Cationic Cure of Epoxy Resin by an Optimum Concentration of N-benzylpyrazinium Hexafluoroantimonate

Similar documents
The Kinetics of B-a and P-a Type Copolybenzoxazine via the Ring Opening Process

In Situ Detection of the Onset of Phase Separation and Gelation in Epoxy/Anhydride/Thermoplastic Blends

Dynamic Mechanical Analysis of Thermosetting Materials

Effect of Electron Beam and γ-ray Irradiation on the Curing of Epoxy Resin

Study on Curing Kinetics and Curing Mechanism of Epoxy Resin Based on Diglycidyl Ether of Bisphenol A and Melamine Phosphate

Kinetics Study of Imidazole-Cured Epoxy-Phenol Resins

An investigation was conducted based on: (1) The study of cure mechanisms and

Modeling and Simulation to Materials Development for Rapid Prototyping Process

APPLICATIONS OF THERMAL ANALYSIS IN POLYMER AND COMPOSITES CHARACTERIZATION. Wei Xie TA Instruments

Modeling of Ultrasonic Processing

Thermal Stability of Trifunctional Epoxy Resins Modified with Nanosized Calcium Carbonate

Polymerization Mechanisms and Curing Kinetics of Novel Polymercaptan Curing System Containing Epoxy/Nitrogen

Rheological behavior during the phase separation of thermoset epoxy/thermoplastic polymer blends

DEVELOPMENT OF IMPROVED METHODS FOR CHARACTERISING THE CURE OF COMPOSITE MATERIALS

Kinetics and Curing Mechanism of Epoxy and Boron Trifluoride Monoethyl Amine Complex System

CHAPTER-5 BISMALEIMIDE-ALLYL NOVOLAC OLIGOMERS: SYNTHESIS AND CURE KINETICS

Synthesis and Characterization of a Novel Silicon-Containing Epoxy Resin

The Effect of Cure Rate on Glassy Polymer Networks

Measuring the Reaction Kinetics of Polyester/PCA-501(HAA) Powder Coatings With Dielectric Analysis (DEA)

Modelling of viscoelastic properties of a curing adhesive

MODELING OF ULTRASONIC PROCESSING

DAMPING PROPERTIES OF SOME EPOXY RESIN - LIQUID CRYSTALLINE DYESTUFF COMPOSITES

CURE DEPENDENT CREEP COMPLIANCE OF AN EPOXY RESIN

Determination of Activation Energy for Glass Transition of an Epoxy Adhesive Using Dynamic Mechanical Analysis

Thermal and mechanical properties of several phthalonitrile resin system

Modulated DSC Paper #8 Use Of Quasi-isothermal Mode for Improved Understanding of Structure Change

THERMOMECHANICAL AND RHEOLOGICAL PROPERTIES OF AN EPOXY RESIN FILLED WITH EPOXY MICROSPHERES

EXPERIMENTALLY DETERMINING THE VISCOELASTIC BEHAVIOR OF A CURING THERMOSET EPOXY R. Thorpe 1, A. Poursartip 1*

Synergy of the combined application of thermal analysis and rheology in monitoring and characterizing changing processes in materials

Differential Scanning Calorimetry study on curing kinetics of diglycidyl ether of bisphenol A with amine curing agents for self-healing systems

TOPEM the new advanced multi-frequency TMDSC technique

Curing Properties of Cycloaliphatic Epoxy Derivatives

Supporting Information for

A FINITE ELEMENT COUPLING MODEL FOR INTERNAL STRESS PREDICTION DURING THE CURING OF THICK EPOXY COMPOSITES

DSC Methods to Quantify Physical Aging and Mobility in Amorphous Systems: Assessing Molecular Mobility

Supporting Information

Cure Kinetics of Aqueous Phenol Formaldehyde Resins Used for Oriented Strandboard Manufacturing: Effect of Zinc Borate

THE MATRIX: EVOLUTIONS II

D1-204 PROPERTIES OF EPOXY-LAYERED SILICATE NANOCOMPOSITES T. SHIMIZU*, T. OZAKI, Y. HIRANO, T. IMAI, T. YOSHIMITSU TOSHIBA CORPORATION.

