinsic isotherms after temperature down jumps, the asymmetry of approach after temperature down and up jumps of the same size, and the memory effect after a two-step history. A new modified TNM model is proposed to quantitatively fit all three signatures of structural recovery with a single set of model parameters. Here, we elucidate the detailed derivation of the new modified model and demonstrate its applicability to the experimental Flash DSC results.}, number={4}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Grassia, Luigi and Koh, Yung P. and Rosa, Mattia and Simon, Sindee L.}, year={2018}, month={Feb}, pages={1549–1558} } @article{koh_simon_2018, title={Enthalpy recovery of ultrathin polystyrene film using Flash DSC}, volume={143}, url={https://doi.org/10.1016/j.polymer.2018.02.038}, DOI={10.1016/j.polymer.2018.02.038}, journal={Polymer}, publisher={Elsevier BV}, author={Koh, Yung P. and Simon, Sindee L.}, year={2018}, month={May}, pages={40–45} } @article{pallaka_unruh_simon_2018, title={Melting behavior of n -alkanes in anodic aluminum oxide (AAO) nanopores using Flash differential scanning calorimetry}, volume={663}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/j.tca.2018.01.016}, DOI={10.1016/j.tca.2018.01.016}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Pallaka, Madhusudhan R. and Unruh, Daniel K. and Simon, Sindee L.}, year={2018}, month={May}, pages={157–164} } @article{zhang_quirk_gerislioglu_wesdemiotis_bekele_tsige_koh_simon_foster_2018, title={Synthesis and Characterization of Well-Defined, Tadpole-Shaped Polystyrene with a Single Atom Junction Point}, volume={51}, url={https://doi.org/10.1021/acs.macromol.8b01384}, DOI={10.1021/acs.macromol.8b01384}, abstractNote={An efficient method for synthesis of well-defined, well-characterized, tadpole-shaped polystyrene with a single atom junction point that is optimal for the study of dynamics has been developed using anionic polymerization, silicon chloride linking chemistry, and metathesis ring closure. The difunctional macromolecular linking agent, ω-methyldichlorosilylpolystyrene, was formed by reacting sec-butyllithium-initiated poly(styryl)lithium with excess (30×) methyltrichlorosilane to eliminate formation of linear dimer and three-arm star polystyrene. The asymmetric, three-arm, star precursor was formed by linking excess α-4-pentenylpoly(styryl)lithium (α-PSLi) with the macromolecular linking agent, and the excess α-PSLi functionalized with ethylene oxide before termination with methanol to facilitate chromatographic separation. Cyclization of the three-arm, star precursor to form tadpole-shaped polystyrene was effected in methylene chloride at high dilution using the Grubbs first generation catalyst, bis(tricyclohexylphosphine)benzylidene ruthenium(IV) chloride. The tadpole product was uniquely characterized by MALDI-MS using peaks that appeared characteristically 28 m/z units lower than those of the corresponding asymmetric, three-arm, star precursor, which corresponds to the loss of an ethylene unit. MD simulations find a smaller hydrodynamic volume for the tadpole-shaped PS as compared to the three-arm star precursor, in quantitative agreement with GPC results. Incorporating one cycle in the molecule, while leaving one chain end, leads to an increase in Tg of only 2.7 °C, much smaller than the increase of 13.6 °C seen when going from the linear chain to cyclic analog with no ends at all. The results are consistent with self-plasticization by free chain ends.}, number={23}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Zhang, Fan and Quirk, Roderic P. and Gerislioglu, Selim and Wesdemiotis, Chrys and Bekele, Selemon and Tsige, Mesfin and Koh, Yung P. and Simon, Sindee L. and Foster, Mark D.}, year={2018}, month={Dec}, pages={9509–9518} } @article{mckenna_simon_2017, title={50th Anniversary Perspective: Challenges in the Dynamics and Kinetics of Glass-Forming Polymers}, volume={50}, url={https://doi.org/10.1021/acs.macromol.7b01014}, DOI={10.1021/acs.macromol.7b01014}, abstractNote={The phenomenology of the glass transition and the associated behavior in the near liquid and glassy states are detailed, including the cooling rate dependence of the glass transition, Kovacs' three signatures of structural recovery, and enthalpy overshoots. Dynamics in the liquid regime just above Tg and the associated temperature dependences are also covered since this behavior is important to understanding the glassy dynamics. The current models of structural recovery and their shortcomings are presented. A number of important unanswered questions are discussed, including how the relaxation time in the glassy state depends on structure, the relationship between the evolution of different properties, the resolution of the Kauzmann paradox, and the behavior of the equilibrium relaxation time below Tg. New experimental approaches are needed to make breakthroughs, such as two that are described: one involving 20 Ma amber to test whether the Vogel temperature dependence continues for the equilibrium state below Tg and another involving an ideal polymer/pentamer mixture to obtain the entropy of the liquid far below TK in a test of the Kauzmann paradox. An unexplored regime of glassy behavior, characterized by ultrastability, high density, and low fictive temperature, is identified, and experiments to understand the material behavior in this region are motivated.}, number={17}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={McKenna, Gregory B. and Simon, Sindee L.}, year={2017}, month={Sep}, pages={6333–6361} } @article{yoon_koh_simon_mckenna_2017, title={An Ultrastable Polymeric Glass: Amorphous Fluoropolymer with Extreme Fictive Temperature Reduction by Vacuum Pyrolysis}, volume={50}, ISSN={0024-9297 1520-5835}, url={http://dx.doi.org/10.1021/acs.macromol.7b00623}, DOI={10.1021/acs.macromol.7b00623}, abstractNote={Vacuum pyrolysis deposition (VPD) has been used to create an ultrastable polymer glass having a fictive temperature Tf of as much as 57 K below the nominal glass transition temperature of the thermally rejuvenated polymer. Amorphous fluoropolymer films 300 to 700 nm thick were created by VPD followed by characterization of the thermal response using rapid-scanning chip calorimetry. The deposition was performed for substrates held at temperatures from 30.0 °C (303.2 K) to 116.7 °C (389.9 K) corresponding to approximately 0.75 to 0.97 times the limiting fictive temperature Tf′ ≈ Tg of the same material determined by cooling then heating at 600 K/s. Consistent with literature observations for small molecules that are vapor deposited in similar conditions relative to the material Tg, large enthalpy overshoots are observed, typical of both highly aged and ultrastable glasses. The 57 K reduction in Tf for the VPD polymers is greater than prior reports for physical vapor deposition of small molecules to form ultrastable glasses as well as greater than the Tf reductions seen in ambers from 20 million to over 200 million years of age. The potential of using such materials to investigate systems extremely deep into the glassy condition is discussed.}, number={11}, journal={Macromolecules}, publisher={American Chemical Society (ACS)}, author={Yoon, Heedong and Koh, Yung P. and Simon, Sindee L. and McKenna, Gregory B.}, year={2017}, month={May}, pages={4562–4574} } @article{bari_simon_2018, title={Determination of the nonlinearity and activation energy parameters in the TNM model of structural recovery}, volume={4}, url={https://doi.org/10.1007/s10973-017-6381-6}, DOI={10.1007/s10973-017-6381-6}, journal={Journal of Thermal Analysis and Calorimetry}, publisher={Springer Nature}, author={Bari, Rozana and Simon, Sindee L.}, year={2018}, month={Jan} } @article{tao_gurung_cetin_mayer_quitevis_simon_2017, title={Fragility of ionic liquids measured by Flash differential scanning calorimetry}, volume={654}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/j.tca.2017.05.008}, DOI={10.1016/j.tca.2017.05.008}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Tao, Ran and Gurung, Eshan and Cetin, M. Mustafa and Mayer, Michael F. and Quitevis, Edward L. and Simon, Sindee L.}, year={2017}, month={Aug}, pages={121–129} } @inbook{simon_mckenna_2017, title={Structural Recovery and Physical Aging of Polymeric Glasses}, booktitle={Polymer Glasses}, publisher={Taylor & Francis}, author={Simon, S.L. and McKenna, G.B.}, editor={Roth, C. B.Editor}, year={2017}, pages={23–54,} } @article{koh_simon_2017, title={The glass transition and enthalpy recovery of a single polystyrene ultrathin film using Flash DSC}, volume={146}, ISSN={0021-9606 1089-7690}, url={http://dx.doi.org/10.1063/1.4979126}, DOI={10.1063/1.4979126}, abstractNote={The kinetics of the glass transition are measured for a single polystyrene ultrathin film of 20 nm thickness using Flash differential scanning calorimetry (DSC). Tg is measured over a range of cooling rates from 0.1 to 1000 K/s and is depressed compared to the bulk. The depression decreases with increasing cooling rate, from 12 K lower than the bulk at 0.1 K/s to no significant change at 1000 K/s. Isothermal enthalpy recovery measurements are performed from 50 to 115 °C, and from these experiments, the temperature dependence of the induction time along the glass line is obtained, as well as the temperature dependence of the time scale required to reach equilibrium, providing a measure of the shortest effective glassy relaxation time and the longest effective equilibrium relaxation time, respectively. The induction time for the ultrathin film is found to be similar to the bulk at all temperatures presumably because the Tg values are the same due to the use of a cooling rate of 1000 K/s prior to the enthalpy recovery measurements. On the other hand, the times required to reach equilibrium for the ultrathin film and bulk are similar at 100 °C, and considerably shorter for the ultrathin film at 90 °C, consistent with faster dynamics under nanoconfinement at low temperatures. The magnitude of the "Tg depression" is smaller when using the equilibrium relaxation time from the structural recovery experiment as a measure of the dynamics than when measuring Tg after a cooling experiment. A relaxation map is developed to summarize the results.}, number={20}, journal={The Journal of Chemical Physics}, publisher={AIP Publishing}, author={Koh, Yung P. and Simon, Sindee L.}, year={2017}, month={Apr} } @article{xue_gurung_tamas_koh_shadeck_simon_maroncelli_quitevis_2016, title={Effect of Alkyl Chain Branching on Physicochemical Properties of Imidazolium-Based Ionic Liquids}, volume={61}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84960901764&partnerID=MN8TOARS}, DOI={10.1021/acs.jced.5b00658}, abstractNote={The branched ionic liquids (ILs) 1-(iso-alkyl)-3-methylimidazolium bis[(trifluoromethane)sulfonyl]amide ([(N – 2)mCN-1C1im][NTf2] with N = 3–7) were synthesized and their physicochemical properties characterized and compared with the properties of linear ILs 1-(n-alkyl)-3-methylimidazolium bis[(trifluoromethane)sulfonyl]amide ([CNC1im][NTf2] with N = 3–7). For N = 4–7, the density of the branched IL [(N – 2)mCN–1C1im][NTf2] is the same as that of its linear analogue [CNC1im][NTf2] within the standard uncertainty of the measurements. In the case of the N = 3 [1mC2C1im][NTf2]/[C3C1im][NTf2] pair, the density of the branched IL is 0.13% higher than that of the linear IL. For a branched/linear IL pair with a given N, the glass transition temperature Tg, melting temperature Tm, and viscosity η are higher for the branched IL than for the linear IL. [2mC3C1im][NTf2] is an exception in that its Tm is lower than that of [C4C1im][NTf2]. Moreover, the viscosity of [2mC3C1im][NTf2] is anomalously higher than what would be predicted based on the trend of the other branched ILs. These trends in the viscosities of the linear and branched ILs are consistent with recent molecular dynamics simulations. Thermal gravimetric analysis indicates that linear ILs are thermally more stable than branched ILs. Pulsed-gradient spin–echo (PGSE) NMR diffusion measurements show that the self-diffusion coefficients of the ions vary inversely with the viscosities according to the Stokes–Einstein (SE) equation. The hydrodynamic radii of the cations and anions of linear ILs calculated from the SE equation however are consistently higher than those of the corresponding branched ILs.}, number={3}, journal={Journal of Chemical and Engineering Data}, author={Xue, L. and Gurung, E. and Tamas, G. and Koh, Y.P. and Shadeck, M. and Simon, S.L. and Maroncelli, M. and Quitevis, E.L.}, year={2016}, pages={1078–1091} } @article{lopez_simon_2016, title={Signatures of Structural Recovery in Polystyrene by Nanocalorimetry}, volume={49}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000372856300043&KeyUID=WOS:000372856300043}, DOI={10.1021/acs.macromol.5b02112}, abstractNote={The intrinsic isotherm, asymmetry of approach, and memory effect experiments are performed for a high fictive temperature polystyrene glass in enthalpy space using nanocalorimetry. Using aging at times as short as 0.01 s and relatively high aging temperatures allows the complete evolution of all three of the signatures of structural recovery to be obtained for the first time in enthalpy space. The results from down jump experiments are compared to those obtained at lower temperatures using conventional differential scanning calorimetry (DSC) with respect to the time required to reach equilibrium, the apparent activation energy, and the Tool–Narayanaswamy–Moynihan (TNM) model parameters, and these results are independent of the measurement method being used; on the other hand, the physical aging rate is higher for the high fictive temperature glass.}, number={6}, journal={Macromolecules}, author={Lopez, Evelyn and Simon, Sindee L.}, year={2016}, pages={2365–2374} } @article{koh_gao_simon_2016, title={Structural recovery of a single polystyrene thin film using Flash DSC at low aging temperatures}, volume={96}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84968764389&partnerID=MN8TOARS}, DOI={10.1016/j.polymer.2016.04.047}, journal={Polymer (United Kingdom)}, publisher={Elsevier BV}, author={Koh, Yung P. and Gao, Siyang and Simon, Sindee L.}, year={2016}, pages={182–187} } @misc{simon_koh_2016, title={The Glass Transition and Structural Recovery Using Flash DSC}, ISBN={9783319313276 9783319313290}, url={http://dx.doi.org/10.1007/978-3-319-31329-0_14}, DOI={10.1007/978-3-319-31329-0_14}, journal={Fast Scanning Calorimetry}, publisher={Springer International Publishing}, author={Simon, Sindee L. and Koh, Yung P.}, year={2016}, pages={433–459} } @article{tao_simon_2015, title={Bulk and shear rheology of silica/polystyrene nanocomposite: Reinforcement and dynamics}, volume={53}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84925431169&partnerID=MN8TOARS}, DOI={10.1002/polb.23669}, abstractNote={Bulk and shear rheological studies were performed on a 10 wt % silica nanoparticle-filled polystyrene nanocomposite. The limiting moduli in glassy and rubbery states are higher for the nanocomposite than for the neat polymer; the increase is consistent with hydrodynamic reinforcement and is slightly higher than the lower bound of the rule of mixtures prediction. All evidence indicates that the presence of nanoparticles does not significantly change the polymer dynamics associated with glass transition, except to increase the Tg by 3 K. Comparison of the bulk and shear retardation spectra suggests that the underlying mechanisms for both responses are similar at short times and that the long-time chain modes available to the shear are not available to the bulk, consistent with Plazek's earlier findings. In addition, T − Tg and TVγ scaling, along with the findings of thermorheological complexity, are discussed. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2015, 53, 621–632}, number={9}, journal={Journal of Polymer Science, Part B: Polymer Physics}, author={Tao, R. and Simon, S.L.}, year={2015}, pages={621–632} } @book{simon_2015, title={Dynamics of confined glass-forming liquids near equilibrium conditions}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84955678973&partnerID=MN8TOARS}, DOI={10.1007/978-3-319-21948-6_11}, journal={Non-equilibrium Phenomena in Confined Soft Matter: Irreversible Adsorption, Physical Aging and Glass Transition at the Nanoscale}, author={Simon, S.L.}, year={2015}, pages={245–263} } @article{zhao_simon_2015, title={Equilibrium free-radical polymerization of methyl methacrylate under nanoconfinement}, volume={66}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000355884100023&KeyUID=WOS:000355884100023}, DOI={10.1016/j.polymer.2015.04.017}, journal={Polymer}, publisher={Elsevier BV}, author={Zhao, H.Y. and Simon, Sindee L.}, year={2015}, pages={173–178} } @article{gao_simon_2015, title={Measurement of the limiting fictive temperature over five decades of cooling and heating rates}, volume={603}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=CCC&KeyUT=CCC:000351646600016&KeyUID=CCC:000351646600016}, journal={Thermochimica Acta}, author={Gao, Siyang and Simon, Sindee L.}, year={2015}, pages={123–127} } @article{tao_simon_2015, title={Pressure-volume-temperature and glass transition behavior of silica/polystyrene nanocomposite}, volume={53}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000357287100005&KeyUID=WOS:000357287100005}, DOI={10.1002/polb.23749}, abstractNote={ABSTRACT The pressure‐volume‐temperature (PVT) behavior and glass transition behavior of a 10 wt % silica nanoparticle‐filled polystyrene (PS) nanocomposite sample are measured using a custom‐built pressurizable dilatometer. The PVT data are fitted to the Tait equation in both liquid and glassy states; the coefficient of thermal expansion α , bulk modulus K , and thermal pressure coefficient γ are examined as a function of pressure and compared to the values of neat PS. The glass transition temperature ( T g ) is reported as a function of pressure, and the limiting fictive temperature ( T f ′) from calorimetric measurements is reported as a function of cooling rate. Comparison with data for neat PS indicates that the nanocomposite has a slightly higher T g at elevated pressures, higher bulk moduli at all pressures studied, and its relaxation dynamics are more sensitive to volume. The results for the glassy γ values suggest that thermal residual stresses would not be reduced for the nanocomposite sample studied. