Cure Thesis

download Cure Thesis

of 76

Transcript of Cure Thesis

  • 8/2/2019 Cure Thesis

    1/76

    RELATIONSHIP BETWEEN THE PHYSICAL PROPERTIES

    AND CURING SYSTEM OF AN EPOXY MATRIX MATERIAL

    by

    Robert Rainer Pittroff

    Submitted in partial fulfilment of the requirements for the degree

    MAGISTER TECHNOLOGIAE: ENGINEERING: MECHANICAL

    in the

    Department of Mechanical Engineering

    FACULTY OF ENGINEERING AND THE BUILT ENVIRONMENT

    TSHWANE UNIVERSITY OF TECHNOLOGY

    Supervisor: Prof O.C. Vorster

    Co-Supervisor: Mr. D Louwrens

    December 2007

  • 8/2/2019 Cure Thesis

    2/76

    ii

    DECLARATION BY CANDIDATE

    I hereby declare that the dissertation submitted for the degree M Tech:

    Engineering: Mechanical, at Tshwane University of Technology, is my own

    original work and has not previously been submitted to any other institution of

    higher education. I further declare that all sources cited or quoted are

    indicated and acknowledged by means of a comprehensive list of

    references.

    Robert R Pittroff

    Copyright Tshwane University of Technology 2007

  • 8/2/2019 Cure Thesis

    3/76

    iii

    Acknowledgement

    I would like to express my sincere gratitude and appreciation to the following:

    My supervisor, Prof. O.C. Vorster for his patience and guidance.

    My co-supervisor, Mr. D. Louwrens for his positive attitude and guidance.

    Tshwane University of Technology for financial assistance.

    National Research Foundation for financial assistance.

    Plastomax Supplies for supplying the material at no cost for this study.

    Members of the Tshwane University of Technology, specifically the

    Department of Polymer Technology and Department of Mechanical

    Engineering for their assistance.

    My family for their continued support

  • 8/2/2019 Cure Thesis

    4/76

    iv

    Abstract

    The effect of curing conditions such as time and temperature on the physical

    properties of an epichlorohydrin-derived, aliphatic amine cured epoxy system

    has been investigated.

    The cure kinetics and thermal behaviour of a setting system was monitored

    experimentally. The effect of the reaction temperature on gelation (tgel) and

    vitrification times (tvit) were determined using parallel plate rheology and

    Differential Scanning Calorometry (DSC) respectively. Rheology was used to

    determine tgel, by monitoring the complex viscosity, complex modulus and tan

    against time, whereas the effect of the curing temperature (Tcure) on the glass

    transition temperature (Tg) of the system was determined from DSC.

    The cure kinetics and thermal behaviour as determined were used to calculate atime-temperature-transformation (TTT) isothermal cure diagram. The TTT

    isothermal cure diagram was used as a framework to monitor and characterize

    changes during the curing process of a thermosetting system.

    The gelation contour corresponds to the point where viscosity tends to infinity,

    while the vitrification contour represents Tg rising to Tcure. Rheology

    measurements under isothermal conditions indicate that Tgel decreases with an

    increase in Tcure.The effect of the curing temperature (Tcure) and curing time (tc) on Tg, dynamic

    mechanical properties, and indentation hardness of the system was determined

    on cured samples characterized with the TTT isothermal cure diagram. Values

    for Tg, and information on the dynamic mechanical properties of the epoxy

    system was determined by means of dynamic mechanical thermal analysis

    (DMTA). Indentation hardness of the cured epoxy system was determined by

    means of measurements on samples using a Hand-Held Portable Hardness

    Tester (Barcol Impressor, Model No. 934-1).

    It is shown that there is a relationship between the cure conditions and the

    mechanical properties of an epoxy matrix material.

  • 8/2/2019 Cure Thesis

    5/76

    v

    Samevatting

    Die effek van die verharding kondisies soos tyd en temperatuur op die fisiese

    einskappe van n epichlorohydrin-derived, aliphatic amine verhardingsreaksie

    epoksie systeem word in hierdie studie pepaal.

    Die verhardingreaksiekinetika en termiese verloop verharding systeem word

    deur eksperimente bepaal. Die effek op jelpunt (tjel) en vitrifikasie (tvit) punt word

    deur parallelplaatspanningsreometrie en differensiele skanderingskalorimetrie

    eksperimente bepaal. Die komplekse viskositeit asook komplekse modulus wat

    deur spanningsreometrie pebaal is, word gebruik om tjel te definer, terwyl

    differensile skandeerkalorimetrie gebruik word om glas transformasie

    temeratuur uitg te werk, deur bepaling van verhardingstye en verhardings

    temperature.

    Die verhardingsreaksiekinetika en termiese verloop wat bepaal is was gebruik

    om n tyd-temperatuur-transformasie (TTT) isotermiese verharding diagram te op

    te trek. Die TTT diagram is gebruik as n raamwerk om die verharding reaksie

    van n reaktiewe systeem te beskryf.

    Die effek op die verharding temperatuur en verhading tyd op die

    glastransformasie temperatuur (Tg) sowel as dinamiese meganiese einskappe is

    deur dinamiese meganiese termiese analise (DMTA) bepaal. Die hardheid van

    verharde stelsels is bepaal deur hardheid analise met n Hand-Held Portable

    Hardness Tester (Barcol Impressor, Model No. 934-1).

    Daar is bewys dat daar wel n verhouding tussen die verharding kondisies en

    meganiese einskappe van n epoksie system bestaan.

  • 8/2/2019 Cure Thesis

    6/76

    vi

    CONTENTS

    PAGE

    ACKNOWLEDGEMENTS. iii

    ABSTRACT. ivEKSERP.. v

    LIST OF TABLES vi

    LIST OF FIGURES. vii

    GLOSSARY. viii

    CHAPTER 1

    1.1 INTRODUCTION ................................................................. 1

    1.2 BACKGROUND INFORMATION.................................................... 21.3 OBJECTIVES.. 3

    CHAPTER 2

    2.0 LITERATURE SURVEY................................................................... 4

    2.1 EPOXY RESIN................................................................................. 4

    2.2 KINETICS OF CURE......................................................... 5

    2.2.1 FACTORS INFLUENCING CURE................................................... 8

    2.2.1.1 FUNCTIONALITY............................................................................ 9

    2.2.1.2 MOLECULAR MASS....................................................................... 9

    2.2.2 COMPETITION BETWEEN CURE AND DEGRADATION.. 10

    2.3 GEL POINT..................................................................................... 10

    2.3.1 MICROSCOPIC GELATION........................................................... 11

    2.3.2 MACROSCOPIC GELATION.......................................................... 12

    2.3.3 GELATION STUDY.................................................................. 12

    2.3.3.1 CURING ANALYSIS: RHEOLOGICAL ANALYSIS......................... 13

    2.4 VITRIFICATION.............................................................................. 15

    2.4.1 CURE TEMPERATURE.................................................................. 16

    2.4.2 GLASS TRANSITION TEMPERATURE......................................... 16

    2.4.3 VITRIFICATION STUDY................................................................. 17

    2.4.3.1 CURING ANALYSIS: DIFFERENTIAL SCANNING

    CALORIMETRY............................................................................... 17

  • 8/2/2019 Cure Thesis

    7/76

    vii

    2.5 CHARACTERIZATION OF THE CURE PROCESS......................... 18

    2.6 MECHANICAL PROPERTIES.......................................................... 19

    2.6.1 MECHANICAL PROPERTIES: MOLECULAR INFLUENCES.. 20

    2.6.1.1 MOLECULAR WEIGHT.................................................................... 20

    2.6.1.2 CROSS-LINKING DENSITY............................................................. 20

    2.6.2 MECHANICAL PROPERTY ANALYSIS TECHNIQUES.................. 21

    2.6.2.1 DYNAMIC MECHANICAL ANALYSIS.............................................. 21

    2.6.2.1.1 INSTRUMENTATION....................................................................... 22

    2.6.2.1.2 DMA AND DMTA ANALYSIS........................................................... 23

    2.6.2.2 INDENTATION HARDNESS............................................................ 24

    CHAPTER 3

    3.0 RESEARCH METHODOLOGY....................................................... 25

    3.1 MATERIALS USED AND THEIR PREPARATION......................... 25

    3.2 GELATION SYUDY........................................................................ 25

    3.3 VITRIFICATION STUDY................................................................ 27

    3.3.1 SAMPLE PREPARATION.............................................................. 27

    3.4 DYNAMIC MECHANICAL THERMAL ANALYSIS......................... 28

    3.4.1 SAMPLE PREPARATION............................................................. 28

    3.5 INDENTATION HARDNESS......................................................... 313.4.2 SAMPLE PREPARATION............................................................. 31

    3.4.3 EXPERIMENTAL PROCEDURE 31

    CHAPTER 4

    4.0 RESULTS AND DISCUSSION........................................................ 32

    4.1 INTRODUCTION............................................................................. 32

    4.2 GELATION STUDY......................................................................... 324.3 VITRIFICATION STUDY................................................................. 34

    4.3.1 DETERMINING Tg.......................................................................... 35

    4.3.1.1 TgVERSUS TCure............................................................................ 36

    4.3.1.2 TgVERSUS tCure............................................................................. 37

    4.3.1.3 ISO- TgCONTOURS ...................................................................... 39

  • 8/2/2019 Cure Thesis

    8/76

    viii

    4.3.1.4 PREDICTION OF tvit........................................................................ 40

    4.4 TIME TEMPERATURE TRANSFORMATION DIAGRAM 45

    4.5 MECHANICAL PROPERTY VS. CURE RELATIONSHIP.............. 46

    4.5.1 AREAS FOR ANALYSIS................................................................. 46

    4.5.2 DYNAMIC MECHANICAL THERMAL ANALYSIS (DMTA)............. 48

    4.5.2.1 STORAGE MODULUS..................................................................... 49

    4.5.2.2 GLASS TRANSITION TEMPERATURE (Tg)................................... 51

    4.5.2.3 STORAGE MODULUS (E') RELATED TO STIFFNESS 52

    4.5.2.4 LOSS MODUDLUS RELATED TO IMPACT RESISTANCE. 53

    4.5.3 HARDNESS ANALYSIS............................................................... 56

    CHAPTER 5

    5.0 CONCLUSIONS AND RECOMMENDATIONS............................... 58

    5.1 CONCLUSIONS.. 58

    5.2 RECOMMENDATIONS.. 60

    BIBLIOGRAPHY. 61

    APPENDIX A: Material data sheet for Epon Resin 826.