POLYBENZOXAZINE BASED NANOCOMPOSITES REINFORCED WITH MODIFIED GRAPHENE OXIDE

Effect of Resin Molecular Architecture on Epoxy Thermoset Mechanical Properties

Cure Kinetics of Aqueous Phenol Formaldehyde Resins Used for Oriented Strandboard Manufacturing: Analytical Technique

Thermal degradation kinetics of Arylamine-based Polybenzoxazines

THEORY AND APPLICATIONS OF MODULATED TEMPERATURE PROGRAMMING TO THERMOMECHANICAL TECHNIQUES

MICROWAVE EFFECTS ON CFRP PROCESSING SUMMARY 1. MICROWAVE BACKGROUND

Mechanical and Electro-optic Properties of Thiol-ene based Polymer Dispersed Liquid Crystal (PDLC)

Common Definition of Thermal Analysis

INFLUENCE OF CARBON NANOFIBERS AND PIEZOELECTRIC PARTICLES ON THE THERMOMECHANICAL BEHAVIOR OF EPOXY MIXTURES

Experimental Study of the Induced Residual Stresses During the Manufacturing Process of an Aeronautic Composite Material

Supporting Information

Polymer Systems and Film Formation Mechanisms in High Solids, Powder, and UV Cure Systems

A Rubber-Modified Thermoplastic where the Morphology Produced by Phase-Separation Induced by Polymerization Disappears at High Conversions

Please do not adjust margins. Electronic Supplementary Material (ESI) for Journal of Materials

Phase Changes and Latent Heat

Dual Cure Phenol - Epoxy Resins: Characterisation and Properties

DSC AS PROBLEM-SOLVING TOOL: BETTER INTERPRETATION OF Tg USING CYCLIC DSC

Phase separation and rheological behavior in thermoplastic modified epoxy systems

Nutshells of Thermal Analysis. Heat it up! Burn it! Thermal Analysis

Rheology, Adhesion, and Debonding of Lightly Cross-linked Polymer Gels

Effect Of Curing Method On Physical And Mechanical Properties Of Araldite DLS 772 / 4 4 DDS Epoxy System

Incorporation of Reaction Chemicals Testing Data in Reactivity Hazard Evaluation. Ken First Dow Chemical Company Midland, MI

Modified Time-temperature Superposition Principle for Viscoelasticity of Thermosetting Resins

Thermally Functional Liquid Crystal Networks by Magnetic Field Driven Molecular Orientation

Thermal and UV-curing Behavior of Inks, Adhesives, and Coatings by Photo-, In-situ DEA and DMA.

Cure kinetics characterization and monitoring of an epoxy resin for thick composite structures

Photopolymerization of Thermoplastic Polyurethane/Acrylate Blends

Bismaleimide Modified Epoxy-Diallylbisphenol System Effect of Bismaleimide Nature on Properties

Effect of Curing Time and Temperature on the Structural Stability of Melamine Formaldehyde Polymers Phisan Katekomol

Lecture 15:The Tool-Narayanaswamy-Moynihan Equation Part II and DSC

A STUDY OF CLAY-EPOXY NANOCOMPOSITES CONSISTING OF UNMODIFIED CLAY AND ORGANO CLAY

Melting and solidi cation of Pb nanoparticles embedded in an Al matrix as studied by temperature-modulated di erential scanning calorimetry

STUDIES ON APPARENT KINETICS AND RHEOLOGICAL BEHAVIOR OF EPOXY/ACRYLATE IPNS AS VACUUM PRESSURE IMPREGNATION RESINS *

Collision Theory. and I 2

QUANTITATIVE EVALUATION OF CURING SHRINKAGE IN POLYMERIC MATRIX COMPOSITES

Cure Kinetics of Ring-Opening Metathesis Polymerization of Dicyclopentadiene *

Thermal Methods of Analysis Theory, General Techniques and Applications. Prof. Tarek A. Fayed