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2015 , 53 , 1131–1138}, number={16}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Tao, Ran and Simon, Sindee L.}, year={2015}, pages={1131–1138} } @article{tao_simon_2015, title={Rheology of Imidazolium-Based Ionic Liquids with Aromatic Functionality}, volume={119}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000360947400052&KeyUID=WOS:000360947400052}, DOI={10.1021/acs.jpcb.5b06163}, abstractNote={A series of imidazolium-based ionic liquids with cyclic and aromatic groups and a bis[(trifluoromethane)sulfonyl]amide anion were characterized using dynamic and steady shear rheology in a wide temperature range. The cations investigated include 1-(cyclohexylmethyl)-3-methylimidazolium ([CyhmC1im](+)), 1-benzyl-3-methylimidazolium ([BnzC1im](+)), 1,3-dibenzylimidazolium ([(Bnz)2im](+)), and 1-(2-naphthylmethyl)-3-methylimidazolium ([NapmC1im](+)). Rheological properties are reported from terminal flow to the glassy state. All ionic liquids show very similar flow behavior with a glassy modulus approaching 1 GPa. The temperature dependences of the shift factors used for time-temperature superposition and the viscosity both follow the same VFT or WLF relationship, and the dynamic fragility at Tg ranges from 117 to 130 for the aromatic ionic liquids investigated, considerably more fragile, in the Angell sense, than the aliphatic analogs. Additionally, an anomalous aging effect in the dynamic viscosity response (i.e., a maximum observed at intermediate frequencies) was found using a strain-controlled rheometer but not using a stress-controlled rheometer.}, number={35}, journal={Journal of Physical Chemistry B}, author={Tao, Ran and Simon, Sindee L.}, year={2015}, pages={11953–11959} } @article{shamim_koh_simon_mckenna_2015, title={The glass transition of trinitrotoluene (TNT) by flash DSC}, volume={620}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84944754512&partnerID=MN8TOARS}, DOI={10.1016/j.tca.2015.10.003}, journal={Thermochimica Acta}, author={Shamim, N. and Koh, Y.P. and Simon, S.L. and McKenna, G.B.}, year={2015}, pages={36–39} } @article{lopez_simon_2015, title={Trimerization Reaction Kinetics and T g Depression of Polycyanurate under Nanoconfinement}, volume={48}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84937230940&partnerID=MN8TOARS}, DOI={10.1021/acs.macromol.5b00167}, abstractNote={Trimerization of a mixture containing a mono- and difunctional cyanate ester is investigated under the nanoporous confinement of silanized hydrophobic controlled pore glass using differential scanning calorimetry. The trimerization reaction of the nanoconfined monomer mixture is accelerated relative to the bulk by as much as 12 times in 8 nm pores, but this acceleration is less than half that observed for nanoconfinement of the individual monomers. The absolute reaction rate of the monomer mixture lies between those of the individual species, being slower than the monocyanate ester and faster than the dicyanate ester. The results are consistent with the hypothesis that the reaction acceleration is due to monomer ordering or layering at the pore surface, leading to a local concentration of reactive groups higher than in the bulk. In addition to the influence of nanoconfinement on trimerization kinetics, the molecular weight and glass transition temperature (Tg) of the polycyanurate formed in the nanopores are investigated. The molecular weight decreases approximately 20% for synthesis in the smallest 8 nm pores relative to the bulk value of 5200 g/mol. Upon extraction from the pores, the polymer Tg is 5–9 K higher than in the bulk. However, in the 8 nm diameter pores, a Tg depression of 44 K is observed relative to the value of the material after extraction from the pores. This depression lies between the values previously observed for the products of the individual cyanate esters which formed a low molecular weight trimer and a cross-linked polymer network. A secondary Tg, associated with a less mobile layer at the pore wall, is 26–40 K above the primary value. The implication is that the origin of confinement effects on reactivity and Tg differ, with changes in reactivity in this system arising from surface layering or ordering and Tg depressions arising from intrinsic size effects.}, number={13}, journal={Macromolecules}, author={Lopez, E. and Simon, S.L.}, year={2015}, pages={4692–4701} } @article{shamim_koh_simon_mckenna_2014, title={Glass transition temperature of thin polycarbonate films measured by flash differential scanning calorimetry}, volume={52}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84907881891&partnerID=MN8TOARS}, DOI={10.1002/polb.23583}, abstractNote={Flash differential scanning calorimetry was used to study the glass transition temperature Tg of polycarbonate ultrathin films. The investigation was made as a function of film thickness from 22 to 350 nm and over a range of cooling rates from 0.1 to 1000 K/s. Polycarbonate spin cast films were floated on a layer of grease on the calorimetric chip. The results show a greatly reduced glass temperature for the thinnest films relative to the macroscopic value. We also observed that the magnitude of the glass temperature reduction decreases as the cooling rate increases with the highest cooling rates showing little thickness dependence of the Tg. Dynamic fragility and activation energy at Tg were found to decrease with decreasing film thickness. The results are discussed in the context of literature reports for supported and freely standing polycarbonate films. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2014, 52, 1462–1468}, number={22}, journal={Journal of Polymer Science, Part B: Polymer Physics}, author={Shamim, N. and Koh, Y.P. and Simon, S.L. and McKenna, G.B.}, year={2014}, pages={1462–1468} } @article{gao_simon_2014, title={Measurement of the limiting fictive temperature over five decades of cooling and heating rates}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84906531686&partnerID=MN8TOARS}, DOI={10.1016/j.tca.2014.08.019}, journal={Thermochimica Acta}, author={Gao, S. and Simon, S.L.}, year={2014} } @article{koh_grassia_simon_2015, title={Structural recovery of a single polystyrene thin film using nanocalorimetry to extend the aging time and temperature range}, volume={603}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84923605555&partnerID=MN8TOARS}, DOI={10.1016/j.tca.2014.08.025}, journal={Thermochimica Acta}, author={Koh, Y.P. and Grassia, L. and Simon, S.L.}, year={2015}, pages={135–141} } @article{zhao_yu_begum_hedden_simon_2014, title={The effect of nanoconfinement on methyl methacrylate polymerization: T-g, molecular weight, and tacticity}, volume={55}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000341906100018&KeyUID=WOS:000341906100018}, DOI={10.1016/j.polymer.2014.07.040}, number={19}, journal={Polymer}, publisher={Elsevier BV}, author={Zhao, H.Y. and Yu, Z.N. and Begum, Fatema and Hedden, Ronald C. and Simon, Sindee L.}, year={2014}, pages={4959–4965} } @article{gao_simon_2014, title={The reaction kinetics of cyclopentadiene dimerization using differential scanning calorimetry: Experiments and modelling}, volume={589}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000340320400030&KeyUID=WOS:000340320400030}, DOI={10.1016/j.tca.2014.05.031}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Gao, Siyang and Simon, Sindee L.}, year={2014}, pages={241–246} } @article{tao_tamas_xue_simon_quitevis_2014, title={Thermophysical Properties of Imidazolium-Based Ionic Liquids: The Effect of Aliphatic versus Aromatic Functionality}, volume={59}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=CCC&KeyUT=CCC:000341619900004&KeyUID=CCC:000341619900004}, number={9}, journal={Journal of Chemical and Engineering Data}, author={Tao, Ran and Tamas, George and Xue, Lianjie and Simon, Sindee L. and Quitevis, Edward L.}, year={2014}, pages={2717–2724} } @article{gao_koh_simon_2013, title={Calorimetric Glass Transition of Single Polystyrene Ultrathin Films}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000314081800027&KeyUID=WOS:000314081800027}, DOI={10.1021/ma3020036}, abstractNote={The calorimetric glass transition (Tg) is measured for single polystyrene ultrathin films using a commercial rapid-scanning chip calorimeter as a function of cooling rate and film thickness. Films have been prepared in two ways: spin-cast films placed on a layer of inert oil or grease and films directly spin-cast on the back of the calorimetric chip. For the films on oil or on grease, the 160 nm thick films show results consistent with those of a bulk sample measured by conventional DSC. On the other hand, the 47 nm thick film on oil and 71 nm thick films both on oil and on grease show a Tg depression which decreases with increasing cooling rate; the magnitude of the Tg depression is similar to results reported in the literature for the most mobile substrate-supported films. For films directly spin-cast onto the sensor, a Tg depression is not observed for 47 and 71 nm thick films but is observed for a 16 nm thick film. These results are also within the range of the data on supported films in the literature but show a smaller depression than films on oil or grease. The effect of annealing is also investigated. For thick films and those directly spin-cast onto the sensor, annealing at 160 °C has no influence on heat flow curves; hence, Tg values remain unchanged. For the 47 and 71 nm thick films on either oil or grease, the depressed Tgs revert to the bulk values over the course of a day at 160 °C. Atomic force microscope (AFM) images show that annealing results in dewetting of the films with hole growth and thickening of the film to 200 nm, the latter of which is presumed to be the reason that Tgs revert to bulk values.}, number={2}, journal={Macromolecules}, author={Gao, S. Y. and Koh, Y. P. and Simon, S. L.}, year={2013}, pages={562–570} } @article{koh_simon_2013, title={Enthalpy Recovery of Polystyrene: Does a Long-Term Aging Plateau Exist?}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000322417100044&KeyUID=WOS:000322417100044}, DOI={10.1021/ma40112361}, number={14}, journal={Macromolecules}, author={Koh, Y. P. and Simon, S. L.}, year={2013}, pages={5815–5821} } @article{koh_simon_2013, title={Enthalpy recovery of polystyrene: Does a long-term aging plateau exist?}, volume={46}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84880651000&partnerID=MN8TOARS}, DOI={10.1021/ma4011236}, abstractNote={A glass is not in thermodynamic equilibrium below its glass transition temperature (Tg), and consequently, its properties, such as enthalpy, volume, and mechanical properties, evolve toward equilibrium in a process known as structural recovery or physical aging. Several recent studies have suggested that the extrapolated liquid line is not reached even when properties have ceased to evolve. In this work, we present measurements of the enthalpy recovery of polystyrene at an aging temperature 15 °C below the nominal Tg, for aging times up to 1 year. The results indicate that the equilibrium liquid enthalpy line can indeed be reached for aging 15 K below Tg. The results are analyzed in the context of the TNM model of structural recovery.}, number={14}, journal={Macromolecules}, author={Koh, Y.P. and Simon, S.L.}, year={2013}, pages={5815–5821} } @article{begum_sarker_simon_2013, title={Modeling Ring/Chain Equilibrium in Nanoconfined Sulfur}, volume={117}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000317552700023&KeyUID=WOS:000317552700023}, DOI={10.1021/jp400145n}, abstractNote={The effect of nanoconfinement on the thermodynamics of free radical polymerization of sulfur is examined. We extend Tobolsky and Eisenberg's model of bulk sulfur polymerization to nanopores accounting for the confinement entropy of the chains and ring using scaling reported in literature. The model quantitatively captures literature data from Yannopoulos and co-workers for the extent of polymerization versus temperature for bulk sulfur polymerization and for polymerization in 20, 7.5, and 2.5 nm diameter Gelsil nanopores, assuming that the change of entropy of nanoconfined chains scales with molecular size to the second power and with nanopore diameter to either the −3.0 or −3.8 power, the former of which fits slightly better. The scaling, which is valid for strong confinement in spherical pores, predicts that the propagation equilibrium constant will depend on both nanopore size and chain length, such that the average chain length decreases significantly upon confinement.}, number={14}, journal={Journal of Physical Chemistry B}, author={Begum, F. and Sarker, R. H. and Simon, S. L.}, year={2013}, pages={3911–3916} } @article{zhao_koh_pyda_sen_simon_2013, title={The kinetics of the glass transition and physical aging in germanium selenide glasses}, volume={368}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000319103700010&KeyUID=WOS:000319103700010}, DOI={10.1016/j.jnoncryso1.2013.02.025}, journal={Journal of Non-Crystalline Solids}, author={Zhao, H. Y. and Koh, Y. P. and Pyda, M. and Sen, S. and Simon, S. L.}, year={2013}, pages={63–70} } @article{zhao_koh_pyda_sen_simon_2013, title={The kinetics of the glass transition and physical aging in germanium selenide glasses}, volume={368}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84875853856&partnerID=MN8TOARS}, DOI={10.1016/j.jnoncrysol.2013.02.025}, number={1}, journal={Journal of Non-Crystalline Solids}, author={Zhao, H.Y. and Koh, Y.P. and Pyda, M. and Sen, S. and Simon, S.L.}, year={2013}, pages={63–70} } @article{zhao_simon_mckenna_2013, title={Using 20-million-year-old amber to test the super-Arrhenius behaviour of glass-forming systems}, volume={4}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000318872100140&KeyUID=WOS:000318872100140}, DOI={10.1038/ncomms2809}, journal={Nature Communications}, author={Zhao, J. and Simon, S. L. and McKenna, G. B.}, year={2013} } @article{guo_grassia_simon_2012, title={Bulk and shear rheology of a symmetric three-arm star polystyrene}, volume={50}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000306956600003&KeyUID=WOS:000306956600003}, DOI={10.1002/polb.23113}, abstractNote={Abstract The bulk and shear rheological properties of a symmetric three‐arm star polystyrene were measured using a self‐built pressurizable dilatometer and a commercial rheometer, respectively. The bulk properties investigated include the pressure–volume–temperature behavior, the pressure‐dependent glass transition temperature ( T g ), and the viscoelastic bulk modulus and Poisson's ratio. Comparison with data for a linear polystyrene indicates that the star behaves similarly but with slightly higher T g s at elevated pressures and slightly higher limiting bulk moduli in glass and rubbery states. The Poisson's ratio shows a minimum at short times similar to what is observed for the linear chain. The horizontal shift factors above T g obtained from reducing the bulk and shear viscoelastic responses are found to have similar temperature dependence when plotted using T − T g scaling; in addition, the shift factors also exhibit a similar temperature dependence to linear polystyrene. The retardation spectra for the bulk and shear responses are compared and show that the long time molecular mechanisms available to the shear response are unavailable to the bulk. At short times, the two spectra have similar slopes, but the short‐time retardation spectrum for the shear response is significantly higher than that for the bulk, a finding that is, as yet, unexplained. © 2012 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys, 2012}, number={17}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Guo, J. X. and Grassia, L. and Simon, S. L.}, year={2012}, pages={1233–1244} } @article{koh_simon_2012, title={Crystallization and Vitrification of a Cyanurate Trimer in Nanopores}, volume={116}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000305933800017&KeyUID=WOS:000305933800017}, DOI={10.1021/jp303318e}, abstractNote={The effects of nanopore confinement on the crystallization and vitrification of a low molecular weight organic material, tris(4-cumylphenol)-1,3,5-triazine, are investigated using differential scanning calorimetry. The material shows cold crystallization and subsequent melting in the bulk state. Under the nanoconfinement of controlled pore glasses (CPG), cold crystallization and melting shift to lower temperatures. Crystallization kinetics are hindered in nanoconfinement, and no crystallization occurs in 13 nm diameter pores over the course of a week. Using a traditional Avrami analysis, the restricted crystallization under nanopore confinement is quantified; for crystallization at 80 °C, the Avrami exponent decreases with decreasing pore size and the overall crystallization rate is approximately 30 times slower for material confined in 50 nm diameter pores than the bulk. When compared at the temperature at which the crystallization rate is a maximum, the Avrami exponent is higher in nanoconfined samples and the crystallization rate is approximately 10 times slower for material confined in 50 nm diameter pores. Under CPG nanoconfinement, the glass transition temperature also decreases and shows two values; interestingly, the T(g) values further decrease with increasing crystallinity.}, number={26}, journal={Journal of Physical Chemistry B}, author={Koh, Y. P. and Simon, S. L.}, year={2012}, pages={7754–7761} } @article{koh_karim_simon_2012, title={Heterogeneous reaction kinetics of epoxide-functionalized regenerated cellulose membrane and aliphatic amine}, volume={543}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/j.tca.2012.04.031}, DOI={10.1016/j.tca.2012.04.031}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Koh, Yung P. and Karim, M. Nazmul and Simon, Sindee L.