  • 8/2/2019 Cure Thesis

    9/76

    ix

    LIST OF FIGURES

    PAGE

    2.1 A generalized time-temperature-transformation (TTT) cure

    diagram. Times to gelation and vitrification during isothermal cure

    against temperature demarks the boundaries of the four

    distinct regions states of matter: liquid, gelled rubber, gelled glass,

    and ungelled glass. (Taken from Prime (1997:1384)).......................... 7

    3.1 Rheometric Dynamic Stress Rheometer (RS-500)............................... 26

    3.2 Typical plot of the loss and storage moduli, tan and complex

    viscosity (*) against time for Rheological measurements................... 26

    3.3 Typical plot of a DSC scan of endothermic heat flow (mW) against

    temperature(C)................................................................................... 28

    3.4 Perkin-Elmer Differential Scanning Calorimeter (DSC-7).................... 28

    3.5 Typical plot of the storage modulus (E') and tan against

    temperature. (Taken from Laza et al. (1998:43)) ................................ 30

    3.6 Rheometric Scientific Solid Analyzer RSA II (DMTA) in dual

    cantilever mode................................................................................... 30

    3.7 The Impressor: A Hand-Held Portable Hardness Tester

    (Barcol Impressor, Model No. 934-1)................................................... 31

    4.1 Time-temperature-transformation (TTT) diagram: Gelation contour.... 34

    4.2 Tg (C) vs. Tcure (C) for different isothermal heating times (1 hour

    through to 72 hours). The reference line of Tg = Tcure is also

    included............................................................................................... 36

    4.3 Tg (C) vs. Log Time (min) for different isothermal temperatures

    (Tcure).................................................................................................... 37

    4.4 Limiting Tg (C) vs. Tcure(C),............................................................... 38

    4.5 Time-temperature-transformation (TTT) diagram: Iso-Tgcontours...... 39

    4.6 Ln time (min) to fixed Tgvalues (from Figure 4.5) vs. 1/Tcure(K)......... 41

    4.7 Ln time (min) to macroscopic gelation vs. 1/Tcure(K)........................... 44

    4.8 Summary time-temperature-transformation (TTT) diagram................. 46

    4.9 Detailed TTT-Isothermal cure diagram showing areas for DTMA

    analysis................................................................................................ 47

  • 8/2/2019 Cure Thesis

    10/76

    x

    4.10 DMTAscans. Tan and storage modulus (E') vs. temperature (C)

    for samples cured for different cure times (tcure, min) at different cure

    temperatures (Tc).

    4.10-a: DMTA scans forTcure= 70C.......................................... 494.10-b: DMTA scans forTcure= 90C.......................................... 494.10-c: DMTA scans forTcure= 110C.......................................... 50

    4.11 Trends in Tg values for samples cured at different temperatures (Tc)

    and for different times (tcure)................................................................. 51

    4.12 Trends in E' values for samples cured at different temperatures (Tc)

    and for different times (tcure)................................................................. 53

    4.13 Determination of the area under the storage modulus curve............... 54

    4.14 Trends in impact strength values for samples cured at different

    temperatures (Tc) for different cure times (tcure)................................... 55

    4.15 Trends in Barcol Hardness values for samples cured at different

    temperatures (Tc) and for different times (tcure).................................... 56

  • 8/2/2019 Cure Thesis

    11/76

    xi

    LIST OF TABLES

    PAGE

    2.1 Outline and classification of epoxy reactive groups, constituents, and

    ammonia and amine reaction groups. Taken from Ellis

    (1993:1,3&12) 5

    2.2 Glossary of curing terms and characteristic curing parameters.

    Taken from Prime (1997:1386)............................................................ 8

    4.1 Average gel times (Log Time) for different frequencies at the

    respective cure temperatures.............................................................. 33

    4.2 Average glass transition temperatures for samples cured for

    different times at different cure temperatures...................................... 35

    4.3 Numerical relationship between Ln time to a fixed Tg and thereciprocal ofTcure(1/K)......................................................................... 40

    4.4 Calculated times to vitrification for Tcure = Tg corresponding to

    isothermal cure temperatures from Figure 4.2..................................... 43

    4.5 Results from DMTA............................................................................. 50

  • 8/2/2019 Cure Thesis

    12/76

    1

    CHAPTER 1

    1.1 Introduction

    Thermoset systems possess good dimensional stability, thermal stability,

    chemical resistance, and electrical properties. Prime (1997:1381) believes that

    the increasing use of thermoset material as adhesives, matrix material in

    reinforced composites, and protective coatings in aeronautical, automotive,

    construction, electrical and medical application is consequence to thermosets

    possessing these properties.

    Using composite materials technology in design and manufacture in engineering

    fields, as explained by Jones (1993:256), exploits the high specific moduli and

    strength of reinforcing fibres, supported and held together by epoxy matrices to

    produce low density high performing structures. According to Morgan

    (1997:2097), epoxy resin systems are generally used for composite matrices,

    due to the suitability in composite fabrication.

    Optimal component design requires an in-depth knowledge of the properties and

    mechanical capabilities and limitations of constituent materials of composites,

    and their correct application. The reinforcing fibers provide strength and stiffness,while the epoxy accounts for the desired shape and load distribution between the

    reinforcing fibers.

    Morgan (1997:2198), Yasmin and Daniel (2004:8211), and Nicholson et al. (1)

    are in general agreement that the failure or degradation of composite component

    performance can be linked to matrix dominated properties. The weak link in the

    composite is the structural integrity, mechanical, and physical response of this

    interfacial region. Favorable response of composite structures to applied loads

    requires competent structural design based on knowledge of the mechanical

    properties of the constituent parts of the composite.

  • 8/2/2019 Cure Thesis

    13/76

    2

    1.2 Background information

    Epoxy is a title given to the class of material that has an epoxide group or oxirane

    ring in its chemical structure. Epoxies exist either as liquids with low viscosity or

    as solids. Curing of an epoxy system as explained by Sun (2002:11), is through

    the ring opening reaction of epoxides where addition of molecules produces a

    higher molecular mass and finally resulting in a three-dimensional structure. The

    active epoxide groups in uncured epoxy resins react with various curing agents

    or hardeners that contain, amongst other, hydroxyl, carboxyl, amine, and amino

    groups.

    Epoxy resins can be divided into two major groups: epichlorohydrin-derived

    resins and cycloaliphatic resins. This particular study will focus on the former

    group, where there is a reaction with a common polyhydroxy compound, such as

    bisphenol A, which is cured by means of an aliphatic amine curing agent.

    Processing of thermosetting based composite structures involve curing cycles of

    the epoxy systems at different isothermal curing temperatures. The degree of

    cure of thermosetting systems is determined by these curing cycles and has an

    important effect on the mechanical properties of the final product. To determine,

    or to a certain extent predict, the mechanical properties through optimized curing

    cycles requires thorough understanding of cure kinetics and characteristics of

    epoxy systems. The kinetics of the curing process was studied by means of

    parallel plate rheology (Halsz and Vorster (2000, 2004), Lange et al. (1996),

    Sun (2002)) and differential scanning calorimetry (DSC) (Gan et al. (1989),

    Gillham (1993), Gumen et al. (2000) and Nez et al. (2000)).

    Information gained from the rheology and DSC analysis was then used to

    generate additional information on the physical state of the epoxy resin. The

    information was used to develop an isothermal time-temperature-transformation

    (TTT) cure diagram (Gillham (1979, 1982, 1990 and 1993), Gan et al.(1989),

    Jiawu et al. (2000), Nez et al. (2000), Sun (2002), and Reghunadhan Nair

  • 8/2/2019 Cure Thesis

    14/76

    3

    (2004:38)). The TTT cure diagram facilitates characterizing the thermosetting

    system.

    By making use of the isothermal TTT cure diagram epoxy samples were

    characterized and cured according to specific cure temperatures (Tcure) and

    curing times (tcure). These samples were then subjected to dynamic mechanical

    thermal analysis (DMTA) scans as well as Barcol hardness test to determine

    mechanical properties of the cured samples under different curing conditions.

    1.3 Objectives

    This study focuses on gaining insight into the properties and mechanical

    capabilities by determining the relationship between the curing conditions and

    physical properties of an epoxy matrix. It aims to:

    Establish the kinetics of cure of the epoxy thermosetting system

    through the use of rheology and differential scanning calorimetry.

    Use the kinetics of cure to develop and construct an isothermal

    time-temperature-transformation (TTT) cure diagram.

    To ascertain the mechanical properties of the cured samplesthrough dynamic mechanical thermal analysis (DMTA) and Barcol

    hardness indentation measurements.

    Using the TTT diagram and mechanical properties from DMTA

    scans and Barcol indentation tests to establish a relationship

    between the curing conditions and physical properties of the epoxy

    thermosetting system.

  • 8/2/2019 Cure Thesis

    15/76

    4

    CHAPTER 2

    2.0 Literature Survey

    2.1 Epoxy Resin

    Commercial epoxy resins were first marketed in the 1940s as result of

    independent work by Pierre Castan in Switzerland, and Sylvan Greenlee in the

    United States. Even though patents for similar resins had been in existence since

    the 1930s, Ellis (1993:1) explains that products that would be called epoxy resins

    in modern times, were synthesized as early as 1891. It is interesting to note that

    early epoxy resins were reaction products of bisphenol A and epichlorohydrin,

    and although various other types of resins are available, this is still the major

    manufacturing route for most resins available today.