Differential scanning calorimetry of phenol formaldehyde resins cure-accelerated by carbonates

Effect of curing time on phenolic resins using latent acid catalyst

CURING KINETIC AND PROPERTIES OF MeHHPA /HYDANTOIN EPOXY RESIN SYSTEM

Influence of the thermodynamic state of bisphenol A and aliphatic epoxy oligomers on the temperature dependences of Newtonian viscosity

Applications of the Micro Reaction Calorimeter

Cure Kinetics of the Ring-Opening Metathesis Polymerization of Dicyclopentadiene

The effect of phase transition of methanol on the reaction rate in the alkylation of hydroquinone

POLARIZATION STABILITY OF AMORPHOUS PIEZOELECTRIC POLYIMIDES

Mechanical and Thermoviscoelastic Behavior of Clay/Epoxy Nanocomposites

(Refer Slide Time: 00:58)

A Technical Whitepaper Polymer Technology in the Coating Industry. By Donald J. Keehan Advanced Polymer Coatings Avon, Ohio, USA

CHEM*3440. Thermal Methods. Thermogravimetry. Instrumental Components. Chemical Instrumentation. Thermal Analysis. Topic 14

Application of Kissinger analysis to glass transition and study of thermal degradation kinetics of phenolic acrylic IPNs

Physical aging of thermosetting powder coatings

The Study on Poly(ether sulfone) Modified Cyanate Ester Resin and Epoxy Resin Cocuring Blends

TECHNICAL UPDATE. Ricon Resins Peroxide Curing Data and Use as a Reactive Plasticizer in Polyphenylene Ether Based CCL and PWB

Physical Aging of Epoxy Polymers and Their Composites

Chapter 4. Results and Discussion. 4.1 Monomer Syntheses

PREPARATION AND CHARACTERIZATION OF ATBN- FUNCTIONALIZED GRAPHENE NANOPLATELETS AND THE EPOXY NANOCOMPOSITES

Raman spectroscopy of epoxy resin crosslinking

Supplementary Figures:

Factors Affecting Reaction Rate

Thermodynamic Processes and Thermochemistry

CHAPTER 4 TOUGHENING/MODIFICATION OF BISMALEIMIDES WITH ALLYL COMPOUNDS PART 1: BISMALEIMIDE-ALLYL PHENOL RESIN SYSTEM (BMIP/DABA)

Transcription:

Macromolecular Research, Vol. 10, No. 1, pp 34-39 (2002) Cationic Cure of Epoxy Resin by an Optimum Concentration of N-benzylpyrazinium Hexafluoroantimonate Jong Keun Lee* and Yusong Choi Department of Polymer Science and Engineering, Kumoh National University of Technology, 188 Shinpyung Dong, Kumi, Kyungbuk 730-701, Korea Jae-Rock Lee Advanced Materials Division, Korea Research Institute of Chemical Technology, P.O. Box 107, Yusong, Taejon 305-600, Korea Jaekyeung Park Department of Chemical Engineering, Sangju National University, 386 Gajang Dong, Sangju, Kyungbuk 742-711, Korea Received Jan. 5, 2002 ; Revised Feb. 2, 2002 Abstract: Cure behavior of an epoxy resin was investigated at different cure temperatures (110, 120, 130, 140, and 150 o C) and cure times in the presence of 2 wt% of an N-benzylpyrazinium hexafluoroantimonate (BPH) cationic catalyst by means of differential scanning calorimetry (DSC) and dynamic mechanical analysis (DMA). The glass transition temperature (T g ) and chemical conversion (x) at the different temperatures were determined from DSC thermograms. The T g and x vs. ln time data were superposed up to T g =100 o C and x = 0.70 by shifting horizontally at a reference temperature of T g = 130 o C. It is interesting that the T g and x of the superposed data increase rather slowly in the early stage of cure and rapidly thereafter. Therefore, the increase of the T g and x can be divided into two regions; R I = -18.4(= T go )~5 o C and R II = 5 ~ 100 o C in T g, and R I = 0 ~ 0.24 and R II = 0.24 ~ 0.70 in x. The R I is closely related to the initiation reactions between BPH and epoxy and between hydroxy group and epoxy in this epoxy/catalyst system. From the kinetic analysis of the T g -shift, activation energy was 12.5 kcal/mol. The relationship between T g and x was also considered. The gelation and vitrification times for different cure temperatures were obtained from DMA curves. Keywords : epoxy, cationic catalyst, glass transition temperature, conversion, cure kinetics. Introduction *e-mail : jklee@kumoh.ac.kr 1598-5032/02/34-06+2002 Polymer Society of Korea Cationic latent cure system for epoxy resin is of interest since it can simplify cure processing and improve long term stability at room temperature and electrical properties, compared to the common amine and anhydride curing agent systems. Various cationic catalysts were introduced in the literature having photo and/or thermal latency for curing epoxy resins. 1-7 Recently, N-benzylpyrazinium salts showing thermal latency for epoxy resins have been synthesized 8 and used to study cure behavior, rheological and thermal/ mechanical properties during and after cure. 8-11 The concentration of the latent catalyst ranging from 0.5 to 5 wt% has been used for the studies. It revealed that the optimum concentration of the catalyst was 1.5 or 2 wt% for various physical properties. During cure of thermosetting materials, the glass transition temperature (T g ) as an index of degree of chemical conversion increases greatly as low molecular weight liquid are transformed into high molecular weight solid. It is known that the T g is an important parameter in analyzing cure behavior so that the variation of T g with cure time has been studied for dicyanate ester-polycyanurate, epoxy/amine and epoxy/anhydride system. 12-17 It has also been described that T g can be used to study the cure kinetics for various epoxy resin and curing agent systems. 15-17 In this study, we investigated the cure behavior and cure kinetics of an epoxy resin in the presence of an optimum amount (2 wt%) of N-benzylpyrazinium hexafluoroantimonate (BPH) as a cationic catalyst in order to understand further the cure process of the system by means of a differential 34

Cationic Cure of Epoxy Resin Figure 1. Chemical structures of the reactants for epoxy resin and catalyst used in this study. scanning calorimetric technique. Dynamic mechanical properties were examined during curing isothermally. Experimental A diglycidyl ether of bisphenol-a (DGEBA, YD 128 from Kukdo Chem. Co., Korea, equivalent weight = 185~ 190 g/eq.) epoxy resin and 2 wt% of an N-benzylpyrazinium hexafluoroantimonate (BPH) catalyst were mixed thoroughly in acetone for 10 min. The chemical structures of the reactants are shown in Figure 1. The unreacted mixture in a vial, covering with aluminum foil to protect from light, was stored in refrigerator. The cure of the epoxy resin was studied using differential scanning calorimetry (DSC, Du Pont 910). DSC measurements were made on the mixture containing approximately 10-15 mg in a hermetic pan. The glass transition temperature (T go ) and the total exothermic heat of cure ( H T ) of the unreacted sample were determined from the DSC scan. The uncured sample in the hermetic pan was partially cured in a dry oven for pre-selected periods (10 to 1,020 min) at different cure temperatures (110 to 150 o C). DSC scans for the partially cured sample were taken to obtain T g and residual heat of cure ( H R ). All experiments for the samples uncured and partially cured were performed in the temperature range from 60 to 300 o C at a heating rate of 10 o C/min in a dry nitrogen atmosphere. When the endothermic peak due to isothermal physical aging in a vitrified region was developed near T g for partially cured samples, the peak was eliminated by heating the samples to just above the endothermic peak, quenching to room temperature. The sample was then subsequently reheated from room temperature to 300 o C. This procedure is essential to determine not only the glass transition temperature but also the residual heat of cure accurately as used in other works. 12,17 A fully cured sample was prepared by curing at 140 o C for 600 min and postcuring at 190 o C for 120 min. The glass transition temperature (T g ) of the fully cured was 175 o C. Dynamic mechanical analysis (DMA, Du Pont 982) was employed to observe the cure behavior of the epoxy mixture that is impregnated into glass mattress. The impregnated sample was placed in the heater after squeezing out the excess resin between aluminum foil. The cure of the sample was analyzed in terms of storage modulus and tanδ. The modulus obtained in this experiment is a relative value because of the composite nature of the specimen (resin mixture + glass mattress). Uncured samples were heated isothermally at the temperatures as in DSC experiments for times up to 2,000 min under a dry nitrogen atmosphere. Frequency and amplitude of the oscillation are 0.1 Hz and 0.2 mm, respectively. Results and Discussion Figure 2 shows typical DSC thermograms of samples Figure 2. DSC thermograms of samples uncured and partially cured, showing the glass transition temperature (T g ) and exothermic peak ( H). Macromol. Res., Vol. 10, No. 1, 2002 35