}, year={2012}, month={Sep}, pages={18–23} } @article{begum_zhao_simon_2012, title={Modeling methyl methacrylate free radical polymerization: Reaction in hydrophilic nanopores}, volume={53}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000306151100021&KeyUID=WOS:000306151100021}, DOI={10.1016/j.polymer.2012.05.023}, number={15}, journal={Polymer}, author={Begum, F. and Zhao, H. Y. and Simon, S. L.}, year={2012}, pages={3238–3244} } @article{begum_zhao_simon_2012, title={Modeling methyl methacrylate free radical polymerization: Reaction in hydrophobic nanopores}, volume={53}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000306151100024&KeyUID=WOS:000306151100024}, DOI={10.1016/j.polymer.2012.04.036}, number={15}, journal={Polymer}, author={Begum, F. and Zhao, H. and Simon, S. L.}, year={2012}, pages={3261–3268} } @article{grassia_simon_2012, title={Modeling volume relaxation of amorphous polymers: Modification of the equation for the relaxation time in the KAHR model}, volume={53}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000306825400024&KeyUID=WOS:000306825400024}, DOI={10.1016/j.polymer.2012.06.013}, number={16}, journal={Polymer}, author={Grassia, L. and Simon, S. L.}, year={2012}, pages={3613–3620} } @article{zheng_mohammed_hines_xiao_martinez_bartsch_simon_russina_triolo_quitevis_2011, title={Effect of Cation Symmetry on the Morphology and Physicochemical Properties of Imidazolium Ionic Liquids}, volume={115}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000290652100020&KeyUID=WOS:000290652100020}, DOI={10.1021/jp1115614}, abstractNote={In this paper, the morphology and bulk physical properties of 1,3-dialkylimidazolium bis{(trifluoromethane)sulfonyl}amide ([(C(N/2))(2)im][NTf(2)]) are compared to that of 1-alkyl-3-methylimidazolium bis{(trifluoromethane)sulfonyl}amide ([C(N-1)C(1)im][NTf(2)]) for N = 4, 6, 8, and 10. For a given pair of ionic liquids (ILs) with the same N, the ILs differ only in the symmetry of the alkyl substitution on the imidazolium ring of the cation. Small-wide-angle X-ray scattering measurements indicate that, for a given symmetric/asymmetric IL pair, the structural heterogeneities are larger in the asymmetric IL than in the symmetric IL. The correlation length of structural heterogeneities for the symmetric and asymmetric salts, however, is described by the same linear equation when plotted versus the single alkyl chain length. Symmetric ILs with N = 4 and 6 easily crystallize, whereas longer alkyl chains and asymmetry hinder crystallization. Interestingly, the glass transition temperature is found to vary inversely with the correlation length of structural heterogeneities and with the length of the longest alkyl chain. Whereas the densities for a symmetric/asymmetric IL pair with a given N are nearly the same, the viscosity of the asymmetric IL is greater than that of the symmetric IL. Also, an even-odd effect previously observed in molecular dynamics simulations is confirmed by viscosity measurements. We discuss in this paper how the structural heterogeneities and physical properties of these ILs are consistent with alkyl tail segregation.}, number={20}, journal={Journal of Physical Chemistry B}, author={Zheng, W. and Mohammed, A. and Hines, L. G. and Xiao, D. and Martinez, O. J. and Bartsch, R. A. and Simon, S. L. and Russina, O. and Triolo, A. and Quitevis, E. L.}, year={2011}, pages={6572–6584} } @article{koh_simon_2012, title={Erratum: Kinetic study of trimerization of monocyanate ester in nanopores (The Journal of Physical Chemistry B (2011) 115 (925-932) DOI: 10.1021/jp110192g)}, volume={116}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84855833437&partnerID=MN8TOARS}, DOI={10.1021/jp210037z}, abstractNote={ADVERTISEMENT RETURN TO ISSUEPREVAddition/CorrectionNEXTORIGINAL ARTICLEThis notice is a correctionCorrection to "Kinetic Study of Trimerization of Monocyanate Ester in Nanopores"Yung P. Koh and Sindee L. Simon*Cite this: J. Phys. Chem. B 2012, 116, 1, 731Publication Date (Web):December 29, 2011Publication History Published online29 December 2011Published inissue 12 January 2012https://pubs.acs.org/doi/10.1021/jp210037zhttps://doi.org/10.1021/jp210037zcorrectionACS PublicationsCopyright © 2011 American Chemical Society. This publication is available under these Terms of Use. Request reuse permissions This publication is free to access through this site. Learn MoreArticle Views272Altmetric-Citations-LEARN ABOUT THESE METRICSArticle Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated. Share Add toView InAdd Full Text with ReferenceAdd Description ExportRISCitationCitation and abstractCitation and referencesMore Options Share onFacebookTwitterWechatLinked InRedditEmail PDF (127 KB) Get e-Alertsclose Get e-Alerts}, number={1}, journal={Journal of Physical Chemistry B}, author={Koh, Y.P. and Simon, S.L.}, year={2012} } @article{koh_simon_2011, title={Kinetic Study of Trimerization of Monocyanate Ester in Nanopores}, volume={115}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000286797700018&KeyUID=WOS:000286797700018}, DOI={10.1021/jp110192g}, abstractNote={A kinetic study of the trimerization of monocyanate ester both in the bulk and in the nanoconfinement of controlled pore glass is performed using differential scanning calorimetry. Both isothermal and dynamic experiments are analyzed. Although the activation energy for the reaction is the same within experimental error for the bulk and nanoconfined samples (approximately 21-23 kcal/mol), the reaction is accelerated under nanoconfinement by approximately 50 times in 13 nm pores compared with bulk.}, number={5}, journal={Journal of Physical Chemistry B}, author={Koh, Y. P. and Simon, S. L.}, year={2011}, pages={925–932} } @article{zhao_simon_2011, title={Methyl methacrylate polymerization in nanoporous confinement}, volume={52}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000294316200024&KeyUID=WOS:000294316200024}, DOI={10.1016/j.polymer.2011.06.048}, number={18}, journal={Polymer}, author={Zhao, H. Y. and Simon, S. L.}, year={2011}, pages={4093–4098} } @article{begum_simon_2011, title={Modeling methyl methacrylate free radical polymerization in nanoporous confinement}, volume={52}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-79952624300&partnerID=MN8TOARS}, DOI={10.1016/j.polymer.2011.02.009}, number={7}, journal={Polymer}, author={Begum, F. and Simon, S.L.}, year={2011}, pages={1539–1545} } @article{guo_simon_2011, title={Thermodynamic scaling of polymer dynamics versus T - T-g scaling}, volume={135}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000294065200037&KeyUID=WOS:000294065200037}, DOI={10.1063/1.3624903}, abstractNote={A thermodynamic scaling law for the relaxation times of complex liquids as a function of temperature and volume has been proposed in the literature: τ(T,V) = ƒ(TVγ), where γ is a material-dependent constant. We test this scaling for six materials, linear polystyrene, star polystyrene, two polycyanurate networks, poly(vinyl acetate), and poly(vinyl chloride), and compare the thermodynamic scaling to T – Tg scaling, where τ = ƒ(T – Tg). The thermodynamic scaling law successfully reduces the data for all of the samples; however, polymers with similar structures but different glass transition (Tg) and pressure-volume-temperature (PVT) behavior, i.e., the two polycyanurates, cannot be superposed unless the scaling law is normalized by TgVgγ. On the other hand, the T – Tg scaling successfully reduced data for all polymers, including those having similar microstructures. In addition, the T – Tg scaling is easier to implement since it does not require knowledge of the PVT behavior of the material. The relationship between TgVgγ/TVγ and T – Tg scaling is clarified and is found to be weakly dependent on pressure.}, number={7}, journal={Journal of Chemical Physics}, author={Guo, J. X. and Simon, S. L.}, year={2011} } @article{simon_2010, title={Journal of Thermal Analysis and Calorimetry: Preface}, volume={102}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-78049463143&partnerID=MN8TOARS}, DOI={10.1007/s10973-010-0990-7}, number={2}, journal={Journal of Thermal Analysis and Calorimetry}, author={Simon, S.}, year={2010} } @article{grassia_d’amore_simon_2010, title={On the viscoelastic Poisson's ratio in amorphous polymers}, volume={54}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000282472100004&KeyUID=WOS:000282472100004}, DOI={10.1122/1.3473811}, abstractNote={Views Icon Views Article contents Figures & tables Video Audio Supplementary Data Peer Review Share Icon Share Twitter Facebook Reddit LinkedIn Tools Icon Tools Reprints and Permissions Cite Icon Cite Search Site Citation Luigi Grassia, Alberto D’Amore, Sindee L. Simon; On the viscoelastic Poisson’s ratio in amorphous polymers. Journal of Rheology 1 September 2010; 54 (5): 1009–1022. https://doi.org/10.1122/1.3473811 Download citation file: Ris (Zotero) Reference Manager EasyBib Bookends Mendeley Papers EndNote RefWorks BibTex toolbar search Search Dropdown Menu toolbar search search input Search input auto suggest filter your search All ContentThe Society of RheologyJournal of Rheology Search Advanced Search |Citation Search}, number={5}, journal={Journal of Rheology}, author={Grassia, L. and D’Amore, A. and Simon, S. L.}, year={2010}, pages={1009–1022} } @article{guo_simon_2010, title={Pressure-Volume-Temperature Behavior of Two Polycyanurate Networks}, volume={48}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000284227500011&KeyUID=WOS:000284227500011}, DOI={10.1002/polb.22155}, abstractNote={Abstract The pressure–volume–temperature (PVT) behavior was studied for two polycyanurate networks having different crosslink densities using a pressurizable dilatometer. The samples were studied at temperatures ranging from 60 to 180 °C and at pressures up to 170 MPa to yield PVT data in both rubbery and glassy states. The Tait equation is found to well describe the isobaric temperature scan and isothermal pressure scan data. The thermal expansion coefficients, instantaneous bulk moduli, and thermal pressure coefficients are extracted from the data and their dependence on crosslink density is examined. The time‐dependent viscoelastic bulk modulus ( K ( t )) is also calculated in the vicinity of the α‐relaxation from previously published pressure relaxation experimental data, and the strength and shape of the dispersion are found to be independent of crosslink density. The limiting bulk moduli depend strongly on temperature with those of the more loosely crosslinked sample being lower at a given temperature and pressure, although at T g ( P ), the limiting moduli of the more loosely crosslinked sample are slightly higher than those of the more highly crosslinked sample. © 2010 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys, 2010}, number={23}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Guo, J. X. and Simon, S. L.}, year={2010}, pages={2509–2517} } @article{zheng_mckenna_simon_2010, title={The viscoelastic behavior of polymer/oligomer blends}, volume={51}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000282354500029&KeyUID=WOS:000282354500029}, DOI={10.1016/j.polymer.2010.08.046}, number={21}, journal={Polymer}, author={Zheng, W. and McKenna, G. B. and Simon, S. L.}, year={2010}, pages={4899–4906} } @article{koh_simon_2010, title={Trimerization of Monocyanate Ester in Nanopores}, volume={114}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77953389823&partnerID=MN8TOARS}, DOI={10.1021/jp912235c}, abstractNote={The effects of nanoconfinement on the reaction kinetics and properties of a monocyanate ester and the resulting cyanurate trimer are studied using differential scanning calorimetry (DSC). On the basis of both dynamic heating scans and isothermal reaction studies, the reaction rate is found to increase with decreasing nanopore size without a change in reaction mechanism. Both the monocyanate ester reactant and cyanurate product show reduced glass transition temperatures (Tgs) as compared to the bulk; the Tg depression increases with conversion and is more pronounced for the fully reacted product, suggesting that molecular stiffness influences the magnitude of nanoconfinement effects. Our results are consistent with the accelerated reaction and the Tg depression found previously for the nanoconfined difunctional cyanate ester, supporting the supposition that intracyclization is not the origin of these effects.}, number={23}, journal={Journal of Physical Chemistry B}, author={Koh, Y.P. and Simon, S.L.}, year={2010}, pages={7727–7734} } @inproceedings{zheng_mckenna_simon_2010, title={Viscoelasticity of polymer/oligomer athermal blends}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77956904901&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Zheng, W. and McKenna, G.B. and Simon, S.L.}, year={2010}, pages={1901–1904} } @article{meng_bernazzani_o’connell_mckenna_simon_2009, title={A new pressurizable dilatometer for measuring the time-dependent bulk modulus and pressure-volume-temperature properties of polymeric materials}, volume={80}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000266442500032&KeyUID=WOS:000266442500032}, DOI={10.1063/1.3122964}, abstractNote={A new piston-cylinder-type pressurizable dilatometer controlled by a stepper motor has been developed to measure the time-dependent bulk modulus and pressure-volume-temperature (PVT) behavior of polymeric materials. The dilatometer can be operated from 35 to 230 degrees C and at pressures of up to 250 MPa. The sample cell, which contains the sample and a fluorinated oil as the confining fluid, is totally submerged into a high precision oil bath to achieve a temperature stability of better than 0.01 degrees C. The instrument is calibrated with mercury and quartz. The total instrument volume is 4.0 cm(3), of which 2.3 cm(3) is the sample cell; the total volume can be measured with an average absolute error of better than 5.0x10(-4) cm(3). To demonstrate the instrument's capabilities, the time-dependent bulk modulus and the PVT behavior of a polystyrene are obtained and compared to the literature.}, number={5}, journal={Review of Scientific Instruments}, author={Meng, Y. and Bernazzani, P. and O’Connell, P. A. and McKenna, G. B. and Simon, S. L.}, year={2009} } @article{dalle-ferrier_simon_zheng_badrinarayanan_fennell_frick_zanotti_alba-simionesco_2009, title={Consequence of Excess Configurational Entropy on Fragility: The Case of a Polymer-Oligomer Blend}, volume={103}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000271352400025&KeyUID=WOS:000271352400025}, DOI={10.1103/PhysRevLett.103.185702}, abstractNote={By taking advantage of the molecular weight dependence of the glass transition of polymers and their ability to form perfectly miscible blends, we propose a way to modify the fragility of a system, from fragile to strong, keeping the same glass properties, i.e., vibrational density of states, mean-square displacement, and local structure. Both slow and fast dynamics are investigated by calorimetry and neutron scattering in an athermal polystyrene-oligomer blend, and compared to those of a pure 17-mer polystyrene considered to be a reference, of the same ${T}_{g}$. Whereas the blend and the pure 17-mer have the same heat capacity in the glass and in the liquid, their fragilities differ strongly. Thus, the difference in fragility is related to an extra configurational entropy created by the mixing process and acting at a scale much larger than the interchain distance, without affecting the fast dynamics and the structure of the glass.}, number={18}, journal={Physical Review Letters}, author={Dalle-Ferrier, C. and Simon, S. and Zheng, W. and Badrinarayanan, P. and Fennell, T. and Frick, B. and Zanotti, J. M. and Alba-Simionesco, C.}, year={2009} } @article{guo_simon_2009, title={Effect of Crosslink Density on the Pressure Relaxation Response of Polycyanurate Networks}, volume={47}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000272561100007&KeyUID=WOS:000272561100007}, DOI={10.1002/polb.21860}, abstractNote={Abstract The effect of crosslink density on the pressure‐volume‐temperature (PVT) behavior and on the pressure relaxation response for two polycyanurate networks is investigated using a custom‐built pressurizable dilatometer. Isobaric cooling measurements were made to obtain the pressure‐dependent glass transition temperature ( T g ). The pressure relaxation studies were carried out as a function of time after volume jumps at temperatures in the vicinity of the pressure‐dependent T g , and the pressure relaxation curves obtained were shifted to construct master curves by time‐temperature superposition. The reduced pressure relaxation curves are found to be identical in shape and placement, independent of crosslink density, when T g is used as the reference temperature. The horizontal shift factors used to create the master curves are plotted as a function of the temperature departure from T g ( T − T g ), and they agree well with their counterparts obtained from the shear response. Moreover, the retardation spectra are derived from bulk compliance and compared to those from the shear. The results, similar to our previous work on polystyrene, indicate that at short times, the bulk and shear responses have similar underlying molecular mechanisms; however, the long‐time mechanisms available to the shear response, which increase with decreasing crosslink density, are unavailable to the bulk response. © 2009 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 2477–2486, 2009}, number={24}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Guo, J. X. and Simon, S. L.}, year={2009}, pages={2477–2486} } @article{simon_mckenna_2009, title={Experimental evidence against the existence of an ideal glass transition}, volume={355}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000266617100013&KeyUID=WOS:000266617100013}, DOI={10.1016/j.jnoncrysol.2008.11.027}, number={10-12}, journal={Journal of Non-Crystalline Solids}, author={Simon, S. L. and McKenna, G. B.