    Epoxy resin is generally found in the form of liquids with low viscosities or as

    solids. Prime (1997:1716) and Sun (2002:11) report that epoxy resins form part

    of a class of material classified and characterized according to the reactive

    epoxide group, or oxirane ring that forms part of the chemical structure. Ellis

    (1993:1) points out that the term epoxy resin is generally applied to both the un-

    reacted prepolymers and cured resins, even thought all the reactive groups may

    have reacted in the cured resins.

    Epoxy resins can be divided into two major groups: epichlorohydrin-derived

    resins and cycloaliphatic resins. This particular study will focus on the former

    group, where there is a reaction with a common polyhydroxy compound, such as

    bisphenol A, which is cured by means of an aliphatic amine curing agent. The

    reactive epoxy group, bisphenol A and epichlorohydrin reagents, as well as the

    aliphatic amine group reaction is outlined and classified in Table 2.1.

  • 8/2/2019 Cure Thesis

    16/76

    5

    Table 2.1 Outline and classification of epoxy reactive groups, constituents, and ammonia and

    amine reaction groups. Taken from Ellis (1993:1,3&12)

    Ellis (1993:72) asserts that the cure of an epoxy resin involves reaction between

    epoxy and hardener reactive groups, and that during cure, a liquid or fluid resin-

    hardener mixture is converted to a solid.

    2.2. Kinetics of cure

    Thermosetting resin is an appropriate title in that it describes the final infusible

    and insoluble state of a set, or cured system. Liquid resin is mixed with a

    hardener or catalyst according to a prescribed ratio in order to initiate the curing

    process. This process is known as polymerization, curing or cross-linking, and is

    a permanent change from a viscous liquid to a viscoelastic fluid, to an elastic gel

  • 8/2/2019 Cure Thesis

    17/76

    6

    (or rubber), and finally to a glassy state. Heat or another form of energy is often

    required to initiate cure. Proper formulation and complete processing will result in

    a polymer with a highly cross-linked, infinite three-dimensional network structure,

    a consequence of the chemical reactions that accompany cure. It is the

    polymerization or curing process that separates thermosets from other polymer

    materials such as thermoplastic materials.

    According to Ellis (1993:72-74) and Prime (1997:1380-1388) the complexity of

    the curing process of a thermoset is due to steps that are necessary during the

    handling and storage of the material, prior to cure, and situational factors that

    influence the process, and state or phase changes that the material undergoes

    during cure.

    The major changes to the properties of thermosetting resins from being a viscous

    fluid to a glassy solid are a result of the cure process. These changes as cure

    proceeds can be represented in a time-temperature-transformation (TTT)

    diagram as represented in Figure 2.1 below.

  • 8/2/2019 Cure Thesis

    18/76

    7

    Figure 2.1.A generalized time-temperature-transformation (TTT) cure

    diagram. Times to gelation and vitrification during isothermal cure against

    temperature demarks the boundaries of the four distinct regions states of

    matter: liquid, gelled rubber, gelled glass, and ungelled glass. (Taken from

    Prime (1997:1384))

    The first phase transformation from a viscoelastic fluid to an elastic gel is known

    as the gel point, and occurs at the gel time (or gelation times), tgel. A distinction

    must be made between molecular gelation and macroscopic gelation. Molecular

    gelation occurs at a specific stage during the chemical reaction, and marks the

    start of the formation of a cross-linked network, whereas the macroscopic

    consequence of gelation shows a rapid increase in viscosity, and development of

    elastic properties.

    The next phase transition from a gel to a glassy solid is vitrification, with glass

    transition measurements of Tg and times to vitrification, tvit. Prime (1997:1383)

    points out that vitrification occurs as a result of the chemical reaction that

    accompanies cure, and is independent of gelation. Vitrification of the system can

    occur without gelation.

  • 8/2/2019 Cure Thesis

    19/76

    8

    Table 2.2 Glossary of curing terms and characteristic curing parameters. Taken from Prime

    (1997:1386)

    Chemical conversion (e.g., of epoxide groups), fraction reacted, degree ofcure

    gel at gel pointult Maximum achievable extent of conversion 1

    tgel Time to gelation (gel time)

    tvit Time to vitrification

    Tcure Cure temperature, a process parameter

    Tcure,0Temperature below which no significant reaction of uncured resin mixtureoccurs in a reasonable time period (cf. storage temperature for uncured resinmixture)

    gelTcure Lowest temperature at which gelation occurs in the liquid state

    Tcure, Minimum temperature at which ultimate conversion occurs in a reasonabletime

    Tg Glass transition temperature, a material property

    Tg0 Tg for thermoset with degree of conversion = 0

    gelTg Tg for thermoset with degree of conversion gel

    Tg, Tg for thermoset with degree of conversion = 1

    Prime (1997:1383) and Gillham (1989:804, 1993) repeatedly point out that the

    onset of vitrification marks a decrease in the rate of reaction. This decrease is a

    result of the change in reaction mechanism from being kinetically controlled to

    diffusion control.

    2.2.1. Factors influencing cure

    Manufacturing and processing of composite parts destined for commercial

    applications calls for the least amount of cycle/process time, yielding improved

    properties. It is therefore imperative that the finished product demand set by themarket be met; Production and cost effective manufacture therefore compels the

    use of additives such as catalysts, accelerators, initiators, hardeners, or fillers.

    According to Prime (1997:1662), catalysts, accelerators, and initiators catalyze

    the cure process, and in doing so reduce the processing times. Hardeners are

    incorporated in the network structure of the polymer chains; thereby having an

  • 8/2/2019 Cure Thesis

    20/76

    9

    effect on the reaction rate, as well as the ultimate properties of the material, such

    as the network structure, glass transition temperature, and dynamic mechanical

    properties. Fillers are generally incorporated to provide or enhance physical

    properties such as modulus, thermal expansion and dimensional stability.

    2.2.1.1 Functionality

    According to Odian (1991:40, 106-108), polymerization of bi-functional

    monomers (f=2) will form linear polymers, whereas polymerization of monomers

    with more than two functional groups per molecule (f>2) will result in the polymer

    being cross-linked.

    With certain monomers, cross-linking will take place, resulting in a network

    structure in which one or more branches from one molecule become attached to

    surrounding molecules.

    2.2.1.2 Molecular Mass

    Odian (1991:19-20) emphasises the importance of the molecular mass of a

    polymer in its synthesis and application. Also that, the useful mechanical

    properties uniquely associated with polymeric materials are a consequence of

    their high molecular weight. It can generally be said that the use of polymer

    material in practical applications require high molecular weights to obtain high

    strengths. It should be noted that polymer chains with strong intermolecular

    forces develop sufficient strength to be useful at lower molecular weights than

    polymers having weaker intermolecular forces (with higher molecular weights).

    There is a significant dependency of the properties (other than strength) of a

    polymer on its molecular weight, albeit different quantities for different properties.

    Optimum or maximum values of properties will be reached at different molecular

    weights and can decrease with a further increase.

    It is therefore necessary for the polymerization process to result in a compromise

    molecular weight, where sufficient strength is obtained without sacrificing other

  • 8/2/2019 Cure Thesis

    21/76

    10

    properties. The control of the molecular weight of a cross-linking process is

    essential for the practical application.

    2.2.2 Competition between Cure and Degradation.

    Prime (1997:1505) explains that the cure process can often become complicated

    as a result of competition between degradation and chemical cross-linking

    reactions. This can often occur as a result of the high cure temperatures required

    to achieve high glass transition temperatures of high performance systems. Yuan

    and Mazeika (1991) as reported by Prime (1997:1431), found that isothermal

    degradation occurs in three stages; during the first stage, there is a decrease in

    modulus and glass transition temperature (Tg), a probable result due to scission

    in the epoxy network. The second stage marks the formation of char, and duringthe third stage char stabilizes, and shows an increase in modulus and Tg. The

    increase may be caused by an increase in the cross-linking density of the

    degraded network. The char region can be seen on Figure 2.1.

    2.3 Gel point

    Gelation is the first noticeable change that is encountered during the cure

    process. Prime (1997:1496) describes gelation as an irreversible change from a

    liquid to a rubber, and that this change marks the start of the formation of the

    cross linked network. Malkin and Kulichikhin (1996:277) used gelation to denote

    the process of fluidity loss, which is a result of the formation of a network of

    chemical bonds.

    Prime further defines gelation on a molecular basis, but states that the

    phenomenon is measured as a macroscopic phenomenon. A distinction between

    molecular gelation and macroscopic gelation may be drawn; macroscopic

    gelation is the consequence of the microscopic gelation. Gelation will be

    discussed on a microscopic and macroscopic level.

  • 8/2/2019 Cure Thesis

    22/76

    11

    2.3.1 Microscopic gelation

    Microscopic gelation marks the start of the formation of the three dimensional

    network structures that is typical of thermosetting systems, and will have a direct

    influence on the final structure of the cured resin. It is vital to understand the

    influence of the occurring changes, and their effect on the final properties of the

    cured resin. Microscopic analysis of gelation focuses on molecular and chemical

    changes that take place during polymerization.

    Ellis (1993:4&72-74), Malkin and Kulichikhin (1996:277), and Prime (1997:1380-

    1383, 1496) are in agreement that the cure of thermoset material on a

    microscopic level is the formation and linear growth of molecular chains. These

    linear chains form branches, cross-links with other branched chains, and result in

    the formation of a rigid three dimensional molecular network, all due to the

    chemical reaction between the epoxy and hardener reactive groups. Malkin and

    Kulichikhin (1996:277) add that the microcrystals formed act as stable points of

    the three-dimensional network. These strong physical bonds will cause fluidity

    loss.