J. K. Lee et al. uncured and partially cured at 140 o C for 30 min in the epoxy resin/n-benzylpyrazinium hexafluoroantimonate (BPH) catalyst system. The glass transition temperature (T go ) and the total exothermic enthalpy ( H T ) of the uncured mixture determined from the DSC curve were -18.4 o C and 400 J/g, respectively. For the partially cured sample, the T g increased to 10 o C and the residual enthalpy value ( H R ) decreased to 200 J/g. Multi-exothermic peaks on the DSC thermogram for the uncured sample were observed in this study; two small peaks at lower temperatures and one large peak at a higher temperature. According to the previous work, 8 it is considered that the small exothermic peaks at lower temperatures are due to the two different initiation reactions between BPH and epoxy and between hydroxyl group and epoxy as shown below: 9 Figure 3. Glass transition temperature (T g ) vs ln(time) during curing at various cure temperatures from 110 to 150 o C. After curing at 140 o C for 30 min, the small peaks are disappeared completely and the main peak at higher temperature becomes smaller. The 2 wt% of BPH in this study is considered to display rather pronounced small exothermic peaks, which disappear after curing at a short period of time before gelation. 8 Cure Characteristics. The glass transition temperature, which is one of the important parameters in analyzing cure behavior, was determined from the inflection point of the step transition on DSC thermogram. The variations of T g with ln(time) during isothermal curing at different temperatures (T c = 110, 120, 130, 140, and 150 o C) are shown in Figure 3. The T g increases rather slowly in the beginning of cure, rapidly in the middle, and levels off thereafter at lower cure temperatures (i.e. 110, 120, and 130 o C), while the T g increases rapidly from the beginning at higher cure temperatures (i.e. 140 and 150 o C). Figure 4 shows the corresponding conversion (x) with ln(time) calculated by the following equation from the total ( H T ) and residual heat of cure Figure 4. Corresponding conversion (x) vs ln(time) during curing at various cure temperatures from 110 to 150 o C. ( H R ). H x = X --------- R (1) H T The increasing trend of conversion with cure time is similar to that of T g in Figure 3. After curing for 1020 min, the T g and conversion increased up to 105 o C and 0.79 at T c = 110 o C and 150 o C and 0.93 at T c = 140 o C, respectively. 36 Macromol. Res., Vol. 10, No. 1, 2002