}, year={2009}, pages={672–675} } @article{li_simon_2009, title={Surface Chemistry Effects on the Reactivity and Properties of Nanoconfined Bisphenol M Dicyanate Ester in Controlled Pore Glass}, volume={42}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000266200800020&KeyUID=WOS:000266200800020}, DOI={10.1021/ma802808v}, abstractNote={Nanoconfinement has been found to have significant effect on the glass transition behavior of low molecular weight and polymeric glass formers. Here, we investigate the influence of nanoconfinement on the cure kinetics and glass transition temperature of a bisphenol M dicyanate ester/polycyanurate material as a function of surface chemistry and nanoconfinement size in native and silanized controlled pore glasses (CPGs). The glass transition temperature and conversion as a function of cure time are examined using differential scanning calorimetry (DSC). The native CPG surface accelerates the cure of bisphenol M dicyanate to a larger extent compared to the silanized hydrophobic CPG presumably because of the catalytic nature of hydroxyl groups on the CPG wall. Two Tgs are observed for both monomer and polycyanurates confined in the native CPGs. The primary Tg of the "fully cured" polycyanurate is depressed by 60 K at 11.5 nm, and the secondary Tg is 10−33 K above the primary Tg; the values are similar to those found previously in silanized CPGs. The length scale associated with the secondary Tg is ∼0.90 nm assuming that the secondary Tg reflects the material at the CPG wall surface. Based on the measurements of Tg, the total heat capacity change at Tg, and the sol content, all as a function of conversion, the network structure does not change upon nanoconfinement.}, number={10}, journal={Macromolecules}, author={Li, Q. X. and Simon, S. L.}, year={2009}, pages={3573–3579} } @article{koh_li_simon_2009, title={T-g and reactivity at the nanoscale}, volume={492}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000268646900006&KeyUID=WOS:000268646900006}, DOI={10.1016/j.tca.2009.06.007}, number={1-2}, journal={Thermochimica Acta}, author={Koh, Y. P. and Li, Q. X. and Simon, S. L.}, year={2009}, pages={45–50} } @article{win_karim_li_simon_mckenna_2010, title={Thermal Pressure Coefficient of a Polyhedral Oligomeric Silsesquioxane (POSS)-Reinforced Epoxy Resin}, volume={116}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000274545000017&KeyUID=WOS:000274545000017}, DOI={10.1002/app.31502}, abstractNote={Abstract The thermal pressure coefficients of a neat, unfilled, epoxy resin and a 10 wt % POSS (polyhedral oligomeric silsesquioxane)‐filled epoxy nanocomposite have been measured using a thick‐walled tube method. It is found that just below the glass transition temperature the thermal pressure coefficient is ∼ 20% smaller for the polymer composite containing 10% POSS than for the neat, unfilled resin. The thermal expansion coefficient and thermal pressure coefficient of the uncured POSS itself are also reported. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010}, number={1}, journal={Journal of Applied Polymer Science}, author={Win, K. Z. and Karim, T. and Li, Q. X. and Simon, S. L. and McKenna, G. B.}, year={2010}, pages={142–146} } @article{li_simon_2008, title={Curing of bisphenol M dicyanate ester under nanoscale constraint}, volume={41}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000253331200037&KeyUID=WOS:000253331200037}, DOI={10.1021/ma702144b}, abstractNote={Nanoscale constraint is known to have a significant impact on the thermal properties of materials. Although thermosetting resins have been cured in the presence of nanoparticles and nanotubes, cure of thermosetting resins under the well-defined nanoscale constraints imposed by controlled pore glass (CPG) or similar matrices has not been previously documented. In this work, we investigate the isothermal curing under nanoscale constraint of a thermosetting resin, bisphenol M dicyanate ester (BMDC), which trimerizes to form a polycyanurate network material. Differential scanning calorimetry is used to monitor the evolution of the glass transition temperature (Tg) and the conversion during cure as a function of the diameter of the silanized control pore glass matrix which is used for confinement. A Tg depression is observed for both the bisphenol M dicyanate ester monomer and the polycyanurate networks; the depression is only a few degrees for the monomer, whereas a 56 K depression is observed for the "fully cured" network in 11.5 nm pores. The nanoscale constraint is also found to strongly increase the rate of cure of bisphenol M dicyanate ester, but it does not affect the normalized Tg vs conversion relationship. The appearance of a secondary Tg above the primary Tg in the smaller pores and the associated length scale are discussed.}, number={4}, journal={Macromolecules}, author={Li, Q. X. and Simon, S. L.}, year={2008}, pages={1310–1317} } @article{effect of structure on enthalpy relaxation of polycarbonate: experiments and modeling_2008, volume={49}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000258437100025&KeyUID=WOS:000258437100025}, DOI={10.1016/j.polymer.2008.05.046}, number={16}, journal={Polymer}, year={2008}, pages={3554–3560} } @article{koh_simon_2008, title={Structural Relaxation of Stacked Ultrathin Polystyrene Films}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000261705400011&KeyUID=WOS:000261705400011}, DOI={10.1002/polb.21598}, abstractNote={Abstract The T g depression and kinetic behavior of stacked polystyrene ultrathin films is investigated by differential scanning calorimetry (DSC) and compared with the behavior of bulk polystyrene. The fictive temperature ( T f ) was measured as a function of cooling rate and as a function of aging time for aging temperatures below the nominal glass transition temperature ( T g ). The stacked ultrathin films show enthalpy overshoots in DSC heating scans which are reduced in height but occur over a broader temperature range relative to the bulk response for a given change in fictive temperature. The cooling rate dependence of the limiting fictive temperature, T f ′, is also found to be higher for the stacked ultrathin film samples; the result is that the magnitude of the T g depression between the ultrathin film sample and the bulk is inversely related to the cooling rate. We also find that the rate of physical aging of the stacked ultrathin films is comparable with the bulk when aging is performed at the same distance from T g ; however, when conducted at the same aging temperature, the ultrathin film samples show accelerated physical aging, that is, a shorter time is required to reach equilibrium for the thin films due to their depressed T g values. The smaller distance from T g also results in a reduced logarithmic aging rate for the thin films compared with the bulk, although this is not indicative of longer relaxation times. The DSC heating curves obtained as a function of cooling rate and aging history are modeled using the Tool‐Narayanaswamy‐Moynihan model of structural recovery; the stacked ultrathin film samples show lower β values than the bulk, consistent with a broader distribution of relaxation times. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 46: 2741–2753, 2008}, number={24}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Koh, Y. P. and Simon, S. L.}, year={2008}, pages={2741–2753} } @inproceedings{koh_simon_2008, title={The glass transition and kinetics in stacked polystyrene ultrathin films}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-52349117603&partnerID=MN8TOARS}, booktitle={Technical Papers, Regional Technical Conference - Society of Plastics Engineers}, author={Koh, Y.P. and Simon, S.L.}, year={2008}, pages={1668–1673} } @article{li_hutcheson_mckenna_simon_2008, title={Viscoelastic Properties and Residual Stresses in Polyhedral Oligomeric Silsesquioxane-Reinforced Epoxy Matrices}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000261705400009&KeyUID=WOS:000261705400009}, DOI={10.1002/polb.21609}, abstractNote={Abstract Fiber‐filled thermosetting polymer composites are extensively used in aerospace industries. One disadvantage of these materials is cure induced or thermally induced residual stresses in the matrix, which may result in deteriorated performance and premature failure. This article explores the use of epoxy/multifunctional polyhedral oligomeric silsesquioxane (POSS) nanocomposites as resins with reduced thermal stress coefficients that result in mitigated residual stresses. The effect of POSS loading on the thermal stress coefficient of the epoxy/POSS nanocomposite resins was investigated from below the β‐relaxation to the α‐relaxation, or glass transition temperature, (i.e., from −100 to 180 °C) by measuring the shear modulus and linear thermal expansion coefficient. The thermal stress coefficient of the epoxy/POSS nanocomposites is found to be a strong function of temperature, decreasing rapidly with decreasing temperature through the α‐relaxation region, increasing in the vicinity of the β‐relaxation, and then decreasing below the temperature associated with the peak in the β‐relaxation. With increasing POSS content, the thermal stress coefficient is reduced compared with the neat resin in the vicinity of the α‐relaxation; however, the thermal stress coefficient increases with increasing POSS content below the temperature of the β‐relaxation peak. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 46: 2719–2732, 2008}, number={24}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Li, Q. X. and Hutcheson, S. A. and McKenna, G. B. and Simon, S. L.}, year={2008}, pages={2719–2732} } @inproceedings{li_hutcheson_mckenna_simon_2008, title={Viscoelastic properties and residual stresses in polyhedral oligomeric silsequioxane (POSS)-reinforced epoxy}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-52349097009&partnerID=MN8TOARS}, booktitle={Technical Papers, Regional Technical Conference - Society of Plastics Engineers}, author={Li, Q. and Hutcheson, S.A. and McKenna, G.B. and Simon, S.L.}, year={2008}, pages={1923–1929} } @article{zheng_simon_2007, title={Confinement effects on the glass transition of hydrogen bonded liquids}, volume={127}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000251106100015&KeyUID=WOS:000251106100015}, DOI={10.1063/1.2793787}, abstractNote={The glass transition behavior of glycerol and propylene glycol confined in nanoporous glass is investigated using differential scanning calorimetry. Both silanized and unsilanized porous glasses are used to confine the liquids with nominal pore sizes ranging from 2.5 to 7.5 nm, and the glass transition temperature (T(g)) and the limiting fictive temperature (T(f )') are measured on cooling and heating, respectively. The effect of pore fullness is also examined. We find that differences in T(g), DeltaC(p), and the enthalpy overshoot behavior observed on heating are significant between partially and completely filled pores for the case of the unsilanized controlled pore glasses (CPGs) but that the effect of pore fullness is insignificant for the silanized CPGs. In general, the behavior in the silanized CPGs is similar to the behavior in the completely filled unsilanized pores. For glycerol, this includes a small depression in T(f )' on the order of 5 K at 2.5 nm. For propylene glycol, similar behavior is found except that an additional glass transition is observed in both silanized and unsilanized systems approximately 30 K higher than the bulk and a slightly smaller depression on the order of 3 K at 2.5 nm is observed in the completely filled unsilanized pores and in partially and completely filled silanized pores. The results are compared to those in the literature, and the confinement effects are discussed.}, number={19}, journal={Journal of Chemical Physics}, author={Zheng, W. and Simon, S. L.}, year={2007} } @inproceedings{li_simon_2007, title={Curing of bisphenol M dicyanate ester under nanoscale constraint}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-34648840786&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Li, Q. and Simon, S.L.}, year={2007}, pages={1750–1755} } @article{li_li_simon_guven_borges_youan_2007, title={Formulation of spray-dried phenytoin loaded poly(epsilon-caprolactone) microcarrier intended for brain delivery to treat epilepsy}, volume={96}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000246132800008&KeyUID=WOS:000246132800008}, DOI={10.1002/jps.20935}, number={5}, journal={Journal of Pharmaceutical Sciences}, author={Li, Z. Z. and Li, Q. X. and Simon, S. and Guven, N. and Borges, K. and Youan, B. B. C.}, year={2007}, pages={1018–1030} } @article{badrinarayanan_zheng_simon_2008, title={Isoconversion analysis of the glass transition}, volume={468}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000254266300015&KeyUID=WOS:000254266300015}, DOI={10.1016/j.tca.2007.11.020}, number={1-2}, journal={Thermochimica Acta}, author={Badrinarayanan, P. and Zheng, W. and Simon, S. L.}, year={2008}, pages={87–93} } @inproceedings{badrinarayanan_zheng_simon_2007, title={Isoconversion analysis of the glass transition}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-34648840782&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Badrinarayanan, P. and Zheng, W. and Simon, S.L.}, year={2007}, pages={1756–1760} } @inproceedings{simon_2007, title={Measurement of the glass transition temperature}, volume={5}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-34648846828&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, S.L.}, year={2007}, pages={2617–2620} } @article{badrinarayanan_simon_2007, title={Origin of the divergence of the timescales for volume and enthalpy recovery}, volume={48}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000245257300006&KeyUID=WOS:000245257300006}, DOI={10.1016/j.polymer.2007.01.064}, number={6}, journal={Polymer}, author={Badrinarayanan, P. and Simon, S. L.}, year={2007}, pages={1464–1470} } @article{meng_simon_2007, title={Pressure relaxation of polystyrene and its comparison to the shear response}, volume={45}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000251556600019&KeyUID=WOS:000251556600019}, DOI={10.1002/polb.21320}, abstractNote={Abstract Isothermal pressure relaxation as a function of temperature in two pressure ranges has been measured for polystyrene using a self‐built pressurizable dilatometer. A master curve for pressure relaxation in each pressure regime is obtained based on the time–temperature superposition principle, and time–pressure superposition of the two master curves is found to be applicable when the master curves are referenced to their pressure‐dependent T g . The pressure relaxation master curves, the shift factors, and retardation spectra obtained from these curves are compared with those obtained from shear creep compliance measurements for the same material. The shift factors for the bulk and shear responses have the same temperature dependence, and the retardation spectra overlap at short times. Our results suggest that the bulk and shear response have similar molecular origin, but that long‐time chain mechanisms available to shear are lost in the bulk response. © 2007 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 45: 3375–3385, 2007}, number={24}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Meng, Y. and Simon, S. L.}, year={2007}, pages={3375–3385} } @inproceedings{zheng_simon_2007, title={Tg in polymer/oligomer athermal blends}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-34648835233&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Zheng, W. and Simon, S.L.}, year={2007}, pages={1788–1792} } @article{zheng_simon_2008, title={The glass transition in athermal poly(alpha-methyl styrene)/oligomer blends}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000253221400008&KeyUID=WOS:000253221400008}, DOI={10.1002/polb.21379}, abstractNote={Abstract The glass transition behavior in athermal blends of poly(α‐methyl styrene) (PaMS) and its hexamer is investigated using differential scanning calorimetry (DSC). The results, along with previous data on similar blends of PaMS/pentamer, are analyzed in the context of the Lodge–McLeish self‐concentration model. A methodology is described to partition the calorimetric transition to obtain effective T g s for each component of the blend. The dependences of these effective T g s on overall blend composition are described by the Lodge–McLeish model, although the self‐concentration effect is less than expected based on the Kuhn length. The length scales of the cooperatively rearranging regions for the two components in the blends are also calculated adapting Donth's fluctuation model to the partitioned DSC transitions and are found to be similar for the two components and show a slight decrease at intermediate concentrations. The kinetics associated with the glass temperature, T g , is examined by studying the cooling rate dependence of T g for the pure components and the blends, as well as by examining the enthalpy overshoots in the heating DSC scans. It is observed that the cooling rate dependence of T g in PaMS/hexamer blends at intermediate concentrations is similar to that of the hexamer, indicating that the kinetics of the glass transition for blends is dominated by the high mobility oligomeric component. Moreover, compared to the pure materials, the PaMS/hexamer blends exhibit a considerably depressed enthalpy overshoot, presumably resulting from their broader relaxation time distribution. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 46: 418–430, 2008}, number={4}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Zheng, W. and Simon, S. L.}, year={2008}, pages={418–430} } @article{badrinarayanan_zheng_li_simon_2007, title={The glass transition temperature versus the fictive temperature}, volume={353}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000248284400013&KeyUID=WOS:000248284400013}, DOI={10.1016/j.jnoncrysol.2007.04.025}, number={26}, journal={Journal of Non-Crystalline Solids}, author={Badrinarayanan, P. and Zheng, W. and Li, Q. X. and Simon, S. L.}, year={2007}, pages={2603–2612} } @article{sun_simon_2007, title={The melting behavior of aluminum nanoparticles}, volume={463}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-34548678652&partnerID=MN8TOARS}, DOI={10.1016/j.tca.2007.07.007}, number={1-2}, journal={Thermochimica Acta}, author={Sun, J. and Simon, S.L.