    Flory (as quoted by Gillham (1989:804)) and Winter (as quoted by Prime

    (1997:1381)) defines the microscopic gel point of a cross-linked system, as the

    point where the weight average molecular weight of the network of chains

    approaches infinity.

    Molecular gelation is therefore a point that may be measured by the initial

    identification of an insoluble gel.

    Flory and Miller et al. (taken from Prime (1997:1382)) are in agreement that

    molecular gelation generally occur during a well-defined and calculable stage in

    the course of the chemical reaction. The only requirements are that the reaction

    mechanism is not dependent on temperature and that there are no non-cross-

    linking side reactions. Molecular gelation is dependent on the functionality,

  • 8/2/2019 Cure Thesis

    23/76

    12

    reactivity, and stoichiometry of the reactants, therefore the theoretical gel point

    can be calculated if the chemistry of the reactants is known.

    2.3.2 Macroscopic gelation

    Ellis (1993:5&72) points out that the first critical feature of a curing system is

    gelation and can be macroscopically described as the point where the resins

    viscosity rapidly approaches infinity. This happens as a consequence of

    microscopic gelation, which is also defined as such by Prime. Gan et al

    (1989:804) explain that the transformation is from a viscous fluid to a state of

    infinite viscosity, or a change from a fluid liquid to a rubbery state. Gillham

    (1990:1) adds that there is a transformation of liquids with low molecular weight

    to an amorphous solid with a high molecular weight by means of chemical

    reactions.

    This irreversible transformation from a fluid to a rubber means that beyond the

    gel point a thermosetting resin loses all ability to flow. This marks a very

    important stage, especially for manufacturing and processing.

    2.3.3 Gelation Study

    The measurements of isothermal times to gelation are important to cure studies

    and will be used to form the basis of a cure diagram and is discussed in later

    chapters of the study.

    The progressive macroscopic change from the liquid state to the rubbery state

    and/or glassy state which occurs during the setting of reactive systems has been

    described on both a molecular and macroscopic level. Prime (1997:1496)

    explains that the detection of an insoluble material, i.e. a cross-linked material

    (gel) that forms from a reacting soluble material (sol), may be taken as the

    starting point of measure of gelation.

  • 8/2/2019 Cure Thesis

    24/76

    13

    According to Winter, Holly et al., Feve, Boiteux et al. (as quoted by Prime

    (1997:1496)), the point where the mechanical loss tangent becomes independent

    of frequency can be used as an alternative measurement starting point.

    Molecular gelation is taken as the point where the loss modulus (G') and storage

    modulus (G'') cross each other, i.e., tan =1. The approximation of gelation

    macroscopically is based on rheological and mechanical changes that occur

    during the transition from liquid to rubber. The rheological aspect accounts for the

    changing viscosity, while the mechanical aspect accounts for the evidence of an

    equilibrium modulus.

    Initial work by Gan et al. (1989) reported on the use of the torsional braid analysis

    (TBA) technique to monitor and study gelation and vitrification of a setting

    system. TBA is a resonant instrument, which according to Gallagher (1997:137),

    was widely used in the past. The detailed mechanics of the system will not be

    discussed here, but it is important to understand that the frequency of oscillation

    is related to the modulus, and that the logarithmic decrement (ratio of the

    amplitude of successive cycles) is related to tan .

    Nez et al. (2000) proposed three methods to experimentally determine the time

    at which gelation occurs (tgel), namely; solubility test, gel-timer and dynamic

    mechanical analysis (DMA). The solubility test, as described by Hagnaver (1983),

    is used to determine the time taken for the setting material to reach a fibriform

    structure in tetrahydrofurane. The DMA measures the dynamic mechanical

    properties of the epoxy.

    2.3.3.1 Curing analysis: Rheological analysis

    Malkin and Kulichikhin (1996:142), Lange et al. (1996), Kulichikin (1997), Halsz

    et al. (2001), Vorster (1995), Vorster et al. (2004), and Sun (2002) used

    rheology, the study of the change of viscosity, as a mean to determine the point

    of gel formation. Ellis (193:87-89) explains that even though the initial rate of

    reaction may be high, the molecular growth is slow. There is a relationship

  • 8/2/2019 Cure Thesis

    25/76

    14

    between the degree of cure and viscosity; this allows for cure monitoring by

    measuring the change in shear viscosity of the setting system. The shear

    viscosity of a resin increases during cure as a result of gelation and/or the onset

    of vitrification.

    According to Halsz et al. (2001:898), rheological analysis of the viscosity of

    curing systems may be carried out using a small strain (5%) at different

    frequencies and temperatures. The components of complex modulus, complex

    viscosity and tan should be measured and recorded against frequency and

    therefore as a function of time, as depicted in work performed by Halsz et al.

    (2001:898). It is interesting to note that prior to gelation the behaviour of the liquid

    resin can be characterized by the zero shear viscosity.

    Vorsteret al. (2004) quoted Flory (1941 & 1959) and Stockmayer (1943 & 1944)

    who predicted theoretical gel point values using the following equation;

    ( )( )

    2

    1

    11

    1

    =

    gfRg [2.1]

    where Ris the molar ratio between the epoxy and hardener, fis the functionality

    of epoxy and gthe functionality of the hardener.

    From the data obtained from experimental rheological measurements the

    gelation times (tg) can be determined using three methods:

    Extrapolating the inverse of zero viscosity

    Equilibrium (intersection) of G and G curves of the dynamic modulus

    Intersection of the tan curves.

    Halsz et al. (2001:899) obtained and compared gel times according to the three

    methods discussed above. The time differences were found to be insignificant,

    thereby permitting single test validity.

  • 8/2/2019 Cure Thesis

    26/76

    15

    Times to gelation (Log time in minutes) as described by Gan et al. (1989:807)

    can be plotted against isothermal cure temperatures to obtain time temperature

    transformation (TTT) diagrams with gelation contours. Gelation contours on a

    TTT diagram shows the relationship between tgeland Tcure. The relationship that

    exists between tgel(Ln time in minutes) to macroscopic gelation versus the

    reciprocal temperature (1/K) is linear. The slope of the linear relationship

    represents the apparent activation energy (Ea) for the system.

    2.4 Vitrification

    Wunderlich (1997:460) defines vitrification of a setting system as the transition

    from a viscous liquid or rubber to an amorphous solid or glass. Gan et al.

    (1989:803) describes vitrification as the point where the system solidifies. This

    phase transition is a consequence of the cure reactions, and as Prime

    (1997:1383 & 1569) explains, vitrification can occur at any stage during

    polymerization. Glass formation (vitrification) is as a result of the glass transition

    temperature (Tg) increasing with conversion. It is accepted that vitrification occurs

    at times to vitrification, tvit, when Tgrises to the cure temperature (Tcure).

    The measurements of the times to vitrification, similar to times to gelation, are

    important and will be used in cure studies. It forms the basis of cure diagrams.

    With reference to Figure 2.1, Malkin and Kulichikhin (1996:120), and Prime

    (1997:1384) explain that in the glassy state beyond the vitrification line, the cure

    reaction rate significantly drops. Wisanrakkit and Gillham (as quoted by Nezet

    al. (2001:3581), Adabbo and Williams, Dillman and Seferis, and Wisanrakkit and

    Gillham (taken from Prime (1997:1623)) explain that isothermal curing of

    thermosets occur in two distinct stages; the first stage reaction mechanism is

    kinetically controlled, but changes to diffusion controlled soon after vitrification.

    There is an increase in the diffusion time scale, whereas segmental mobility, and

    reacting groups decrease.

  • 8/2/2019 Cure Thesis

    27/76

    16

    According to Prime (1997:179), gel conversion is dependent on functionality and

    stoichiometry, whereas the glass transition is dependent on the frequency of

    measurements. Therefore indicating the dependence on the dynamic mechanical

    technique used for monitoring it.

    2.4.1 Cure temperature

    According to Prime (1997:1381), most thermosetting formulations require heat or

    irradiation to cure. Energy in the form of heat can be supplied to the reaction and

    will be elevating the cure temperature of the system. The cure temperature (Tcure)

    is a processing parameter and is the temperature at which the polymerisation or

    cure process takes place.

    2.4.2 Glass transition temperature

    Scherer, Rekhson and Elliott (from Ellis, 1993:81) point out that glass-forming

    systems exhibit a glass transition phenomenon. Prime (1997:1401) further

    defines the point as a transition from a supercooled liquid or rubbery state to a

    metastable glassy state.

    According to Wunderlich (1997:226 & 380), there is a definite link between the

    liquid state and glassy state of a thermosetting resin. The glass transition

    temperature (Tg) is the main characteristic temperature of the two states, and is

    the process where the molecular orientation of the liquid state is frozen in

    position as a glass. Because there is no change in molecular order at Tg, the

    macroconformation of the molecules in the glass state correspond to that of the

    liquid.

    Odian (1991:24 & 29) defines the glass transition temperature, Tg, of a polymeric

    system, as the temperature where the amorphous or unordered regions of a

    polymer take on the characteristic properties of the glassy state, typically

    brittleness, stiffness, and rigidity that is associated with the solid phase.

  • 8/2/2019 Cure Thesis

    28/76

    17

    According to Gan et al. (1989:803), Gillham (1990, 1993), and Wang (1992), the

    sensitivity and one-to-one relationship between Tg and the conversion for

    thermosetting resins, implies that Tg can be used as an index to measure

    conversion during cure. The glass transition temperature is a useful, sensitive,

    easily measured macroscopic parameter for monitoring the cure process of

    thermosetting systems.

    Prime (1997:1506&1414) points out that softening of a thermoset is affiliated with

    Tg, defines the upper use temperature of the system. This information is critical

    when designing and manufacturing composite structures to be used at elevated

    temperatures.

    2.4.3 Vitrification study

    As stated before, onset of vitrification of a thermosetting system is when there is

    a change from a liquid to a glassy solid. This transition is further defined as the

    point where Tgrises to Tcure. It is therefore important to be able to determine Tg.