Cationic Cure of Epoxy Resin Figure 5. Log(relative storage modulus) and tanδ vs ln(time) at various cure temperatures from 110 to 150 o C. Cure of thermosetting polymers, especially epoxy, can be characterized from the variations of the storage and loss modulus (or tanδ) by means of dynamic mechanical analysis. Figure 5 has the variations of (a) relative storage modulus and (b) tanδ during isothermal curing at isothermal temperatures, ranging from 110 to 150 o C. For all the temperatures, the storage modulus increases slowly and then rapidly, and levels off vs ln(time). Whole curve shifts to a shorter time scale with increase of cure temperature, meaning that cure reaction is accelerated. All tanδ curves show two peaks as marked by arrows, which correspond to gelation at a shorter cure time and vitrification at a longer cure time. At gelation, viscosity increases dramatically by the formation of branched molecules with infinite molecular weight. A glassy material is formed as a consequence of the network becoming tighter through further chemical crosslinking reaction at vitrification. As in the modulus, the tanδ curves move to a shorter time scale. Gelation and vitrification time of 145 and 1,355 min at T c = 110 o C are shortened to 12 and 230 min at T c =150 o C, respectively. The times for the gelation and vitrification are Table I. Gelation and Vitrification Times Obtained from DMA Cure Temperature ( o C) Gelation Time (min) Vitrification Time (min) 110 145 1,355 120 96 984 130 40 593 140 23 420 150 12 230 Figure 6. Horizontal shift of T g vs ln(time) data (T r = 130 o C), showing superposition below T g =100 o C. Note the break in the increasing slope at T g =5 o C. listed in Table I. Horizontal Shift of T g and Conversion. The data of T g vs ln(time) in Figure 3 were horizontally shifted at a reference temperature (T r ) of 130 o C and the result is displayed in Figure 6. All the glass transition temperature data are superposed and formed a master curve below T g =100 o C. The superposition is known to occur in a kinetically-controlled regime where the reaction mechanism changes to a diffusioncontrolled (above T g = 100 o C in this study) as observed in other studies. 15-17 In this Figure, it is very interesting that there is a distinct change of slope in the increase of T g in the superposed region. Therefore, the master curve can be divided into two regions in term of T g : R I (T g ) = -18.4 to 5 o C and R II (T g ) = 5 to 100 o C, as marked in the figure on a vertical axis. The horizontal shift of the data of x vs ln(time) in Figure 4 were also made at the same reference temperature (T r = 130 o C) and the result is shown in Figure 7. The shift produced a master curve and the break in the conversion increase was observed as shown in the shift of T g. The two regions in terms of x are R I (x) = 0 to 0.24 and R II (x) = 0.24 to 0.70. The R I where T g or x increases gradually with isothermal cure time appears to be closely related to the small cure exotherms at lower temperatures on the DSC thermogram for uncured samples shown in Figure 2. The slope change in the T g and x confirms the existence of the initiation reactions (epoxy/bph and epoxy/hyroxyl group) involved in the cure, proposed in earlier reports. 8,9 Consequently, the initiation reactions were found to take place in the early stage of cure below T g =5 o C or x = 0.24. Macromol. Res., Vol. 10, No. 1, 2002 37

J. K. Lee et al. Figure 7. Horizontal shift of x vs ln(time) data (T r =130 o C), showing superposition below x = 0.70. Note the break in the increasing slope at x = 0.24. Cure Kinetics. Among the methods in analyzing experimental results for the cure kinetics, a time-temperature superposition method has been used to determine the activation energy (E a ). 18 When the rate of reaction is kineticallycontrolled, the rate is given by the Arrhenius rate equation. dx ----- = A dt E exp -------- a RT c fx ( ) where E a is the apparent activation energy for the overall reaction, T c is the cure temperature, and f(x) is a function of conversion (x) independent of temperature. Rearranging eq. 2 and integrating yields, x E a dx ln -------- = lna+ lnt -------- W fx ( ) RT c From a one-to-one relationship between x and T g, eq. 3 can be written as: E a FT ( g ) = lna+ lnt -------- RT c At the arbitrary reference temperature (T r ), eq. 4 becomes E FT ( g ) = lna+ lnt a r ------- RT r (2) (3) (4) (5) Figure 8. Shift factor vs 1/T to obtain activation energy. Table II. Activation Energies for Epoxy Systems Obtained by a Time-temperature Superposition Method Materials Activation Energy References # (kcal/mol) Epoxy Novolac/DDS a 14.6 15 DGEBA/TMAB b 15.2 16 DGEBA/Anhydride/Imidazole 15.2 17 DGEBA/BPH 12.5 This work a DDS, 4,4-diaminodiphenylsulfone. b TMAB, trimethylene glycol di-p-aminobenzoate. The shift factor, A(T), is defined as ln time difference at constant T g, subtracting eq. 4 from eq. 5, E a AT ( ) lnt r lnt ---- X X = = ---- ---- R T c The shift factor at a reference temperature (i.e. T r = 130 o C) in Figure 6 was obtained and plotted against inverse of absolute temperature in Figure 8. The shift factor varies linearly with 1/T. Using eq. 6, activation energy was determined from the slope of the figure. The activation energy was 12.5 kcal/mol in this epoxy/bph system, and compared with other values obtained by the same method in Table II. The result shows that the activation energy of the catalytic system is smaller than that of amine or anhydride systems, meaning that the curing may proceed readily at the cure temperatures tested. Relationship between T g vs Conversion. For many thermosetting systems, there is a unique relationship between T g T r (6) 38 Macromol. Res., Vol. 10, No. 1, 2002