}, year={2007}, pages={32–40} } @article{pitchimani_zheng_simon_hope-weeks_burnham_weeks_2007, title={Thermodynamic analysis of pure and impurity doped pentaerythritol tetranitrate crystals grown at room temperature}, volume={89}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000248172000021&KeyUID=WOS:000248172000021}, DOI={10.1007/s10973-006-8462-9}, number={2}, journal={Journal of Thermal Analysis and Calorimetry}, author={Pitchimani, R. and Zheng, W. and Simon, S. L. and Hope-Weeks, L. J. and Burnham, A. K. and Weeks, B. L.}, year={2007}, pages={475–478} } @article{li_simon_2007, title={Viscoelastic shear response and network structure in polycyanurates}, volume={40}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000244855500064&KeyUID=WOS:000244855500064}, DOI={10.1021/ma0624144}, abstractNote={The shear response of polycyanurate networks with different cross-link densities, varied by changing the ratio of difunctional to monofunctional cyanate ester, is measured from shear stress relaxation and dynamic experiments. Master curves are constructed following the time−temperature superposition principle, and the temperature dependence of the shift factors is examined. The discrete relaxation time spectra are calculated from the viscoelastic responses and are found to be independent of cross-link density over the time/frequency range measured. The cross-link density, determined from the rubbery modulus, and the sol content, measured from sol extraction experiments, are modeled for the fully cured polycyanurate networks using the recursive method; a monomer cyclization reaction is assumed in the modeling based upon the chemical composition of the sol which was determined by mass spectroscopy. The effect of monomer cyclization on the conversion at gelation of dicyanate esters is discussed.}, number={6}, journal={Macromolecules}, author={Li, Q. X. and Simon, S. L.}, year={2007}, pages={2246–2256} } @article{koh_mckenna_simon_2006, title={Calorimetric glass transition temperature and absolute heat capacity of polystyrene ultrathin films}, volume={44}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000242467100014&KeyUID=WOS:000242467100014}, DOI={10.1002/polb.21021}, abstractNote={Abstract The absolute heat capacity and glass transition temperature ( T g ) of unsupported ultrathin films were measured with differential scanning calorimetry with the step‐scan method in an effort to further examine the thermodynamic behavior of glass‐forming materials on the nanoscale. Films were stacked in layers with multiple preparation methods. The absolute heat capacity in both the glass and liquid states decreased with decreasing film thickness, and T g also decreased with decreasing film thickness. The magnitude of the T g depression was closer to that observed for films supported on rigid substrates than that observed for freely standing films. The stacked thin films regained bulk behavior after the application of pressure at a high temperature. The effects of various preparation methods were examined, including the use of polyisobutylene as an interleaving layer between the polystyrene films. © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 3518–3527, 2006}, number={24}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Koh, Y. P. and McKenna, G. B. and Simon, S. L.}, year={2006}, pages={3518–3527} } @article{sun_pantoya_simon_2006, title={Dependence of size and size distribution on reactivity of aluminum nanoparticles in reactions with oxygen and MoO3}, volume={444}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000238054500001&KeyUID=WOS:000238054500001}, DOI={10.1016/j.tca.2006.03.001}, number={2}, journal={Thermochimica Acta}, author={Sun, J. and Pantoya, M. L. and Simon, S. L.}, year={2006}, pages={117–127} } @article{li_simon_2006, title={Enthalpy recovery of polymeric glasses: Is the theoretical limiting liquid line reached?}, volume={47}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33746911952&partnerID=MN8TOARS}, DOI={10.1016/j.polymer.2006.04.046}, number={13}, journal={Polymer}, author={Li, Q. and Simon, S.L.}, year={2006}, pages={4781–4788} } @inproceedings{li_simon_2006, title={Enthalpy recovery: Do materials reach the equilibrium LINE?}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33745605220&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Li, Q. and Simon, S.L.}, year={2006}, pages={1674–1678} } @inproceedings{meng_simon_2006, title={Measurement of the bulk modulus using pressurizable dilatometry}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33745594834&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Meng, Y. and Simon, S.L.}, year={2006}, pages={1730–1734} } @article{zheng_simon_2006, title={Polystyrene freeze-dried from dilute solution: Tg depression and residual solvent effects}, volume={47}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33646153766&partnerID=MN8TOARS}, DOI={10.1016/j.polymer.2006.03.035}, number={10}, journal={Polymer}, author={Zheng, W. and Simon, S.L.}, year={2006}, pages={3520–3527} } @article{lahlouh_rajagopalan_biswas_sun_huang_simon_lubguban_gangopadhyay_2006, title={Post treatments of plasma-enhanced chemical vapor deposited hydrogenated amorphous silicon carbide for low dielectric constant films}, volume={497}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000234957300018&KeyUID=WOS:000234957300018}, DOI={10.1016/j.tsf.2005.10.085}, number={1-2}, journal={Thin Solid Films}, author={Lahlouh, B. and Rajagopalan, T. and Biswas, N. and Sun, J. and Huang, D. and Simon, S. L. and Lubguban, J. A. and Gangopadhyay, S.}, year={2006}, pages={109–114} } @article{simon_bernazzani_2006, title={Structural relaxation in the glass: Evidence for a path dependence of the relaxation time}, volume={352}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000242821800064&KeyUID=WOS:000242821800064}, DOI={10.1016/j.jnoncrysol.2006.01.151}, number={42-49}, journal={Journal of Non-Crystalline Solids}, author={Simon, S. L. and Bernazzani, P.}, year={2006}, pages={4763–4768} } @inproceedings{koh_mckenna_simon_2006, title={The calorimetric glass transition of free standing polystyrene thin films}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33745685815&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Koh, Y.P. and McKenna, G.B. and Simon, S.L.}, year={2006}, pages={1506–1509} } @inproceedings{badrinarayanan_zheng_simon_2006, title={The glass transition temperature versus the fictive temperature}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33745592787&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Badrinarayanan, P. and Zheng, W. and Simon, S.L.}, year={2006}, pages={1669–1673} } @inproceedings{meng_o’connell_mckenna_simon_2005, title={A new pressurizable dilatometer for measuring bulk modulus of thermosets}, volume={8}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33644995646&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Meng, Y. and O’Connell, P. and McKenna, G.B. and Simon, S.L.}, year={2005}, pages={335–339} } @inproceedings{huang_simon_mckenna_2005, title={Chain length dependence of heat capacity}, volume={8}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33644998802&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Huang, D. and Simon, S.L. and McKenna, G.B.}, year={2005}, pages={101–103} } @article{huang_simon_mckenna_2005, title={Chain length dependence of the thermodynamic properties of linear and cyclic alkanes and polymers}, volume={122}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000227372200082&KeyUID=WOS:000227372200082}, DOI={10.1063/1.1852453}, abstractNote={The specific heat capacity was measured with step-scan differential scanning calorimetry for linear alkanes from pentane (C(5)H(12)) to nonadecane (C(19)H(40)), for several cyclic alkanes, for linear and cyclic polyethylenes, and for a linear and a cyclic polystyrene. For the linear alkanes, the specific heat capacity in the equilibrium liquid state decreases as chain length increases; above a carbon number N of 10 (decane) the specific heat asymptotes to a constant value. For the cyclic alkanes, the heat capacity in the equilibrium liquid state is lower than that of the corresponding linear chains and increases with increasing chain length. At high enough molecular weights, the heat capacities of cyclic and linear molecules are expected to be equal, and this is found to be the case for the polyethylenes and polystyrenes studied. In addition, the thermal properties of the solid-liquid and the solid-solid transitions are examined for the linear and cyclic alkanes; solid-solid transitions are observed only in the odd-numbered alkanes. The thermal expansion coefficients and the specific volumes of the linear and cyclic alkanes are also calculated from literature data and compared with the trends in the specific heats.}, number={8}, journal={Journal of Chemical Physics}, author={Huang, D. H. and Simon, S. L. and McKenna, G. B.}, year={2005} } @article{merzlyakov_mckenna_simon_2006, title={Cure-induced and thermal stresses in a constrained epoxy resin}, volume={37}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000236338500009&KeyUID=WOS:000236338500009}, DOI={10.1016/j.compositesa.2005.05.013}, number={4}, journal={Composites Part a-Applied Science and Manufacturing}, author={Merzlyakov, M. and McKenna, G. B. and Simon, S. L.}, year={2006}, pages={585–591} } @article{merzlyakov_simon_mckenna_2005, title={Instrumented thick-walled tube method for measuring thermal pressure in fluids and isotropic stresses in thermosetting resins}, volume={76}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000229962000070&KeyUID=WOS:000229962000070}, DOI={10.1063/1.1930069}, abstractNote={We have developed a method for measuring the thermal pressure coefficient and cure-induced and thermally induced stresses based on an instrumented thick-walled tube vessel. The device has been demonstrated at pressures up to 330 MPa and temperatures to 300 °C. The method uses a sealed stainless steel thick-walled tube to impose three-dimensional isotropic constraints. The tube is instrumented with strain gauges in hoop and in axial directions and can be used in open or closed configurations. By making measurements of the isotropic stresses as a function of temperature, the method allows determination of the thermal pressure coefficient in both the glassy and rubbery (or liquid) states. The method also can be used to measure isotropic stress development in thermosetting resins during cure and subsequent thermal cycling. Experimental results are presented for sucrose benzoate, di-2-ethylhexylsebacate, and an epoxy resin. The current report shows that the method provides reliable estimates for the thermal pressure coefficient. The thermal pressure coefficient is determined with resolution on the order of 10kPa∕K. Among advantages of the method is that the tubes are reusable, even when measurements are made for cure response of thermosetting resins.}, number={6}, journal={Review of Scientific Instruments}, author={Merzlyakov, M. and Simon, S. L. and McKenna, G. B.}, year={2005} } @article{rajagopalan_lahlouh_lubguban_biswas_gangopadhyay_sun_huang_simon_toma_butler_2006, title={Investigation on hexamethyldisilazane vapor treatment of plasma-damaged nanoporous organosilicate films}, volume={252}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000239735100030&KeyUID=WOS:000239735100030}, DOI={10.1016/j.apsusc.2005.08.060}, number={18}, journal={Applied Surface Science}, author={Rajagopalan, T. and Lahlouh, B. and Lubguban, J. A. and Biswas, N. and Gangopadhyay, S. and Sun, J. and Huang, D. H. and Simon, S. L. and Toma, D. and Butler, R.}, year={2006}, pages={6323–6331} } @article{doshi_simon_2005, title={Modeling nanoporosity development in polymer films for low-k applications}, volume={45}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000228939900002&KeyUID=WOS:000228939900002}, DOI={10.1002/pen.20317}, abstractNote={The bubble growth dynamics of a polymer supersaturated with CO2 have been modeled for micron-size films after nucleation. The model equations are based on the shell model of Arefmanesh, Advani, and Michaelides in which a nucleated bubble is surrounded by a finite concentric shell of polymer supersaturated with gas. Bubbles grow by mass transfer of dissolved gas from this shell. The model is extended to allow for diffusion of dissolved gas out of the shell in addition to diffusion into the bubble. A parametric analysis is performed to examine the effects of film thickness, temperature, diffusivity at the Tg and Henry's law constant. POLYM. ENG. SCI., 45:640–651, 2005. © 2005 Society of Plastics Engineers}, number={5}, journal={Polymer Engineering and Science}, author={Doshi, P. and Simon, S.}, year={2005}, pages={640–651} } @article{meng_simon_2005, title={Relation between mobility factor and diffusion factor for thermoset cure}, volume={437}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000232322800029&KeyUID=WOS:000232322800029}, DOI={10.1016/j.tca.2005.06.032}, number={1-2}, journal={Thermochimica Acta}, author={Meng, Y. and Simon, S. L.}, year={2005}, pages={179–189} } @inproceedings{kolla_simon_2005, title={Volume recovery and the tau-effective paradox}, volume={6}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33645034617&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Kolla, S. and Simon, S.L.}, year={2005}, pages={284–288} } @inproceedings{merzlyakov_mckenna_simon_2004, title={Effect of cure cycle on residual stress development in thermosetting materials}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442597266&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Merzlyakov, M. and McKenna, G.B. and Simon, S.L.}, year={2004}, pages={2227–2230} } @inproceedings{simon_bernazzani_mckenna_2004, title={Effects of freeze-drying on the glass temperature of cyclic polystyrenes}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442655675&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, S.L. and Bernazzani, P. and McKenna, G.B.}, year={2004}, pages={2252–2257} } @article{merzlyakov_meng_simon_mckenna_2004, title={Instrumented sphere method for measuring thermal pressure in fluids and isotropic stresses and reaction kinetics in thermosetting resins}, volume={75}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000224755800036&KeyUID=WOS:000224755800036}, DOI={10.1063/1.1790584}, abstractNote={A novel technique is described for measuring thermal pressure in fluids and for measuring isotropic stress development and reaction kinetics in thermosetting resins during cure and thermal cycling. The method uses a 12.7-mm-diam sealed stainless steel spherical pressure vessel to impose three-dimensional isotropic constraints. The vessel is instrumented with strain gauges and thermocouples. Both isotropic stresses and reaction kinetics during cure at cure temperatures as high as 300 °C can be measured. In addition, measurement of the isotropic stress as a function of temperature yields the thermal pressure coefficient in both the glassy and rubbery (or liquid) states. Experimental results are presented for sucrose benzoate, a pressure-transmitting oil di-2-ethylhexylsebacate and an epoxy resin. The method provides reproducible estimates for the thermal pressure coefficient and the stresses are highly isotropic. A suggestion for improved versions of the device is: thicker walled vessels can be used to increase the upper stress limit (currently at 30 MPa). Also if a lower temperature range is to be studied, then aluminum can be used as a vessel material. Since epoxy resins have better adhesion to aluminum than to stainless steel, there may be an advantage to this.}, number={10}, journal={Review of Scientific Instruments}, author={Merzlyakov, M. and Meng, Y. and Simon, S. L. and McKenna, G. B.}, year={2004}, pages={3327–3334} } @inproceedings{meng_simon_2004, title={Relation between mobility and diffusion factors for thermoset cure}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442601546&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Meng, Y. and Simon, S.L.}, year={2004}, pages={2212–2216} } @article{lubguban_gangopadhyay_lahlouh_rajagopalan_biswas_sun_huang_simon_mallikarjunan_kim_et al._2004, title={Supercritical CO2 extraction of porogen phase: An alternative route to nanoporous dielectrics}, volume={19}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000224843700015&KeyUID=WOS:000224843700015}, DOI={10.1557/jmr.2004.0413}, number={11}, journal={Journal of Materials Research}, author={Lubguban, J. A. and Gangopadhyay, S. and Lahlouh, B. and Rajagopalan, T. and Biswas, N. and Sun, J. and Huang, D. H. and Simon, S. L. and Mallikarjunan, A. and Kim, H. C. and et al.}, year={2004}, pages={3224–3233} } @article{kolla_simon_2005, title={The tau-effective paradox: new measurements towards a resolution}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000226430600016&KeyUID=WOS:000226430600016}, DOI={10.1016/j.polymer.2004.11.086}, number={3}, journal={Polymer}, author={Kolla, S. and Simon, S. L.}, year={2005}, pages={733–739} } @article{alcoutlabi_mckenna_simon_2003, title={Analysis of the development of isotropic residual stresses in a bismaleimide/spiro orthocarbonate thermosetting resin for composite materials}, volume={88}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000180754900028&KeyUID=WOS:000180754900028}, DOI={10.1002/app.11649}, abstractNote={Abstract In this article, we extend our model of isotropic residual stress development in thermosets to a novel thermosetting resin system: bismaleimide/spiro orthocarbonate. In this system, the cure shrinkage and resulting isotropic residual stresses are reduced through a ring‐opening reaction that occurs independently of the addition reaction. The modeling effort includes a parametric analysis of the effects of various parameters, including the volume changes involved in the reactions, the relative rates and orders of the reactions, the cure history, and the values of the bulk moduli and thermal expansion coefficients. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 227–244, 2003}, number={1}, journal={Journal of Applied Polymer Science}, author={Alcoutlabi, M. and McKenna, G. B. and Simon, S. L.