    Ellis (1993:72) points out that when Tcure is too low, vitrification may occur before

    gelation and inhibit further reaction.

    2.4.3.1 Curing analysis: Differential scanning calorimetry

    Differential scanning calorimetry (DSC) has been used in work by various groups

    such as Gillham (1990 & 1993), Nez et al. (2001) and Mafi et al. (2005) to

    determine the glass transition temperature (Tg) of setting systems. According to

    Turi (1981) and Wunderlich (1990) as quoted by Prime (1997:1388-1390), DSC

    forms part of a family of measuring techniques that record the response of a

    material to temperature scans. This is achieved by measuring and recording the

    heat flow into a material (endothermic) or out of a material (exothermic). DSC

    can therefore be used to determine Tg for thermosets; by assuming that the

    measured heat flow (dH/dt) is proportional to the reaction rate (d/dt), or after

    integration, the total heat detected is identical to the heat evolved by the curing

  • 8/2/2019 Cure Thesis

    29/76

    18

    reaction. Other characteristics that can be determined through the use of DSC

    are cure, and physical ageing of systems.

    This technique monitors the setting system during the change from a liquid to

    rubbery state and finally glass state. According to Gan et al(1989:804) Tgcan be

    measured in three types of experiments:-

    1 Times to vitrification are obtained when Tgrises to Tcure

    2 The progress of setting before vitrification at Tcure can be obtained

    from measurements ofTgobtained during intermitted cooling

    3 Temperature scans first to a temperature below and then to

    temperatures above Tcure, give Tg values which reflect the state of

    cure obtained at Tcure, and the stability of the material to

    temperatures higher than Tcure.

    Tg values for samples cured at Tcure fortcure determined from DSC scans can be

    used to monitor cure conversion simply by plotting Tg against Tcure for isothermal

    heating times. The relationship is linear at lower isothermal cure temperatures.

    All Tcure

    versus Tg

    curves features a maximum. Gan et al. (1989:807) explain that

    this inflection is due to the competition between cure (which raises Tg) and

    thermal degradation (which decreases Tg).

    2.5 Characterization of the cure process

    Gillham (taken from Prime (1997:1568)) provides a useful frameworks for the

    understanding and conceptualization of the changes that occur during the curing

    process of a thermosetting resin. This is through the development of the

    isothermal time-temperature-transformation (TTT) cure diagram as depicted in

    Figure 2.1. The diagram depicts the various physical states of a thermoset (liquid,

    rubber, and glass) in relation to the two critical phenomena: gelation and

    vitrification. Sun (2002:91) proposes that the characteristics of the cure process

    can be understood by combining and describing the rheological analysis and

  • 8/2/2019 Cure Thesis

    30/76

    19

    DSC thermal analysis by a time-temperature-transformation (TTT) isothermal

    cure diagram.

    Construction of a TTT isothermal cure diagram requires that the relationship

    between degree of cure and cure time at isothermal cure temperatures be

    established.

    The relationship between gel times (tgel), defined as the elapsed time to achieve

    equal storage and loss moduli, and isothermal cure temperature is represented

    by the gelation contour. Earlier discussions verify that Tg is a function of degree

    of cure and dependent on cure temperature. The relationship between Tgand the

    isothermal cure time can be determined. This allows that the time taken for

    isothermal cure, or vitrification (tvit), when Tg rises to the cure temperature (Tcure),

    can be determined. The relationship between tvit and Tcure is represented by the

    vitrification curve. By plotting the gelation and vitrification curves on the same

    graph, the isothermal TTT cure diagram for epoxy resin system is constructed.

    2.6 Mechanical properties

    Proper application of epoxy resins requires that the material be a rigid solid at the

    use temperature Tuse

    . The use temperature in all these applications is below the

    glass transition temperature of the cured resin, TuseTg.

    Work by various authors (Nielsen (1969:69-103), Ellis (1993:159) Paul

    (1989:187), Gillham (1993), Lange et al. (1996), Morgan (1997:2103), Fabrice

    and Redford (2002:340), Ahamed et al. (2004), and Mafi et al. (2005)) confirm

    the influence of molecular weight and cross-linking density on the final

    mechanical properties of thermosetting systems. Ellis explains that epoxies with

    higher cross-linking densities have higher Tgs, which often succumbs to brittle

    failures, whereas epoxies with low cross-linking densities exhibit structures with

    more mobility that allows for greater plastic deformation. Higher levels of

    deformation often result in higher fracture toughness. Misev (1991:196-197)

    substantiates that the mechanical properties are directly proportional to the

  • 8/2/2019 Cure Thesis

    31/76

    20

    molecular weight. But only to a point, as the molecular weight is dependent on

    the resin and cross-linker stoichiometry.

    Bell as well as Misra et al. (as reported by Ellis (1993:159)) found significant

    increases in impact strength with an increase in molecular mass between cross-

    links. A higher molecular masss between cross-links denotes a lower cross-link

    density.

    2.6.1 Mechanical properties: Molecular influences

    2.6.1.1 Molecular weight

    According to Misev (1991:176), the molecular weight of a thermoset influences

    the following mechanical properties; tensile strength, impact resistance and melt

    viscosity. This was confirmed through studies performed by Nielsen, Matsuoka,

    Walsh and Termonia (taken from Nicholson et al. (date:1)), the molecular weight

    distribution effects fracture toughness, and impact strength of the polymer.

    The critically molecular weight dependent property is the glass transition

    temperature (Tg), which represents the temperature region where there is a

    phase change, either form a glass to a rubber, or from a rubber to a glass. Tg

    generally increases with the extend of cure, which denotes molecular growth, and

    confirms the relationship.

    Good mechanical properties after curing requires that the molecular weight be

    high. There is however a constraint; thermosetting systems with higher molecular

    weights consequently have high viscosities, which causes production problems.

    2.6.1.2 Cross-linking density

    The cross-linking density is representative of the number of bonds or cross-links

    that has formed as result of the chemical reaction between the constituents

    reagents during the cure. Ellis (1993:159), Nielsen (1969:69-103), Paul

    (1989:187) and Morgan (1997:2103) are in agreement that the fundamental

  • 8/2/2019 Cure Thesis

    32/76

    21

    physical and mechanical properties of epoxy resins are determined by the nature

    and density of its cross-linked network. There is influence on Tg of the polymer,

    the ability to undergo plastic deformation, therefore its toughness characteristics

    at a given temperature. Gillham (1993) explains that the maximum cross-linking

    density for a fully-cured material will yield the minimum internal stress, and

    maximum modulus.

    There is of course a limit to the cross-linking density, explained by Reghunadhan

    Nair (2004:406) and Mimura and Ito (2002:7559), as epoxies with high cross-

    linking densities tend to have high Tgs, but often exhibit brittle structures.

    Almeida and Monteiro (1996:330) add that higher cross-linking densities

    generally result in stiffer structures with higher ultimate strengths.

    Excessive cross-linking reactions result in a rigid, brittle network structure with a

    diminished load bearing capacity.

    2.6.2 Mechanical property analysis techniques

    2.6.2.1 Dynamic mechanical analysis

    Methods such as torsional braid analysis (TBA), dynamic mechanical analysis

    (DMA), and dynamic thermal mechanical analysis (DMTA) are methods of

    measuring changes in complex modulus, compliance, and viscosity of samples

    during oscillatory deformation of material samples. The different modes of

    deformation that a sample material can be subjected to are as follows; flexure,

    tensile, shear, bending, and compression deformations. Principal applications of

    DMA for thermosetting material is the measurement of Tg, analysis of cross-

    linking, mechanical properties developed during the curing process and

    characterizing cured or partially cured materials.

    Prime (1997:1464-1465) states that Tg for thermosetting materials can be defined

    to occur at the maximum value of the loss modulus (E'), the loss compliance, or

  • 8/2/2019 Cure Thesis

    33/76

    22

    tan . Mechanical properties such as storage and loss moduli or compliances for

    samples during or after cure can be measured.

    The dynamic mechanical modulus, E*, is defined as the ratio of the stress, *, to

    the strain, *, in the material being tested. The strain is a response to the stress

    applied, or the stress is a response to the applied strain. E*is a complex quantity

    and is defined as:

    '''**

    *

    iEEE +==

    [2.2]

    '

    ''tan

    E

    E= [2.3]

    where E' is the storage modulus, a measure of the stress stored in the sample

    as mechanical energy; E''is the loss modulus, a measure of stress dissipated as

    heat; tan is the phase lag between the stress and strain.

    2.6.2.1.1 Instrumentation

    o Torsional braid analysis (TBA)

    TBA is a dynamic mechanical technique where the specimen for a freelyoscillating torsion pendulum experiment consists of an inert substrate or braid

    coated with the reactive system. The changes from the liquid or rubbery state to

    the glassy state can be monitored during cure. Gan et al. (1989) and Jiawu et

    al.(2000) used the TBA to study the cure behaviour and measure macroscopic

    gelation times tgel and glass transition measurements such as Tg and times to

    vitrification (tvit) to characterize reactive coatings.

    o Dynamic mechanical analysis (DMA)

    DMA is a thermal analysis system module that is designed for four modes of

    oscillation: fixed frequency oscillation, resonant frequency oscillation, stress

    relaxation, and creep. Direct measurements of frequency, temperature, time,

    stress, stain, and phase angles are taken during scans. Measurements can be

  • 8/2/2019 Cure Thesis

    34/76

    23

    made in the following modes; three-point bending, dual cantilever, fiber

    extension, film extension, and parallel plate.