Cationic Cure of Epoxy Resin Research Fund from Kumoh National University of Technology. References Figure 9. Relationship between T g and conversion for all isothermal cure temperatures. and conversion independent of cure temperature. 12 The relationship can be described by the empirical DeBenedetto equation: 19 T g T ------------------- go T g T go λx = -------------------------- X ( X λ)x where T go and T g are the glass transition temperatures of the unreacted and the fully cured material, respectively, and λ is a material constant. All the data for T g were plotted vs conversion in Figure 9 with the best fitted line of the DiBenedetto equation. The material constant value (λ) was 0.45. Note that the T g vs conversion relationship in the figure is not influenced by the presence of the initiation reactions (i.e., epoxy/bph and epoxy/hyroxyl group). (7) Acknowledgement. This paper is supported by a 2000 (1) O. Shimomura, I. Tomita, and T. Endo, J. Polym. Sci.: Part A: Polym. Chem., 39, 868 (2001). (2) K. Morio, H. Murase, and H. Tsuchiya, J. Appl. Polym. Sci., 32, 5727 (1986). (3) S. Murai, Y. Nakano, and S. Hayase, J. Appl. Polym. Sci., 80, 181 (2001). (4) A.P. Pappas and L.H. Hill, J. Coating Tech., 53, 43 (1981). (5) T. Endo, A. Kikkawa, H. Uno, and H. Sato, J. Polym. Sci., Polym. Lett. Ed., 27, 73 (1989). (6) J. Gu, S. C. Narang, and E. M. Pearce, J. Appl. Polym. Sci., 30, 2997 (1985). (7) J. Park, Korea Polym. J., 9, 206 (2001). (8) Y.C. Kim, S.-J. Park, and J.-R. Lee, Polym. J., 29, 759 (1997). (9) S.-J. Park, T.-J. Kim, and J.-R. Lee, J. Polym. Sci.:Part B:Polym. Phys., 38, 2114 (2000). (10) G.-H. Kwak, S.-J. Park, and J.-R. Lee, J. Appl. Polym. Sci., 78, 290 (2000). (11) S.-J. Park and H.-C Kim, J. Polym. Sci.:Part B:Polym. Phys., 39, 121 (2001). (12) G. Wisanrakkit and J. K. Gillham, J. Coating Tech., 62, 35 (1990). (13) A. Osei-Owusu, G. C. Martin, and J. T. Grotto, Polym. Eng. Sci., 31, 1604 (1991) (14) V. Micro, Z. Q. Cao, F. Mechin, and J. P. Pascault, Am. Chem. Soc., Polym. Mater. Sci., Eng. Prepr., 66, 451 (1992). (15) P. A. Oyanguren and R. J. J. Williams, J. Appl. Polym. Sci., 47, 1371 (1993). (16) S. L. Simon and J. K. Gillham, J. Appl. Polym. Sci., 46, 1245 (1992). (17) W. H. Park, J. K. Lee, and K. J. Kwon, Polym. J., 28, 407 (1996). (18) R. B. Prime, Thermal Characterization of Polymeric Materials, E. A. Turi, Ed., Academic Press, New York, 1997, Vol. 2. (19) L. E. Nelson, J. Macromol. Sci. Rev. Macromol. Chem., C3, 69 (1969). Macromol. Res., Vol. 10, No. 1, 2002 39