}, year={2003}, pages={227–244} } @inbook{lahlouh_rajagopalan_lubguban_biswas_gangopadhyay_sun_huang_simon_kim_volksen_et al._2003, title={Creating nanoporosity by selective extraction of porogens using supercritical carbon dioxide/cosolvent processes}, volume={766}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000187620400041&KeyUID=WOS:000187620400041}, booktitle={Materials, Technology and Reliability For Advanced Interconnects and Low-K Dielectrics-2003}, author={Lahlouh, B and Rajagopalan, T and Lubguban, JA and Biswas, N and Gangopadhyay, S and Sun, J and Huang, D and Simon, SL and Kim, HC and Volksen, W and et al.}, year={2003}, pages={291–296} } @inproceedings{lahlouh_rajagopalan_lubguban_biswas_gangopadhyay_sun_huang_simon_kim_volksen_et al._2003, title={Creating nanoporosity by selective extraction of porogens using supercritical carbon dioxide/cosolvent processes}, volume={766}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0348199041&partnerID=MN8TOARS}, booktitle={Materials Research Society Symposium - Proceedings}, author={Lahlouh, B. and Rajagopalan, T. and Lubguban, J.A. and Biswas, N. and Gangopadhyay, S. and Sun, J. and Huang, D. and Simon, S.L. and Kim, H.C. and Volksen, W. and et al.}, year={2003}, pages={291–296} } @inproceedings{bernazzani_simon_2003, title={Depression of Tg in polystyrene by freeze-drying}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0141728393&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Bernazzani, P. and Simon, S.L.}, year={2003}, pages={1987–1991} } @article{simon_bernazzani_mckenna_2003, title={Effects of freeze-drying on the glass temperature of cyclic polystyrenes}, volume={44}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000186949800029&KeyUID=WOS:000186949800029}, DOI={10.1016/j.polymer.2003.10.010}, number={26}, journal={Polymer}, author={Simon, S. L. and Bernazzani, P. and McKenna, G. B.}, year={2003}, pages={8025–8032} } @article{echeverria_kolek_plazek_simon_2003, title={Enthalpy recovery, creep and creep-recovery measurements during physical aging of amorphous selenium}, volume={324}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000184878800005&KeyUID=WOS:000184878800005}, DOI={10.1016/s0022-3093(03)}, number={3}, journal={Journal of Non-Crystalline Solids}, author={Echeverria, I. and Kolek, P. L. and Plazek, D. J. and Simon, S. L.}, year={2003}, pages={242–255} } @article{echeverría_kolek_plazek_simon_2003, title={Enthalpy recovery, creep and creep-recovery measurements during physical aging of amorphous selenium}, volume={324}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0041561174&partnerID=MN8TOARS}, DOI={10.1016/S0022-3093(03)00245-X}, number={3}, journal={Journal of Non-Crystalline Solids}, author={Echeverría, I. and Kolek, P.L. and Plazek, D.J. and Simon, S.L.}, year={2003}, pages={242–255} } @article{huang_simon_mckenna_2003, title={Equilibrium heat capacity of the glass-forming poly(alpha-methyl styrene) far below the Kauzmann temperature: The case of the missing glass transition}, volume={119}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000184474100002&KeyUID=WOS:000184474100002}, DOI={10.1063/1.1599271}, abstractNote={The absolute specific heat capacity of poly(α-methyl styrene) and mixtures with its pentamer were found to be independent of concentration at temperatures from 240 to 480 K. Extrapolation to 100% polymer yielded the equilibrium specific heat capacity for the polymer at temperatures as much as 180 K below the glass temperature or 130 K below the Kauzmann temperature. We find no evidence of a second order transition or a smeared transition in the equilibrium heat capacity, the entropy, the excess configurational entropy or the enthalpy over the entire range of temperatures investigated. The observations indicate that the Kauzmann paradox must be resolved without invoking a thermodynamic glass transition.}, number={7}, journal={Journal of Chemical Physics}, author={Huang, D. H. and Simon, S. L. and McKenna, G. B.}, year={2003}, pages={3590–3593} } @inproceedings{merzlyakov_meng_simon_mckenna_2003, title={Evaluation of different methods of measurement for the isotropic stress development in curing thermosets}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0141728409&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Merzlyakov, M. and Meng, Y. and Simon, S.L. and McKenna, G.B.}, year={2003}, pages={1898–1902} } @inbook{simon_2003, title={Physical Aging}, booktitle={Encyclopedia of Polymer Science and Technology}, publisher={John Wiley & Sons}, author={Simon, S.L.}, editor={Kroschwitz, JacquelineEditor}, year={2003} } @inproceedings{simon_wiesner_heinze_2003, title={Program improvements resulting from completion of one ABET 2000 assessment cycle}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-8744310747&partnerID=MN8TOARS}, booktitle={ASEE Annual Conference Proceedings}, author={Simon, S.L. and Wiesner, T.F. and Heinze, L.R.}, year={2003}, pages={9161–9166} } @article{rajagopalan_lahlouh_lubguban_biswas_gangopadhyay_sun_huang_simon_mallikarjunan_kim_et al._2003, title={Supercritical carbon dioxide extraction of porogens for the preparation of ultralow-dielectric-constant films}, volume={82}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0038110799&partnerID=MN8TOARS}, DOI={10.1063/1.1583139}, abstractNote={Supercritical carbon dioxide extraction of poly(propylene glycol) porogen from poly(methylsilsesquioxane) (PMSSQ) cured to temperatures adequate to initiate matrix condensation, but still below the decomposition temperature of the porogen, is demonstrated to produce nanoporous, ultralow-dielectric-constant thin films. Both closed and open cell porous structures were prepared simply by varying the porogen load in the organic/inorganic hybrid films. 25 and 55 wt % porogen loads were investigated in the present work. Structural characterization of the samples conducted using transmission electron microscope, small angle x-ray scattering, and Fourier transform infrared spectroscopy, confirms the extraction of the porogen from the PMSSQ matrix at relatively low temperatures (⩽200 °C). The standard thermal decomposition process is performed at much higher temperatures (typically in the range of 400 °C–450 °C). The values of dielectric constants and refractive indices measured are in good agreement with the structural properties of these samples.}, number={24}, journal={Applied Physics Letters}, author={Rajagopalan, T. and Lahlouh, B. and Lubguban, J.A. and Biswas, N. and Gangopadhyay, S. and Sun, J. and Huang, D.H. and Simon, S.L. and Mallikarjunan, A. and Kim, H.-C. and et al.}, year={2003}, pages={4328–4330} } @book{mckenna_simon_2002, title={Chapter 2 The glass transition: Its measurement and underlying physics}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-67649586633&partnerID=MN8TOARS}, DOI={10.1016/S1573-4374(02)80005-X}, journal={Handbook of Thermal Analysis and Calorimetry}, author={McKenna, G.B. and Simon, S.L.}, year={2002}, pages={49–109} } @article{bernatz_echeverria_simon_plazek_2002, title={Characterization of the molecular structure of amorphous selenium using recoverable creep compliance measurements}, volume={307}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000178038100106&KeyUID=WOS:000178038100106}, DOI={10.1016/s0022-3093(02)01522-3}, journal={Journal of Non-Crystalline Solids}, author={Bernatz, K. M. and Echeverria, I. and Simon, S. L. and Plazek, D. J.}, year={2002}, pages={790–801} } @article{bernazzani_simon_plazek_ngai_2002, title={Effects of entanglement concentration on T-g and local segmental motions}, volume={8}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000177546200012&KeyUID=WOS:000177546200012}, DOI={10.1140/epje/i2001-10087-5}, number={2}, journal={European Physical Journal E}, author={Bernazzani, P. and Simon, S. L. and Plazek, D. J. and Ngai, K. L.}, year={2002}, pages={201–207} } @article{simon_park_mckenna_2002, title={Enthalpy recovery of a glass-forming liquid constrained in a nanoporous matrix: Negative pressure effects}, volume={8}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000177546200013&KeyUID=WOS:000177546200013}, DOI={10.1140/epje/i2001-10072-0}, number={2}, journal={European Physical Journal E}, author={Simon, S. L. and Park, J. Y. and McKenna, G. B.}, year={2002}, pages={209–216} } @article{bernatz_echeverría_simon_plazek_2002, title={Erratum: Viscoelastic properties of amorphous boron trioxide (Journal of Non-Crystalline Solids (2001) 289 (9-16) PII: S0022309301007256)}, volume={297}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0036468543&partnerID=MN8TOARS}, DOI={10.1016/S0022-3093(01)00969-3}, number={2-3}, journal={Journal of Non-Crystalline Solids}, author={Bernatz, K.M. and Echeverría, I. and Simon, S.L. and Plazek, D.J.}, year={2002} } @article{lubguban_rajagopalan_mehta_lahlouh_simon_gangopadhyay_2002, title={Low-k organosilicate films prepared by tetravinyltetramethylcyclotetrasiloxane}, volume={92}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000176600000059&KeyUID=WOS:000176600000059}, DOI={10.1063/1.1483916}, abstractNote={Low-k films with k of 2.5–2.9 were deposited under different conditions of pressures and temperatures using a plasma-enhanced chemical vapor deposition (PECVD) system. These films were prepared using a new liquid precursor, tetravinyltetramethylcyclotetrasiloxane (TVTMCTS) and H2 carrier gas. The rf power was kept as low as possible to maintain the original ring structure in the films. The as-deposited films were annealed and the dielectric and optical properties were investigated. Identification of the absorption bands in the IR spectra for as-deposited films reveals a broadband around 950–1200 cm−1 arising from the Si–O stretching mode of the ring (1065 cm−1) and chain structure (1000 cm−1), respectively; a band at 750–900 cm−1 due to Si–O bending (790 cm−1); Si–CH3 rocking mode (760 cm−1); a sharp band centered at 1260 cm−1 due to a Si–CH3 bending mode; and a broadband at 2800–3000 cm−1 due to the CH group. A comparison of the IR spectra of the PECVD film and TVTMCTS liquid reveals that vinyl vibrations (Si–CH=CH2) at 960, 1410, and 3030–3095 cm−1 for CH2 and at 1598 cm−1 for C=C present in the liquid were not detected in the CVD films. Hence C=C bonds were broken in the plasma polymerization process. As the pressure and the deposition temperature (TD) increased, the intensity of the Si–O vibration arising from the ring structure increased and decreased, respectively. Thus by tuning the pressure and TD we can control the structure of the film. There is a good correlation found between the Si–CH3 and Si–O ring intensities and k values; the increasing Si–CH3 and Si–O ring is accompanied by decreasing k. The films were thermally stable up to 400 °C annealing temperature.}, number={2}, journal={Journal of Applied Physics}, author={Lubguban, J. and Rajagopalan, T. and Mehta, N. and Lahlouh, B. and Simon, S. L. and Gangopadhyay, S.}, year={2002}, pages={1033–1038} } @article{zheng_simon_mckenna_2002, title={Modeling structural recovery in glasses: An analysis of the peak-shift method}, volume={40}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000177633900007&KeyUID=WOS:000177633900007}, DOI={10.1002/polb.10270}, abstractNote={Abstract Amorphous polymers below their glass‐transition temperature are inherently not at equilibrium. As a result, their structures continuously relax in an attempt to reach the equilibrium state. The current models of structural recovery can quantitatively describe the process. One of the parameters needed for the models is the nonlinearity parameter x . It has been proposed that x can be obtained from experimental data with the so‐called peak‐shift method. In this work, we use the Tool–Narayanaswamy–Moynihan model to identify the factors that determine the accuracy of the peak‐shift method and to quantify the errors in the value of x obtained from the peak‐shift method. In addition, we determine the influence of the error in x on the evaluation of the nonexponential model parameter β. Finally, the peak‐shift method is compared with the traditional curve‐fitting method for model parameter determination. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2027–2036, 2002}, number={18}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Zheng, Y. and Simon, S. L. and McKenna, G. B.}, year={2002}, pages={2027–2036} } @article{lubguban_sun_rajagopalan_lahlouh_simon_gangopadhyay_2002, title={Supercritical carbon dioxide extraction to produce low-k plasma enhanced chemical vapor deposited dielectric films}, volume={81}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000179481900032&KeyUID=WOS:000179481900032}, DOI={10.1063/1.1525390}, abstractNote={A treatment to reduce the dielectric constant of plasma enhanced chemical vapor deposition (PECVD) films is presented. The method involved extracting low molecular weight or CO2 soluble species in the films by post deposition supercritical CO2 pressurization (SCCO2). We observed a decrease in k value of about 10%–14% in a composite film of organosilicate and a-C:F after SCCO2 treatment at 200 °C for 8 h. The composite films were deposited by PECVD using C4F8 and tetravinyltetramethylcyclotetrasiloxane (TVTMCTS) liquid source and H2 carrier gas at room temperature. As-deposited films were also annealed at 200 °C for 8 h in N2 atmosphere to compare the effect of thermal annealing without SCCO2 treatment. The result shows that there is no change in the k of the films after annealing. Thus, SCCO2 extraction is a good method for reducing the dielectric constant of these PECVD composite films. Supercritical CO2 pressurization of the film deposited using TVTMCTS and H2 only without the addition of C4F8 has no effect on the dielectric properties of the film while SCCO2 treatment of a-C:F samples deposited using C4F8 only dissolved the film. Therefore, in the composite film, we expect that CFx species dissolve during SCCO2 treatment while the organosilicate structure is preserved. Analysis of the Fourier-transform infrared spectra of the samples supports this hypothesis based on the decrease in the C–F absorption intensity after SCCO2 treatment.}, number={23}, journal={Applied Physics Letters}, author={Lubguban, J. A. and Sun, J. and Rajagopalan, T. and Lahlouh, B. and Simon, S. L. and Gangopadhyay, S.}, year={2002}, pages={4407–4409} } @article{bernazzani_simon_2002, title={Volume recovery of polystyrene: evolution of the characteristic relaxation time}, volume={307}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000178038100064&KeyUID=WOS:000178038100064}, DOI={10.1016/s0022-3093(02)01463-1}, journal={Journal of Non-Crystalline Solids}, author={Bernazzani, P. and Simon, S. L.}, year={2002}, pages={470–480} } @article{prasatya_mckenna_simon_2001, title={A viscoelastic model for predicting isotropic residual stresses in thermosetting materials: Effects of processing parameters}, volume={35}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000169096400002&KeyUID=WOS:000169096400002}, DOI={10.1177/002199801772662424}, number={10}, journal={Journal of Composite Materials}, author={Prasatya, P. and McKenna, G. B. and Simon, S. L.}, year={2001}, pages={826–848} } @article{prasatya_mckenna_simon_2001, title={A viscoelastic model for predicting isotropic residual stresses in thermosetting materials: Effects of processing parameters}, volume={35}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0034934918&partnerID=MN8TOARS}, DOI={10.1106/D0JJ-V6Q1-891M-3GJR}, abstractNote={The isotropic residual stresses in a composite subjected to three-dimensional constraints are calculated by extending a thermo-viscoelastic model developed previously by Simon et al. [1] to describe the time, temperature, and conversion dependence of the shear modulus for a commercial thermosetting material during cure. Experimental residual stress data as a function of cure time are fit to obtain limiting values for the rubbery and glassy bulk moduli. The effects of cure history on isotropic residual stresses are then investigated via simulations using the bulk moduli, a model of the cure kinetics, the relationship beween T g and conversion, and the stress relaxation function obtained in the thermo-viscoelastic model which includes the dependence of the shift factor on temperature and conversion. The isotropic residual stresses at room temperature can be directly related to the cure temperature at which gelation occurred for cases when vitrification does not occur during cure.}, number={10}, journal={Journal of Composite Materials}, author={Prasatya, P. and McKenna, G.B. and Simon, S.L.}, year={2001}, pages={826–848} } @book{simon_2001, title={Fitting differential scanning calorimetry heating curves for polyetherimide using a model of structural recovery}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0345447631&partnerID=MN8TOARS}, number={710}, journal={ACS Symposium Series}, author={Simon, S.L.}, year={2001}, pages={188–198} } @article{simon_2001, title={Temperature-modulated differential scanning calorimetry: theory and application}, volume={374}, ISSN={0040-6031}, url={http://dx.doi.org/10.1016/s0040-6031(01)00493-2}, DOI={10.1016/s0040-6031(01)00493-2}, number={1}, journal={Thermochimica Acta}, publisher={Elsevier BV}, author={Simon, Sindee L}, year={2001}, month={Jun}, pages={55–71} } @article{bernatz_echeverria_simon_plazek_2001, title={Viscoelastic properties of amorphous boron trioxide}, volume={289}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000170756600002&KeyUID=WOS:000170756600002}, DOI={10.1016/s0022-3093(01)00725-6}, number={1-3}, journal={Journal of Non-Crystalline Solids}, author={Bernatz, K. M. and Echeverria, I. and Simon, S. L. and Plazek, D. J.}, year={2001}, pages={9–16} } @article{simon_sobieski_plazek_2001, title={Volume and enthalpy recovery of polystyrene}, volume={42}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000165934200029&KeyUID=WOS:000165934200029}, DOI={10.1016/s0032-3861(00)00623-6}, number={6}, journal={Polymer}, author={Simon, S. L. and Sobieski, J. W. and Plazek, D. J.