    Yasmin and Daniel (2004:8213) determined thermomechanical properties such

    as storage modulus (E), loss modulus (E), damping factor tan, and Tg of

    samples through DMA scans. Park et al. (2004) used DMA purely to determine

    the temperature dependence of the loss factor tan for samples modified through

    synthesis of epoxidized soybean oil. Ahamed et al. (2004) used DMTA to study

    the effect of stoichiometry on the dynamic mechanical properties of resins

    systems. Cain et al. (2004:3) performed DMA in dual-cantilever mode on

    samples.

    o Dynamic mechanical thermal analysis (DMTA)

    The DMTA is a fixed frequency thermal analysis module that measures time,

    temperature, frequency, sample, and suspension stiffness (in-phase and out-of-

    phase) directly. Three measurement modes are possible: bending mode, single-

    cantilever (and double cantilever), and shear sandwich geometry.

    Dynamic mechanical thermal analysis (DMTA) can be used to measure the

    complex modulus, compliance, and viscosity of materials as functions of

    temperature in several different modes of oscillatory deformation.

    Mafi et al. (2005) and Remiro et al. (2001) used DMTA to gauge the effect of

    curing conditions such as time and temperature on dynamic mechanical

    behaviour (adhesion, viscoelasticity, and stiffness) of a thermoset system. Laza

    et al. (1998) used DMTA to study the effect of amine concentrations on the

    dynamic mechanical properties of resins systems.

    2.6.2.1.2 DMA and DMTA analysis

    DMA according to Prime (1997:1581-1597), is commonly used in characterization

    of polymeric materials. This is due to the viscoelastic behaviour of the material

    which changes with time. A reacting system therefore requires a description of

    the kinetics of the system and the effect of the changing system on the

    viscoelastic model. The non-isothermal dynamic mechanical characterization of

    cured or partially cured systems will include effects of additional cure. The

  • 8/2/2019 Cure Thesis

    35/76

    24

    viscoelastic model is described by the dynamic mechanical behaviour shown

    when the measured storage (E), and loss (E) moduli are plotted against

    temperature.

    2.6.2.2 Indentation hardness

    The indentation hardness of cured epoxy samples can be measured by means of

    a Barcol Impressor.

    Pharr (1998) reviewed and discussed ultra-low load indentation (or

    nanoindentation) techniques commonly used for measuring mechanical

    properties of thin films and small volumes of material. The different indentation

    techniques vary and can be classified according to the type of indenter used in

    the measurement. Although the above mentioned work deals with small

    quantities of material on a sub-micron scale, it still bears justice to this study as

    mechanical properties can be determined by analysis of the indentation load-

    displacement data.

    According to Reghunadhan Nair (2004:89), enhanced modulus, compressive

    strength and hardness are desirable improvements in epoxy resins. Good

    mechanical performance offer immediate solutions for challenging problems.

    Zhang et al. (2004) investigated the enhancement of wear resistance of epoxies

    and found that the friction and wear behaviour can be improved through

    enhancement of hardness, stiffness and compressive strength.

  • 8/2/2019 Cure Thesis

    36/76

    25

    CHAPTER 3

    3.0 Research Methodology

    3.1 Materials used and their preparation

    The base resin used in this study was prepared from the reaction between

    epichlorohydrin and bisphenol-A (Epon Resin 826 (ARH)) manufactured by

    Resolution Performance Products, and supplied by Plastomax. The curing agent

    used in this study was an aliphatic amine (Ancamine) also from Plastomax. Data

    sheets representing the materials are presented in Annexure A.

    The components of the system, the epoxy resin and hardener (curing agent),

    were carefully weighed on an electronic scale and homogeneously mixed to a

    ratio of 100:25 by mass (as per instruction from the supplier).

    3.2 Gelation Study

    A Rheometric Dynamic Stress Rheometer (RS-500) (Figure 3.1) equipped with

    Orchestrator software was used under oscillatory dynamic mode to obtain the

    experimental data reported in this study.

    A parallel 25 mm diameter plate measuring head with a 0,2 mm gap was

    employed. The measurements were carried out by using a small strain (5%) at

    four frequency values (50, 15, 10 and 1 rad/s) and nine temperatures for each

    frequency (30, 40, 50, 60, 70, 80, 90, 100 and 110C). The components of

    complex modulus, complex viscosity and tan were measured and plotted

    against frequency and therefore as a function of time.

    The gel point was taken as the crossover value of the two components of the

    complex modulus curves, namely the storage and loss modulus. (See Figure

    3.2). Gelation occurs at time tgel=196 s, indicated by the crossover of the

    components of complex modulus.

  • 8/2/2019 Cure Thesis

    37/76

    26

    Figure 3.1. Rheometric Dynamic Stress Rheometer (RS-500)

    Figure 3.2. Typical plot of the loss and storage moduli, tan and

    complex viscosity * against time for Rheological measurements

    G'/G ' Crossover Point: 196

  • 8/2/2019 Cure Thesis

    38/76

    27

    The average time for the system to reaction to reach gelation, time Tgel,

    was calculated from the data and plotted in Figure 3.2. This shows the

    relationship between the time to gelation (tgel) and curing temperature

    (Tcure) of the system.

    3.3 Vitrification Study

    3.3.1 Sample preparation

    Small amounts of the initial reactive mixture as per section 3.1 (approximately 10

    mg) were transferred into small DSC aluminium pans. These samples were

    grouped in batches of 10 and sealed in plastic zip lock bags. The plastic bags

    were stored in a freezer, at approximately -15C to prevent any cure taking place.

    Sample pans were removed from the sealed plastic bags, and were allowed at

    least 20 minutes to reach room temperature before placing in an oven for

    isothermal curing at 60, 80, 100,120 and 140C for periods up to 4320 minutes.

    A Perkin Elmer differential scanning calorimeter, Model DSC 7 (see Figure 3.4),

    equipped with Pyris software was employed to determine the glass transition

    temperatures. The heating profile consisted of a heating run from 25C to 150C

    at 10C/min. The test was carried out in a nitrogen atmosphere with a flow rate of

    20 5 ml/min. The parameters measured were the glass transition temperature

    (Tg) and residual exotherm (Hr) of the reaction after the material had been

    subjected to isothermal curing at the different curing temperatures.

    Tgappears as an endothermic shift over a temperature interval in the DSC scan

    as shown in Figure 3.3. In this study, Tg was taken as the mid-point of the step-

    transition, while Hr of the remaining reaction appears as an exothermic peak in

    the temperature range of the rubbery state of the material.

  • 8/2/2019 Cure Thesis

    39/76

    28

    Figure 3.3. Typical plot of a DSC scan of endothermic heat flow (mW) against temperature(C).

    Figure 3.4 Perkin-Elmer Differential Scanning Calorimeter (DSC-7)

    3.4 Dynamic Mechanical Thermal Analysis

    3.4.1 Sample preparation

    The reactive mixture (prepared as described in section 3.1) was poured into

    specially created flat moulds made from impression material (Zetalabor Platinum

    Laboratory High Precision Addition Silicone. Manufactured by Zhermack).

    These samples were stored in a freezer, at approximately -15C to prevent any

    cure taking place.

  • 8/2/2019 Cure Thesis

    40/76

    29

    Samples were removed from the freezer, and were allowed at least 20 minutes to

    reach room temperature before moving to an oven for isothermal curing.

    The cured samples were cut to approximately 5mm x 25mm x 1.5 mm, carefully

    measured and mounted in the DMTA.

    A Rheometric Scientific Solid Analyzer RSA II (DMTA), equipped with

    Orchestrator software was used to obtain the experimental data (see Figure 3.6).

    A Dynamic Temperature Ramp using the dual cantilever mode was employed

    and measurements of the storage modulus (E'), loss modulus (E) and loss

    tangent (tan ) of fully cured samples were carried out using a small strain (0.01

    %) at a frequency of=6.28 rad/s.

    The cured samples were subjected to temperature scans from 40 to 160C at a

    rate of 10C/min.

    DMTA analyses were performed on samples cured at three different

    temperatures (70, 90 and 110C) for three different periods (30, 60 and 120

    minutes). A typical DMTA scan is shown in Figure 3.5

  • 8/2/2019 Cure Thesis

    41/76

    30

    Figure 3.5. Typical plot of the storage modulus (E') and tan against temperature. (Taken from

    Laza et al. (1998:43))

    Figure 3.6. Rheometric Scientific Solid Analyzer RSA II (DMTA) in dual cantilever mode

  • 8/2/2019 Cure Thesis

    42/76

    31

    3.5 Indentation hardness

    3.5.1 Sample preparation

    The samples were prepared as described in Section 3.4 under Dynamic

    Mechanical Thermal Analysis

    3.5.2 Experimental procedure

    A Hand-Held Portable Hardness Tester (Barcol Impressor, Model No. 934-1) was

    used to obtain the experimental data (see Figure 3.7). The indentor consisted of

    a hardened steel truncated cone with an angle of 26 with a flat tip of 0.157 mm

    in diameter.

    The test method used to determine the indentation hardness was performed in

    accordance with Standard Test Method for Indentation Hardness of Rigid Plastics

    by Means of a Barcol Impressor, ASTM Designation: D 2583 87.

    Barcol indentation analyses were performed on samples cured at three different

    temperatures (70, 90 and 110C) for three different times (30, 60 and 120

    minutes). It must be noted that the impression hardness tests were carried out on

    samples at room temperature.

    Figure 3.7. The Impressor: A Hand-Held Portable Hardness Tester(Barcol Impressor, Model No. 934-1)

  • 8/2/2019 Cure Thesis

    43/76

    32

    CHAPTER 4

    4.0 Results and discussion

    4.1 Introduction

    The results obtained from the experimental work will be dealt with individually in

    this chapter. Results from gelation and glass transition temperature studies are

    used to monitor the extent of cure of the thermoset system, and forms the basis

    of characterizing the material. The results will be summarized in the form of an

    isothermal time-temperature-transformation (TTT) diagram.

    The results obtained from dynamic mechanical thermal analysis (DMTA) and

    hardness tests will be used to illustrate the relationship between mechanical

    properties and the molecular structure of the cured material as the molecular

    structure of the system is dependent on curing conditions.