}, year={2001}, pages={2555–2567} } @article{simon_mckenna_sindt_2000, title={Modeling the evolution of the dynamic mechanical properties of a commercial epoxy during cure after gelation}, volume={76}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000085583600007&KeyUID=WOS:000085583600007}, DOI={10.1002/(sici)1097-4628(20000425)76:4<495::aid-app7>3.0.co;2-b}, abstractNote={The cure kinetics for a commercial epoxy have been established and the influence of the degree of cure on the glass transition determined. Time-temperature and time-conversion superposition principles have been built into a model that successfully predicts the development of the viscoelastic properties of the epoxy during isothermal cure from gelation to after vitrification. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 76: 495–508, 2000}, number={4}, journal={Journal of Applied Polymer Science}, author={Simon, S. L. and McKenna, G. B. and Sindt, O.}, year={2000}, pages={495–508} } @article{simon_mckenna_sindt_2000, title={Modeling the evolution of the dynamic mechanical properties of a commercial epoxy during cure after gelation}, volume={76}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0033902073&partnerID=MN8TOARS}, DOI={10.1002/(SICI)1097-4628(20000425)76:4<495::AID-APP7>3.0.CO;2-B}, number={4}, journal={Journal of Applied Polymer Science}, author={Simon, S.L. and Mckenna, G.B. and Sindt, O.}, year={2000}, pages={495–508} } @article{simon_mckenna_2000, title={Quantitative analysis of errors in TMDSC in the glass transition region}, volume={348}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000086576800010&KeyUID=WOS:000086576800010}, DOI={10.1016/s0040-6031(99)00482-7}, number={1-2}, journal={Thermochimica Acta}, author={Simon, S. L. and McKenna, G. B.}, year={2000}, pages={77–89} } @inbook{mckenna_simon_2000, title={Time dependent volume and enthalpy responses in polymers}, volume={1357}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000086082200002&KeyUID=WOS:000086082200002}, DOI={10.1520/stp15827s}, abstractNote={Most users of polymeric materials have a good sense that the glass transition event is kinetic in nature, i.e., it depends on cooling or heating rate in conventional experiments, and that the glassy state is a non-equilibrium state. However, it is often not appreciated that the structural recovery which occurs as a glass attempts to reach equilibrium is non-linear (e.g., the rate of volume recovery depends on the instantaneous volume). The non-linear viscoelastic nature of structural recovery can lead to surprising behaviors in certain kinds of measurements. It results in, for example, the asymmetry of approach in up and down-temperature jumps. The features of enthalpy and volume recovery, including the sub-Tg peaks and excess enthalpy overshoots in differential scanning calorimetry, are well-described by models of structural recovery developed in the 1970's. The purpose of the present work is to describe structural recovery and physical aging and their impacts on material performance and measurement of material properties. In addition, we present new results from calculations using the structural recovery models in which we demonstrate that new analytical tools, such as Temperature Modulated Differential Scanning Calorimetry (TMDSC), need to be used with caution when glass-forming systems are studied because of the nonlinear viscoelastic nature of structural recovery.}, booktitle={Time Dependent and Nonlinear Effects in Polymers and Composites}, author={McKenna, G. B. and Simon, S. L.}, editor={Schapery, R. A. and Sun, C. T.Editors}, year={2000}, pages={18–46} } @inproceedings{mckenna_simon_2000, title={Time dependent volume and enthalpy responses in polymers}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0033882869&partnerID=MN8TOARS}, number={1357}, booktitle={ASTM Special Technical Publication}, author={McKenna, G.B. and Simon, S.L.}, year={2000}, pages={18–46} } @article{karmana_simon_enick_1999, title={Carbon-dioxide-based microsortation of postconsumer polyolefins and its effect on polyolefin properties}, volume={38}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0032860242&partnerID=MN8TOARS}, number={3}, journal={Polymer-Plastics Technology and Engineering}, author={Karmana, E. and Simon, S. and Enick, R.}, year={1999}, pages={433–444} } @article{zacharia_simon_beckman_enick_1999, title={Improving the thermal stability of a polymer through liquid carbon dioxide extraction of a metal compound}, volume={63}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000077760500014&KeyUID=WOS:000077760500014}, DOI={10.1016/s0141-3910(98)00065-2}, number={1}, journal={Polymer Degradation and Stability}, author={Zacharia, R. E. and Simon, S. L. and Beckman, E. J. and Enick, R. M.}, year={1999}, pages={85–88} } @article{simon_mckenna_1999, title={Measurement of thermal conductivity using TMDSC: Solution to the heat flow problem}, volume={18}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0032641503&partnerID=MN8TOARS}, number={6}, journal={Journal of Reinforced Plastics and Composites}, author={Simon, S.L. and McKenna, G.B.}, year={1999}, pages={559–571} } @article{bernatz_giri_simon_plazek_1999, title={Physical aging by periodic creep and interrupted creep experiments}, volume={111}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000081698700052&KeyUID=WOS:000081698700052}, DOI={10.1063/1.479495}, abstractNote={A newly developed “interrupted creep” experiment has been used to study the physical aging of a low molecular weight polystyrene, Tg∼69 °C. The results of the new experiment are compared to those obtained from traditional “periodic creep” experiments. The interrupted creep experiment provides information about the viscosity, the recoverable creep compliance and the steady-state compliance, Js, during aging. Low molecular weight polystyrene was chosen because it exhibits a steady-state compliance that is a strong function of temperature. Aging was conducted at three temperatures, 68.2, 65.7 and 61.0 °C, using both down-jump and up-jump experiments. The behavior observed in the new experiments mirrors the behavior observed in the traditional experiments. In addition, the new experiments allow the first ever determination of how Js evolves during aging. The change of Js with aging time was calculated using the relationship between the shift factors, obtained from the recoverable creep compliance data, and the average relaxation times, obtained from the viscosity. The advantage of the new experiment is that it provides both the short-time recoverable creep compliance information and the long-time viscous flow. By combining these contributions to the creep compliance in a simple additive fashion, one can obtain a more complete picture of how the material is behaving during aging.}, number={5}, journal={Journal of Chemical Physics}, author={Bernatz, K. M. and Giri, L. and Simon, S. L. and Plazek, D. J.}, year={1999}, pages={2235–2241} } @inproceedings{sindt_simon_mckenna_liang_1998, title={Cure, shrinkage and properties of an epoxy material}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0031676841&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Sindt, Olivier and Simon, Sindee L. and McKenna, Gregory B. and Liang, Erwin}, year={1998}, pages={1658–1662} } @article{zacharia_simon_1998, title={Dynamic and isothermal thermogravimetric analysis of a polycyanurate thermosetting system}, volume={38}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000073552100003&KeyUID=WOS:000073552100003}, DOI={10.1002/pen.10219}, abstractNote={Abstract Isothermal and dynamic thermogravimetric analysis (TGA) were performed on a polycyanurate thermosetting material. The effects on thermal stability of the copper naphthenate catalyst were studied by comparing two systems with different amounts of catalyst. Tests were performed isothermally at temperatures ranging from 200°C to 240°C for times up to one week. Dynamic tests were also performed at rates from 0.1°C/min to 40°C/min at temperatures ranging from 180°C to 600°C. It is found that increasing the concentration of copper increased the weight loss in isothermal tests, but did not affect the overall activation energy. Conversely, in dynamic tests, copper concentration had no effect on weight loss. The results demonstrate the difficulties of extrapolating dynamic TGA data to use conditions in order to predict long‐term thermal stability.}, number={4}, journal={Polymer Engineering and Science}, author={Zacharia, R. E. and Simon, S. L.}, year={1998}, pages={566–572} } @inbook{simon_1998, title={Fitting DSC Heating Curves with the Tool-Narayanaswamy-Moynihan Model of Structural Recovery}, number={710}, booktitle={Structure and Properties of Glassy Polymers}, publisher={Oxford University Press, UK}, author={Simon, S.L.}, editor={Tant, M.R. and Hill, A.J.Editors}, year={1998} } @book{simon_1998, title={Fitting Differential Scanning Calorimetry Heating Curves for Polyetherimide Using a Model of Structural Recovery}, volume={710}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0041854210&partnerID=MN8TOARS}, journal={ACS Symposium Series}, author={Simon, S.L.}, year={1998}, pages={188–198} } @inproceedings{simon_mckenna_1998, title={Measurement of thermal conductivity using temperature-modulated differential scanning calorimetry: Solution to the heat flow problem}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0031702966&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, Sindee L. and McKenna, Gregory B.}, year={1998}, pages={2042–2046} } @inbook{echeverria_simon_1998, place={Trivandrum, India}, title={Structural Relaxation of Poly(Ether Imide): A Comparison of the Times to Reach Equilibrium for Different Properties" in "Recent Research Developments in Polymer Science}, volume={2}, booktitle={Recent Research Developments in Polymer Science}, author={Echeverria, I. and Simon, S.L.}, editor={Salamone, A.B. and Brandrup, J. and Ottenbrite, R.M. and Pandalai, D.G.Editors}, year={1998} } @inproceedings{wolf_simon_plazek_1998, title={Volume and enthalpy recovery of polystyrene}, volume={3}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0031649650&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Wolf, Elliot M. and Simon, Sindee L. and Plazek, Donald J.}, year={1998}, pages={3411–3414} } @article{sobieski_simon_plazek_1997, title={Dilatometric studies of physical aging of polyetherimide}, volume={49}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030654102&partnerID=MN8TOARS}, number={1}, journal={Journal of Thermal Analysis}, author={Sobieski, J.W. and Simon, S.L. and Plazek, D.J.}, year={1997}, pages={455–460} } @article{zacharia_simon_1997, title={Effect of catalyst on the thermal degradation of a polycyanurate thermosetting system}, volume={64}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0031552336&partnerID=MN8TOARS}, number={1}, journal={Journal of Applied Polymer Science}, author={Zacharia, R. E. and Simon, S. L.}, year={1997}, pages={127–131} } @inproceedings{simon_mckenna_1997, title={Effects of structural recovery in modulated DSC}, volume={2}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030719137&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, Sindee L. and McKenna, Gregory B.}, year={1997}, pages={2232–2237} } @article{simon_1997, title={Enthalpy recovery of poly(ether imide): Experiment and model calculations incorporating thermal gradients}, volume={30}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1997XK73400017&KeyUID=WOS:A1997XK73400017}, DOI={10.1021/ma9614508}, abstractNote={Enthalpy recovery data on poly(ether imide) from differential scanning calorimetry (DSC) is fit using the Moynihan−Tool−Narayanswamy model of structural recovery. A self-consistent phenomenological equation is used to describe the experimentally observed structure and temperature dependence of the relaxation time in both glass and equilibrium regimes. Temperature gradients in the DSC sample are calculated based on the thermal diffusivity of the material and are, for the first time, incorporated into the model calculations. When no thermal gradients are assumed, model parameters are found to vary with thermal history despite the use of the self-consistent equation for the relaxation time. There is also a discrepancy between experimental data and model calculations with respect to the shape of the DSC annealing peaks. Accounting for the presence of thermal gradients in the DSC sample is found to affect the values of the model parameters needed to fit the data. However, thermal gradients are unable to account for the thermal history dependence of the model parameters or for the discrepancy between observed and calculated shapes of DSC annealing peaks.}, number={14}, journal={Macromolecules}, author={Simon, S. L.}, year={1997}, pages={4056–4063} } @inproceedings{simon_1997, title={Enthalpy recovery of polyetherimide: experiment and model calculations}, volume={76}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030643363&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, Sindee L.}, year={1997}, pages={489–490} } @article{simon_mckenna_1997, title={Interpretation of the dynamic heat capacity observed in glass-forming liquids}, volume={107}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1997YG95800049&KeyUID=WOS:A1997YG95800049}, DOI={10.1063/1.475020}, abstractNote={Slow structural relaxations can complicate the interpretation of thermodynamic measurements on glass-forming liquids. Here we demonstrate using model calculations that structural recovery can lead to an apparent frequency-dependent heat capacity in ac calorimetry experiments. The model is shown to describe the complex heat capacity data reported in the literature for glycerol and poly(vinyl acetate). Importantly, the model does not invoke a complex heat capacity; rather, only static heat capacities are used. The analysis further suggests that ac calorimetry should provide a powerful way of testing models of structural recovery.}, number={20}, journal={Journal of Chemical Physics}, author={Simon, S. L. and McKenna, G. B.}, year={1997}, pages={8678–8685} } @inbook{simon_1997, title={Modeling DSC annealing peaks for polyetherimide: Incorporation of temperature gradients}, volume={455}, booktitle={Structure and Dynamics of Glasses and Glass Formers}, author={Simon, S. L.}, editor={Angell, C. A. and Ngai, K. L. and Kieffer, J. and Egami, T. and Nienhaus, G. U.Editors}, year={1997}, pages={177–182} } @inproceedings{simon_1997, title={Modeling DSC annealing peaks for polyetherimide: incorporation of temperature gradients}, volume={455}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030659915&partnerID=MN8TOARS}, booktitle={Materials Research Society Symposium - Proceedings}, author={Simon, Sindee L.}, year={1997}, pages={177–182} } @article{simon_plazek_sobieski_mcgregor_1997, title={Physical aging of a polyetherimide: Volume recovery and its comparison to creep and enthalpy measurements}, volume={35}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1997WQ33000007&KeyUID=WOS:A1997WQ33000007}, DOI={10.1002/(sici)1099-0488(19970430)35:6<929::aid-polb7>3.0.co;2-c}, abstractNote={Journal of Polymer Science Part B: Polymer PhysicsVolume 35, Issue 6 p. 929-936 Article Physical aging of a polyetherimide: Volume recovery and its comparison to creep and enthalpy measurements Sindee L. Simon, Corresponding Author Sindee L. Simon Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorDonald J. Plazek, Donald J. Plazek Materials Science and Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorJ. William Sobieski, J. William Sobieski Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorEric T. McGregor, Eric T. McGregor Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this author Sindee L. Simon, Corresponding Author Sindee L. Simon Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorDonald J. Plazek, Donald J. Plazek Materials Science and Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorJ. William Sobieski, J. William Sobieski Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this authorEric T. McGregor, Eric T. McGregor Chemical Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261Search for more papers by this author First published: 07 December 1998 https://doi.org/10.1002/(SICI)1099-0488(19970430)35:6<929::AID-POLB7>3.0.CO;2-CCitations: 93AboutPDF ToolsRequest permissionExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL Share a linkShare onEmailFacebookTwitterLinkedInRedditWechat Abstract Volume recovery measurements have been used to study the physical aging behavior of a polyetherimide. Isothermal aging temperatures near Tg were studied with aging times ranging up to several days. The volume decreases during physical aging and levels off at equilibrium. For comparison purposes, the data are normalized to yield the departure from equilibrium which varies from unity at very short aging times to zero when equilibrium is reached. As the aging temperature decreases, the normalized curves are shifted to longer times without a significant change in shape. Hence, the data can be reduced by aging time—temperature superposition. The temperature dependence of the shift factors used to reduce the volume recovery data and the times to reach equilibrium for the volume recovery follow the WLF equation and agree within experimental error with the values from enthalpy and creep measurements obtained in previous work. However, the approach to equilibrium for volume appears to differ from that of enthalpy, with volume recovery being faster than the enthalpy recovery at short times. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35: 929–936, 1997 References and Notes 1 L. C. E. Struik, Physical Aging in Amorphous Polymers and Other Materials, Elsevier, New York, 1978. 2 G. B. McKenna, in Comprehensive Polymer Science, Vol. 2, Polymer Properties, C. Booth and C. Price, Eds., Pergamon, Oxford, 1989, Chap. 2. 3 J. O'Reilly, CRC Crit. Rev. Solid State Mater. Sci., 13 (3), 259 (1987). 4 I. Hodge, J. Non-Cryst. Solids, 169, 211 (1994). 5 G. B. McKenna, et al., Polym. Eng. Sci., 35(5), 403 (1995). G. B. McKenna, C. R. Schultheisz, and Y. Leterrier, in Deformation Yield and Fracture of Polymers, Proc. 9th Int'l. Conf., Cambridge, UK, 1994, pp. 31/1–31/4. 6 C. A. Bero and D. J. Plazek, J. Polym. Sci. B: Polym. Phys., 29, 39 (1991). 7 E. F. Oleinik, Polym. J., 19(1), 105 (1987). 8 K. Adachi and T. Kotaka, Polym. J., 14, 959 (1982). 9 R. J. Roe and G. M. Millman, Polym. Eng. Sci., 23, 318 (1983). 10 A. Weitz and B. Wunderlich, J. Polym. Sci.: Polym. Phys., 12, 2473 (1974). 11 J. Perez, J. Y. Cavaille, R. D. Calleja, J. L. G. Ribelles, M. M. Pradas, and A. R. Greus, Makromol. Chem., 192, 2141 (1991). 12 J. Mijovic and T. Ho, Polymer, 34, 3865 (1993). 13 H. Sasabe and C. T. Moynihan, J. Polym. Sci. B: Polym. Phys., 16, 1447 (1978). 14 J. M. G. Cowie, S. Elliott, R. Ferguson, and R. Simha, Polym. Commun., 28, 298 (1987). 15 G. W. Scherer, Relaxation in Glass and Composites, Wiley, New York, 1986. 16 I. Echeverria, P-C. Su, S. L. Simon, and D. J. Plazek, J. Polym. Sci. B: Polym. Phys., 33, 2457 (1995). 17 N. Wachenhut, Air Force Office of Scientific Research Contract F33615-77-C5235, Technical Report, 1981. 18 P. Zoller, P. Bolli, V. Pahud, and H. Ackerman, Rev. Sci. Instrum., 47(8), 948 (1976). 19 R. N. Haward, J. Macromol. Sci. Rev. Macromol. Chem., C4(2), 191 (1970). 20 E. A. DiMarzio and F. Dowell, J. Appl. Phys., 50, 6061 (1979). 21 Unpublished data of B. Lander and D. J. Plazek, University of Pittsburgh. 22 D. J. Plazek and K. L. Ngai, Macromolecules, 24, 1222 (1991). 23 R. Bohmer and C. A. Angell, Phys. Rev. B, 45(17), 10091 (1992). 24 P.-C. Su, Master's thesis, Chemical Engineering Department, University of Pittsburgh, 1994. 25 C. A. Angell, J. Non-Cryst. Solids, 131–133, 13 (1991). Citing Literature Volume35, Issue630 April 1997Pages 929-936 ReferencesRelatedInformation}, number={6}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Simon, S. L. and Plazek, D. J. and Sobieski, J. W. and McGregor, E. T.}, year={1997}, pages={929–936} } @inproceedings{simon_plazek_harper_holden_1997, title={Physical aging of polystyrene: volume and enthalpy recovery}, volume={76}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030646244&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, Sindee L. and Plazek, Donald J. and Harper, Brant J. and Holden, Thomas}, year={1997}, pages={334–335} } @article{simon_mckenna_1997, title={The effects of structural recovery and thermal lag in temperature-modulated DSC measurements}, volume={307}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:000071690100001&KeyUID=WOS:000071690100001}, DOI={10.1016/s0040-6031(97)00275-x}, number={1}, journal={Thermochimica Acta}, author={Simon, S. L. and McKenna, G. B.}, year={1997}, pages={1–10} } @article{zacharia_simon_1997, title={Thermogravimetric analysis of a polycyanurate thermosetting material}, volume={49}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1997XD48000038&KeyUID=WOS:A1997XD48000038}, DOI={10.1007/bf01987452}, abstractNote={The thermal degradation of a fully cured polycyanurate thermosetting material was examined by monitoring the mass loss at various temperatures ranging from 200 to 220‡C. The effect of the copper naphthenate catalyst, generally used to facilitate curing, is also studied. A decrease in weight is observed with increasing time at elevated temperatures in the systems containing copper naphthenate, with the onset of degradation occurring sooner with higher concentrations of the copper compound. The apparent activation energy for degradation is 220±30 kJ mol−1.}, number={1}, journal={Journal of Thermal Analysis}, author={Zacharia, R. E. and Simon, S. L.}, year={1997}, pages={311–315} } @inproceedings{simon_plazek_1996, title={Physical aging of polyetherimide: enthalpy, volume, and creep measurements}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030360547&partnerID=MN8TOARS}, booktitle={Technical Papers, Regional Technical Conference - Society of Plastics Engineers}, author={Simon, S.L. and Plazek, D.J.}, year={1996} } @inproceedings{simon_1996, title={Thermal stability of a polycyanurate thermosetting system: effects of cure catalysts}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0030364342&partnerID=MN8TOARS}, booktitle={Technical Papers, Regional Technical Conference - Society of Plastics Engineers}, author={Simon, S.L.}, year={1996} } @article{echeverria_su_simon_plazek_1995, title={PHYSICAL AGING OF A POLYETHERIMIDE - CREEP AND DSC MEASUREMENTS}, volume={33}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1995TG02600017&KeyUID=WOS:A1995TG02600017}, DOI={10.1002/polb.1995.090331717}, abstractNote={Abstract Creep and differential scanning calorimetry (DSC) measurements have been used to study the physical aging behavior of a polyetherimide. Isothermal aging temperatures ranged from 160°C to T g with aging times ranging from 10 min to 8 days. The only measurable effect of physical aging on the short‐time creep curves is a shift of the creep compliance to longer times. Andrade plots of the compliance versus the cube root of time are linear at short times with the slope β decreasing with increasing aging time to a constant value once equilibrium is reached. Log β 3 is related directly to the degree to which the creep curves shift to longer times with physical aging, and is used in this work as a measure of physical aging. A reduced curve of log β 3 versus log aging time is obtained for the aging temperatures investigated by appropriate vertical and horizontal shifts. The enthalpy change during aging increases linearly with the logarithm of the aging time, t a , leveling off at equilibrium at values which increase with decreasing aging temperature. Hence, both nonequilibrium and equilibrium temperature shift factors can be calculated from the DSC data. Good agreement is observed between the equilibrium temperature shift factors obtained from the creep and DSC data. The temperature dependence of the nonequilibrium temperature shift factors is found to be an order of magnitude smaller than that of the equilibrium shift factors. The time scales to reach equilibrium for enthalpy and for mechanical measurements are found to be the same within experimental error. © 1995 John Wiley & Sons, Inc.}, number={17}, journal={Journal of Polymer Science Part B-Polymer Physics}, author={Echeverria, I. and Su, P. C. and Simon, S. L. and Plazek, D. J.}, year={1995}, pages={2457–2468} } @inproceedings{simon_gillham_1994, title={Application of thermosetting cure diagrams}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0027989115&partnerID=MN8TOARS}, number={pt 2}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, S.L. and Gillham, J.K.}, year={1994}, pages={2192–2195} } @article{simon_gillham_1994, title={CONVERSION TEMPERATURE PROPERTY DIAGRAM FOR A LIQUID DICYANATE ESTER HIGH-TG POLYCYANURATE THERMOSETTING SYSTEM}, volume={51}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1994MX48200007&KeyUID=WOS:A1994MX48200007}, DOI={10.1002/app.1994.070511007}, abstractNote={Abstract The effects of conversion and temperature on the dynamic mechanical properties (at ≈ 1 Hz) of a dicyanate ester/polycyanurate thermosetting system are investigated after cure using torsional braid analysis (TBA). Extent of conversion is measured by T g . The isothermal glassy‐state modulus at measurement temperatures below the glass transition temperature of the monomer ( T g 0 ) decreases with increasing conversion. The isothermal modulus at temperatures above T g 0 passes through a maximum due to competition between increase in the isothermal glassy‐state modulus at the measurement temperature due to the vitrification process during cooling and the aforementioned decrease in the modulus with increasing conversion, which is considered to be due primarily to steric constraints in the developing network. The maximum in the isothermal modulus is associated with the boundary between the glass and glass transition regions. The experimental results are summarized in a conversion ( T g )–temperature–property diagram, the T g TP diagram, which is a framework for understanding relationships between transitions and material properties for thermosetting systems. © 1994 John Wiley & Sons, Inc.}, number={10}, journal={Journal of Applied Polymer Science}, author={Simon, S. L. and Gillham, J. K.}, year={1994}, pages={1741–1752} } @misc{simon_gillham_1994, title={Cyanate ester/polycyanurate systems: structure-property relationships}, ISBN={9789401045773 9789401113267}, url={http://dx.doi.org/10.1007/978-94-011-1326-7_4}, DOI={10.1007/978-94-011-1326-7_4}, journal={Chemistry and Technology of Cyanate Ester Resins}, publisher={Springer Netherlands}, author={Simon, S. L. and Gillham, J. K.}, year={1994}, pages={87–111} } @article{simon_gillham_1994, title={THERMOSETTING CURE DIAGRAMS - CALCULATION AND APPLICATION}, volume={53}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1994PB11500001&KeyUID=WOS:A1994PB11500001}, DOI={10.1002/app.1994.070530601}, abstractNote={Abstract The time–temperature–transformation (TTT) isothermal cure diagram and the continuous‐heating‐transformation (CHT) cure diagram are calculated from a reaction model for a high‐ T g epoxy/amine system that has been developed to describe both epoxy/amine and etherification reactions in kinetically and diffusion‐controlled reaction regimes. The cure diagrams are applied to various processing operations. The optimization of processing and of material properties by exploiting gelation and/or vitrification during cure is discussed. © 1994 John Wiley & Sons, Inc.}, number={6}, journal={Journal of Applied Polymer Science}, author={Simon, S. L. and Gillham, J. K.}, year={1994}, pages={709–727} } @article{simon_gillham_1993, title={COATINGS APPLICATIONS OF THERMOSETTING CURE DIAGRAMS - OFF-STOICHIOMETRIC HIGH-TG EPOXY-AMINE SYSTEMS}, volume={65}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0027648874&partnerID=MN8TOARS}, number={823}, journal={Journal of Coatings Technology}, author={Simon, S. L. and Gillham, J. K.}, year={1993}, pages={57–65} } @article{simon_gillham_1993, title={CURE KINETICS OF A THERMOSETTING LIQUID DICYANATE ESTER MONOMER HIGH-TG POLYCYANURATE MATERIAL}, volume={47}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1993KE25200008&KeyUID=WOS:A1993KE25200008}, DOI={10.1002/app.1993.070470308}, abstractNote={Abstract The cure of a liquid dicyanate ester monomer, which reacts to form a high‐ T g (≈200°C) polycyanurate network, has been investigated using differential scanning calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR), and a dynamic mechanical technique, torsional braid analysis (TBA). The monomer is cured with and without catalyst. The same one‐to‐one relationship between fractional conversion and the dimensionless glass transition temperature is found from DSC data for both the uncatalyzed and catalyzed systems, independent of cure temperature, signifying that the same polymeric structure is produced. T g is the parameter used to monitor the curing reactions since it is uniquely related to conversion, is sensitive, is accurately determined, and is also directly related to the solidification process. The rate of uncatalyzed reaction is found to be much slower than that of the catalyzed reaction. The apparent overall activation energy for the uncatalyzed reaction is found to be greater than that of the catalyzed reaction (22 and 13 kcal/mol, respectively) from time–temperature superposition of experimental isothermal T g vs. In time data to form kinetically‐controlled master curves for the two systems. Although the time–temperature superposition analysis does not necessitate knowledge of the rate expression, it has limitations, because if the curing process consists of parallel reactions with different activation energies, as is considered to be the case from analysis of the FTIR data, there should not be a kinetically‐controlled master curve. Consequently, a kinetic model, which can be satisfactorily extrapolated, is developed from FTIR isothermal cure studies of the uncatalyzed reaction. The FTIR data for the uncatalyzed system at high cure temperatures, where the material is in the liquid or rubbery states throughout cure, 190 to 220°C, are fitted by a model of two parallel reactions, which are second‐order and second‐order autocatalytic (with activation energies of 11 and 29 kcal/mol), respectively. Using the model parameters determined from the FTIR studies and the relationship between T g and conversion from DSC studies, T g , vs. time curves are calculated for the uncatalyzed system and found to agree with DSC experimental results for isothermal cure temperatures from 120 to 200°C to even beyond vitrification. The DSC data for the catalyzed system are also described by the same kinetic model after incorporating changes in the pre‐exponential frequency factors (due to the higher concentration of catalyst) and after incorporating diffusion‐control, which occurs prior to vitrification in the catalyzed system (but well after vitrification in the uncatalyzed system). Time–temperature‐transformation (TTT) isothermal cure diagrams for both systems are calculated from the kinetic model and compared to experimental TBA data. Experimental gelation is found to occur at a conversion of approximately 64% in the catalyzed system by comparison of experimental macroscopic gelation at the various curing temperatures and iso‐ T g (iso‐conversion) curves calculated from the kinetic model. © 1993 John Wiley & Sons, Inc.}, number={3}, journal={Journal of Applied Polymer Science}, author={Simon, S. L. and Gillham, J. K.}, year={1993}, pages={461–485} } @inproceedings{simon_gillham_1992, title={Cure and TTT diagram for a high-Tg epoxy-/amine system: Application to powder coatings}, volume={67}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0027004447&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K.}, year={1992}, pages={342–343} } @inproceedings{simon_gillham_1992, title={Cure of a dicyanate ester/polycyanurate system}, volume={66}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0026973293&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K.}, year={1992}, pages={453–454} } @inproceedings{simon_gillham_1992, title={Physical aging of a dicyanate ester/polycyanurate material}, volume={66}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0026980695&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K.}, year={1992}, pages={277–278} } @article{simon_gillham_1992, title={REACTION-KINETICS AND TTT CURE DIAGRAMS FOR OFF-STOICHIOMETRIC RATIOS OF A HIGH-TG EPOXY/AMINE SYSTEM}, volume={46}, url={http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcAuth=ORCID&SrcApp=OrcidOrg&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=WOS:A1992JT35700014&KeyUID=WOS:A1992JT35700014}, DOI={10.1002/app.1992.070460714}, abstractNote={Abstract The cure behavior of two off‐stoichiometric ratios, one amine‐rich and the other epoxy‐rich, of a high‐ T g aromatic difunctional epoxy/aromatic tetrafunctional diamine system are studied using differential scanning calorimetry (DSC). The curing reactions are presumed to consist of competing epoxy/amine reactions and a subsequent reaction of epoxy with hydroxyl, termed etherification, that is significant in the epoxy‐rich system after depletion of amino hydrogens. A one‐to‐one relationship between conversion and T g independent of cure temperature is found for each of the off‐stoichiometric systems in spite of the competing reactions. Since it is uniquely related to fractional conversion, sensitive, and easily measured, T g is used to monitor the curing reactions. Kinetically controlled master curves for isothermal cure are obtained for each system by shifting T g vs. In time data to an arbitrary reference temperature; apparent activation energies for the epoxy/amine and the etherification reaction regimes appear to be identical at 15.5 kcal/mol. Experimental DSC data are satisfactorily described for both systems by a kinetic model of the competitive epoxy/amine and etherification reactions. The ratio of the rate constants for the reactions of epoxy with the secondary amine group and epoxy with the primary amine group, α = k 2 / k 1 , is found to be approximately 0.5 (i.e., equal reactivity of amino hydrogens), whereas the ratio of the rate constants for the reactions of epoxy with hydroxyl and epoxy with the primary amine group, β = k 3 / k 1 , is found to be 0.001. Diffusion control is incorporated in the model by the use of a free‐volume theory. Vitrification and iso‐ T g contours in the time–temperature–transformation (TTT) isothermal‐cure diagram are calculated for both systems from the kinetic model. © 1992 John Wiley & Sons, Inc.}, number={7}, journal={Journal of Applied Polymer Science}, author={Simon, S. L. and Gillham, J. K.}, year={1992}, pages={1245–1270} } @inproceedings{simon_gillham_1992, title={TgTP diagram for a dicyanate ester/polycyanurate system}, volume={66}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0026981611&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K.}, year={1992}, pages={502–503} } @inproceedings{simon_gillham_1991, title={Cure kinetics of a liquid high-Tg dicyanate ester/polycyanurate system}, volume={37}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0025748929&partnerID=MN8TOARS}, booktitle={Annual Technical Conference - ANTEC, Conference Proceedings}, author={Simon, S.L. and Gillham, J.K.}, year={1991}, pages={1554–1558} } @inproceedings{simon_gillham_shimp_1990, title={Cure of a high-Tg liquid thermosetting dicyanate ester/polycyanurate system}, volume={62}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0025254997&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K. and Shimp, D.A.}, year={1990}, pages={96–100} } @inproceedings{simon_gillham_1990, title={Effect of conversion and temperature on the physical annealing of a high-Tg dicyanate ester/polycyanurate material}, volume={63}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0025637443&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Gillham, J.K.}, year={1990}, pages={760–766} } @inproceedings{simon_wisanrakkit_gillham_1989, title={Reaction kinetics and calculation of iso-T$/g$/ and vitrification curves in the TTT diagram for a high Tg epoxy/amine system}, volume={61}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0024890883&partnerID=MN8TOARS}, booktitle={Polymeric Materials Science and Engineering, Proceedings of the ACS Division of Polymeric Materials Science and Engineering}, author={Simon, S.L. and Wisanrakkit, G. and Gillham, J.K.}, year={1989}, pages={799–805} }