    4.2 GelationStudy

    The aim of the gelation study is to determine at what point during the curing

    process of the thermosetting system gelation (Tgel) occurs. The experimental

    results obtained from parallel plate rheology tests, as per Section 3.1., were

    used to obtain crossover values of the storage and loss modulus curves as a

    function of time. A derived summary of the average gel times (Log Time) for

    tests performed at the four frequencies (50, 15, 10 and1 rad/s) and seven cure

    temperatures (30, 40, 50, 60, 70, 80, 90, 100 and 110C) are shown in Table

    4.1. Times to gelation for samples tested at low curing temperatures (30, 40

    and 50C) are extrapolated from storage and loss modulus curves, as the

    system solidified prior to reaching gelation. We can classify these as having

    vitrified before the curing system could form a gel. The greyed out cells

    represent measurements that could not be recorded due to premature

    solidification of the resin system. The values of pre-gelation and premature

    solidification will not form part of the calculations.

  • 8/2/2019 Cure Thesis

    44/76

    33

    Table 4.1.Average gel times (Log Time) for different frequencies

    at the respective cure temperatures

    Frequency (rad/s)

    Gelation

    Times

    (Log t) 50.00 15.00 10.00 1.00

    Average

    (s)

    Average

    (min)

    Log

    Time

    (min)

    30 8326.95 8273.35 8973.95 10009.80 8896.01 148.27 2.17

    40 4628.70 4830.05 4597.20 5722.45 4944.60 82.41 1.92

    50 3007.05 3107.15 3273.30 3615.80 3250.83 54.18 1.73

    60 1504.70 1677.15 1727.45 1661.25 1642.64 27.38 1.44

    70 1020.21 1126.38 893.51 1008.60 1012.18 16.87 1.23

    80 663.49 688.20 650.42 680.04 670.54 11.18 1.05

    90 443.39 435.13 416.90 409.87 358.65 5.98 0.78

    100 344.07 297.42 312.99 318.16 5.30 0.72

    110 224.10 243.00 289.00 252.03 4.20 0.62

    Further analysis of data from Table 4.1 presented in Figure 4.1 shows the

    decrease in gelation times (Tg) for samples cured at higher cure temperatures

    (Tcure). This indicates that there is a definite relationship between the time taken

    for the polymer system to react and form a three dimensional network structure

    and the temperature at which the system is cured. It is obvious that the time

    taken for the system to form a gel decreases as the cure temperature

    increases. The oscillation frequency at which the tests were conducted does

    not have a real effect on gel times.

  • 8/2/2019 Cure Thesis

    45/76

    34

    20

    30

    40

    50

    60

    70

    80

    90

    100

    110

    120

    0.40 0.60 0.80 1.00 1.20 1.40 1.60

    Log Time (min)

    Tcure

    (C)

    Figure 4.1. Time-temperature-transformation (TTT) diagram: Gelation contour.

    4.3 Vitrification Study

    The aim of the vitrification study is to determine at which point the

    thermosetting system vitrifies during the curing process. Vitrification is a

    transition from a viscous liquid or rubber to a glass and according to Gan et al.

    (1989:803), vitrification is the point where the system solidifies. According to

    Prime (1997:1383 &1569), vitrification can occur at any stage during

    polymerization, either as a consequence of solvent evaporation or chemical

    reaction. It is therefore important to be able to track and monitor the cure

    reaction and define the phase changes that occur during cure and in doing so,

    demonstrate the conditions where vitrification takes place.

    Work by Gan et al. (1989:803), Gillham (1990, 1993), and Wang (1992),

    illustrated that there is a sensitive one-to-one relationship between the glass

    transition temperature, Tg, and the conversion in thermosetting resins. This

    implies that Tg can be used as an index for measuring conversion during cure.

    The glass transition temperature is a useful, sensitive, easily measured

    macroscopic parameter for monitoring the cure process of thermosetting

    systems. Measurements taken on samples cured at different cure

    temperatures, Tcure, for different periods, tcure, were measured by differential

  • 8/2/2019 Cure Thesis

    46/76

  • 8/2/2019 Cure Thesis

    47/76

    36

    4.3.1.1 Tg versus Tcure

    Values ofTg determined by DSC and the corresponding Tcure values for cure

    times tcure (1 to 72 h) are plotted in Figure 4.2. The contours show the

    relationship between Tg, and Tcure.

    Figure 4.2.Tg (C) vs. Tcure (C) for different isothermal heating times (1 hour through to72 hours). The reference line ofTg= Tcure is also included.

    The reference lines included are ofTcure = Tg, and Tg= 107.6C. The value for

    Tg is determined in Section 4.3.1.2.

    By examining Figure 4.2, especially Tgvs. Tcure for the tcure=10 hours contour,

    we note a definite increase in glass transition temperature with an increase in

    cure temperature. It is important to note that samples cured at a constant Tcure

    for increased time increments, show continuous increases in Tg values.Keeping in mind that Tg is used to monitor cure, we can see from the positive

    contour slopes ofTgvs. Tcure at lower curing temperatures, that the cure is not

    complete. Contours at higher curing temperatures, for longer curing times,

    exhibit a flexure point in the slope of the plot and a definite maximum is

    observed. This is a result of the competition between cure (which increases Tg)

    Tg=107,6C

    Tg=Tcure

  • 8/2/2019 Cure Thesis

    48/76

    37

    and thermal degradation (which decreases Tg) and is indicative of a

    deteriorating molecular structure. The change is possibly due to de-vitrification,

    and molecular breakdown as a result of thermal influences on the molecular

    structures.

    According to Gan et al. (1989:807), a system that exhibits contrasting values of

    Tg>Tcure forTcure >Tg, even for prolonged cure times (tcure = 72 hours) indicates

    that the system is highly reactive.

    4.3.1.2 Tgversus tcure

    The variation of Tg with log time at different isothermal temperatures is

    illustrated in Figure 4.3. It shows that samples cured at constant temperaturesfor increasing cure times exhibit a general increase in Tgwhich is a result of the

    curing reaction taking place and, is in accordance with that reported by Gan et

    al. (1989:808).

    Figure 4.3.Tg (C) vs. Log Time (min) for different isothermal temperatures (Tcure).

  • 8/2/2019 Cure Thesis

    49/76

  • 8/2/2019 Cure Thesis

    50/76

    39

    4.3.1.3 Iso-Tgcontours

    The data derived from Figure 4.3 was used to develop iso-Tg contours in the

    format of a time-temperature-transformation (TTT) diagram and is shown in

    Figure 4.5.

    Figure 4.5. Time-temperature-transformation (TTT) diagram: Iso-Tgcontours.

    As previously stated, the glass transition temperature is a useful tool to monitor

    the reactions that occur during cure. The iso-Tgcontours can therefore be seen

    as an extention of conversion curves. Vitrification of the system occurs when Tg

    rises to Tcure. Vitrification will be discussed in the next paragraph.

    Tcure (C)against LogTime (min)

  • 8/2/2019 Cure Thesis

    51/76

    40

    4.3.1.4 Prediction oftvit

    Data from Figure 4.5 was used to compile Table 4.3, which, in turn, was used

    to compile Figure 4.6. This illustrates the relationship between Ln time to a

    fixed Tgand the reciprocal ofTcure (1/K).

    The apparent activation energies, E, can be obtained from the slopes of the

    linear relationships in Figure 4.6 (as in Section 4.1.2): E values are 18.19,

    16.90, 26.80, and 46.82 kJ/mol, for the iso-Tg values 95, 98, 100, and 102C,

    respectively.

    Table 4.3 Numerical relationship between Ln time to a

    fixed Tgand the reciprocal ofTcure (1/K).

    Ln Time (min) Iso-Tg

    Tcure

    (C)

    Tcure

    (K) 1000/Tcure 93 95 98 100 102 104 105

    60 333 3.00 5.73 8.01

    80 353 2.83 3.73 4.14 5.00 6.26 7.71

    100 373 2.68 2.58 2.95 3.82 4.49 4.79 5.66 7.44

    120 393 2.54 2.30 3.09 3.73 3.89 4.37 5.07

    140 413 2.42 2.60 3.89 4.31 4.81 5.04

  • 8/2/2019 Cure Thesis

    52/76

    41

    1.00

    2.00

    3.00

    4.00

    5.00

    6.00

    7.00

    8.00

    2.30 2.40 2.50 2.60 2.70 2.80 2.90

    1000/Tcure (K)

    Ln

    Time

    (min)

    Tg: 95C

    Tg: 98C

    Tg: 100C

    Tg: 102C

    Figure 4.6. Ln time (min) to fixed Tgvalues (from Figure 4.5) vs. 1/Tcure (K).

    According to Gan et al. (1989:810) and Prime (1997:1616-1617) (Arrhenius

    Reference) the slope of the linear part of the Arrhenius plot, Figure 4.6,

    represents the apparent activation energies, E, of an epoxy (thermoset)

    system. Prime further explains that the activation energy may be viewed as a

    time-temperature shift factor where the cure process can be described by a

    single overall activation energy. The activation energies, E, as determined

    from the slopes of curves from Figure 4.6 were used to determine vitrification

    times as illustrated below in equations 4.1 to 4.4.

  • 8/2/2019 Cure Thesis

    53/76

    42

    The linear equations for the iso-Tgplots emanating from Figure 4.6 are listed

    below:

    186.181000.8827.7)(

    =Tcure

    tLn (r = 1) for iso-Tg= 95C [4.1]

    904.161000

    .7311.7)(

    =

    TcuretLn (r = 1) for iso-Tg= 98C [4.2]

    803.261000

    .672.11)(

    =

    TcuretLn (r = 1) for iso-Tg= 100C [4.3]

    824.461000

    .252.19)(

    =

    TcuretLn (r = 1) for iso-Tg= 102C [4.4]

    The times to vitrification, tvit, were calculated by setting Tg = Tcure in the iso-Tg

    equations, since vitrification is define to occur at Tg = Tcure. ForTcure = 95, 98,

    100 and 102C, the logarithm10 of times to vitrification are 1.447, 1.750, 2.077,

    and 2.26 minutes respectively.

    The calculated values were used to construct the isothermal TTT cure diagram

    simply because of the long experimental times for the system to vitrify.

    The method used by Gan et al. (1989) to calculate vitrification times for a

    thermosetting system, assumes the absence of diffusion control, solvent

    evaporation, and thermal degradation. Their method describes the relationship

    between the rate of reaction (-dx/dt), reactant concentration (x), activation

    energy (E), temperature (T,K) and time (t) with the Arrhenius equation:

    ( ) )(/exp xfRTEAdt

    dxa = [4.5]

    or by integrating

    ( ) ( ) tRTEAxF a = /exp [4.6]

  • 8/2/2019 Cure Thesis

    54/76

    43

    A being the Arrhenius pre-exponential factor, E the activation energy, R the

    gas constant (8,314 J/(mol.K)), and F(x), f(x) functions of x.

    The analysis of thermal data obtained in this study, with support from work

    performed by other researchers is summarized in an isothermal time-

    temperature-transformation (TTT) cure diagram. The vitrification curve in a TTT

    system cure diagram corresponds to a stage in the reaction at temperature Tg,

    for time tvit.

    The iso-Tgcontour for any corresponding pair of values ofTgand tvit is given by

    the following equation;

    ( ) )vitgacurecureatRTEtRTE = /exp/exp [4.7]

    Or

    ( ) curegcure

    avit tTT

    REt ln11

    /ln +

    = [4.8]

    The use of equation [4.8], in conjunction with data fortcure = 3 hours from Figure

    4.2, and E/R= 7882.7, obtained from time temperature conversion behaviour

    forTg = 95C (assumed to be constant for lower conversions), indicates thatcure at 60C for 3 hours gives Tg = 92.5C, with log tvit = 1.88 min forTcure =

    92.5C. Vitrification times were calculated for isothermal cure and presented in

    Table 4.4.

    Table 4.4 Calculated times to vitrification forTcure = Tg

    corresponding to isothermal cure temperatures from Figure 4.2.

    Tcure(C,

    From Fig 4.2)Tg (C)

    Tcure

    (C)

    tvit (Log

    min)

    60 92.5 92.5 1.88

    70 95.2 95.2 1.95

    80 97.7 97.7 2.02

    90 99.1 99.1 2.05

    100 101.4 101.4 2.11

    110 103.2 103.2 2.15

  • 8/2/2019 Cure Thesis

    55/76

    44

    Figure 4.7 indicates the relationship between Ln time to gelation and the

    reciprocal temperature (K);

    583,111000.9374,4)(

    =Tcure

    tLn (r = 0.986) [4.9]

    for which the apparent activation energy is 11.583 kJ/mol. The value of Tg at

    gelation is expected to vary with Tcure since the measurement of gelation in this

    study was carried out by means of parallel plate rheology.

    0.00

    0.50

    1.00

    1.50

    2.00

    2.50

    3.00

    3.50

    2.5 2.6 2.7 2.8 2.9 3 3.1

    1000/Tcure (K)

    Ln

    Time

    (min)

    Figure 4.7. Ln time (min) to macroscopic gelation vs. 1/Tcure (K).

    The isothermal temperature at which the time to vitrification equals the time to

    macroscopic gelation (extrapolated) is indicated as gelTg. This is calculated

    using Eq. (8), where tvit corresponds to reaction temperature gelTg, equation

    [4.9], and equation [4.10] which summarizes isochronal Tg versus Tcure data for

    tcure= 3 hours of Figure 4.2.

    5,77.25,0 += gcure TT (Tg

  • 8/2/2019 Cure Thesis

    56/76

    45

    The relevant expression becomes:

    193.5

    5,77'.25,0

    1

    '

    17.7882583,11

    '

    4,4937+

    +

    =ggelggelggel TTT

    [4.11]

    Solving the expression, gelTg = 40.204C, log tvit = 1.82 min for Tcure =

    40.204C.

    4.4 Time-Temperature-Transformation Diagram

    The results of the work done in the first two parts of this chapter can be

    summarized in the form of a time-temperature-transformation (TTT) isothermal

    cure diagram, as in Figure 4.8 where the progress of transformation of liquid to

    glass or to rubber is shown by the gelation, vitrification and iso-Tgcontours.

  • 8/2/2019 Cure Thesis

    57/76

    46

    Figure 4.8. Summary time-temperature-transformation (TTT) diagram.

    4.5 Mechanical property Vs. cure relationship

    4.5.1 Areas for Analysis

    The present investigation is aimed at indicating the relationship between cure

    and mechanical properties. To facilitate this, a TTT isothermal cure diagram

    (Figure 4.8) was used to determine regions or zones to illustrate the

  • 8/2/2019 Cure Thesis

    58/76

    47

    relationship and is presented in Figure 4.9 below. Figure 4.9 is a detailed

    version of Figure 4.8.

    Zone 1 through to zone 3 represents samples cured at Tcure = 70C for cure

    times (tcure) of 30, 60, and 120 minutes, zones 4 to 6 are representative of

    samples subjected to Tcure = 90C, and zones 7 to 9 to Tcure = 110C, for cure

    times of 30, 60, and 120 minutes respectively.

    Figure 4.9. Detailed TTT-Isothermal cure diagram showing areas

    for DTMA analysis.

    With reference to the generalised time-temperature-transformation (TTT) cure

    diagram (taken from Prime (1997:1384)), Figure 2.1, zones 1 and 2 fall into a

    conversion area beyond gelation but ahead of vitrification, zone 3 is after

    vitrification.

  • 8/2/2019 Cure Thesis

    59/76

    48

    4.5.2 Dynamic Mechanical Thermal Analysis (DMTA)

    As it would have been very time and material consuming, DMTA was chosen

    as method of determining mechanical properties instead of tensile, impact and

    bending tests. The results for storage modulus, E', and loss modulus, E as

    obtained from DMTA scans in this study are dealt with separately for samples

    cured isothermally at differentcure times (tc). The values forTcure and tc were

    determined by means of the time-temperature-transformation (TTT) diagram

    constructed in the previous section of this chapter.

    The DMTA temperature scans were performed at a rate of 10C/minute, for

    temperatures between 25 and 160C.

    4.5.2.1 Storage Modulus

    The scans show that systems cured at different temperatures (Tcure) display

    different moduli but broadly similar dynamic mechanical behaviour.

    Distinct changes in log(E') are evident and coincide with the glass transition

    temperature (Tg) which is represented by the peak tan values.

  • 8/2/2019 Cure Thesis

    60/76

    49

    DMTA Graphs : 70C

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    20 40 60 80 100 120 140 160

    Temperature (C)

    TanDelta

    6.0

    6.5

    7.0

    7.5

    8.0

    8.5

    9.0

    9.5

    10.0

    StorageModulusLogE'

    Tan Delta 30 min

    Tan Delta 60 min

    Tan Delts 120 min

    30 min @ 70C

    60 min @ 70C

    120 min @ 70C

    Figure 4.10-a. DMTA scans forTcure = 70C

    DMTA Graphs for 90C

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    20 40 60 80 100 120 140 160

    Temperature C

    TanDelta

    6.0

    6.5

    7.0

    7.5

    8.0

    8.5

    9.0

    9.5

    10.0

    StorageModulusL

    ogE'

    Tan Delta 30 min

    Tan Delta 60 min

    Tan Delta 120 min

    30 min @ 90C

    60 min @ 90C

    120 min @ 90C

    Figure 4.10-b. DMTA scans forTcure = 90C

  • 8/2/2019 Cure Thesis

    61/76

    50

    DMTA Graphs for 110C

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    20 40 60 80 100 120 140 160

    Temperature (C)

    TanDelta

    6.0

    6.5

    7.0

    7.5

    8.0

    8.5

    9.0

    9.5

    10.0

    StorageModulusLogE'

    Tan Delta 30 min

    Tan Delta 60 min

    Tan Delta 120 min

    30 min @ 110C

    60 min @ 110C

    120 min 110C

    Figure 4.10-c. DMTAscans forTcure=110C. Tan and storage modulus (E') vs. temperature

    (C) for samples cured for different cure times (tcure, min) at different cure temperatures (Tc).

    The results from Figure 4.10 are tabulated in Table 4.4 below. Subsequent

    Tables and Figures were derived from this Table.

    Table 4.4. Results from DMTA

    Tan E' (Pa) Tg (C) Time (min) Tan E' (Pa) Tg (C) Time (min) Tan E' (Pa) Tg (C) Time (min)

    70 0.74 1.69E+08 91.30 7.95 1.10 7.12E+07 96.06 8.45 0.83 8.45E+07 98.14 8.65

    90 0.99 6.01E+07 109.82 9.75 0.83 5.65E+07 111.13 9.93 1.15 8.60E+07 113.37 10.27

    110 0.75 1.08E+08 110.71 9.83 0.95 1.05E+08 114.84 10.42 0.79 5.51E+07 112.39 10.05

    Cure

    Temperatures

    (Tcure,

    C)

    30 min 60 min 120 min

  • 8/2/2019 Cure Thesis

    62/76

    51

    4.5.2.2 Glass transition temperature (Tg)

    A dynamic mechanical thermal analyser (DMTA) was used to determine the

    modulus and tan as a function of temperature and/or frequency. The samples

    were tested under three-point bending mode, known as dual cantilever. According to Laza et al. (1998), and Mafi et al. (2005), the glass transition

    temperature, Tg, is generally defined as the maximum in the tan curve.

    The contours of Tg values against tc (log time) as in Figure 4.11 below,

    indicates the relationship between Tg, Tcure, and tcas determined by DMTA.

    80

    90

    100

    110

    120

    0 20 40 60 80 100 120 140

    Cure Time (min)

    Tg

    (C)

    Tcure : 70C

    Tcure : 90C

    Tcure : 110C

    Figure 4.11. T