1-s2.0-S0143749611000194-main

download 1-s2.0-S0143749611000194-main

of 15

Transcript of 1-s2.0-S0143749611000194-main

  • 7/28/2019 1-s2.0-S0143749611000194-main

    1/15

    Nanocomposite adhesives: Mechanical behavior with nanoclay

    Erol Sancaktar n, Jason Kuznicki

    Department of Polymer Engineering, The University of Akron, Akron, OH 44325, USA

    a r t i c l e i n f o

    Article history:

    Accepted 22 September 2010Available online 21 February 2011

    Keywords:

    Novel adhesives

    Epoxy/epoxides

    Microscopy

    Mechanical properties of adhesives

    Epoxynanoclay composites

    a b s t r a c t

    The major objective for this research was to examine the role of epoxyclay nanocomposites in the area

    of epoxy bonding to porous stone (granite) substrates. Two bisphenol A epoxy systems were selected

    based on the prior work that determined optimal adhesive properties from a larger set of epoxy

    systems to determine the role of viscosity on the intercalation and exfoliation of the clay tactiods in the

    epoxy resin. The systems were characterized and mechanically tested at varying levels of intercalated

    and exfoliated organic clay tactiods. In the first stage of the work, epoxyclay systems were

    characterized by wide-angle X-ray diffraction (WAXD) to detect inter-laminar distances of clay layers

    and to determine if the mixing procedures had indeed dispersed and exfoliated the clay layers

    sufficiently. The second stage of the work involved examining mechanical properties of the epoxy

    nanoclay systems. Fracture behavior was studied using granite stone substrates in notched double lap

    configuration. Compressing a wedge between the cover plates induced the fracture. Fracture toughness

    was approximated by the load at fracture. Tensile properties were measured using cast dog bone tensile

    samples. The better layered silicate nanocomposite performance was seen with the lower viscosity

    resin. The most noticeable improvements in mechanical properties for the lower viscosity resin system

    were found to be maximum stress, elastic modulus, and yield stress. Increased toughness and stress

    whitening at 1% by weight nanoclay loading revealed that the clay can act as a shear-yielding

    toughening agent in this epoxy system.

    & 2011 Elsevier Ltd. All rights reserved.

    1. Introduction

    Nanotechnology provides improvements in the mechanical,

    thermal, and permeation properties of epoxy adhesives. In addi-

    tion, the use of nanoparticles such as nanofibers, nanotubes and

    graphite, various forms of nanosilicates offer promising improve-

    ments in the above mentioned adhesive properties with practi-

    cality and low cost.

    Clay particles of higher aspect ratio with approximately

    1 nm 10-100 nm dimensions can be successfully dispersed in

    the epoxy matrix to produce nanocomposites containing layered

    silicate of the smectic type. The particles are typically montmor-illonite, but can also be saponite and fluorohectorite forms of clay

    and have a structure different from the more common clays. They

    possess ion exchange properties, and thus, surface activity with a

    sodium cationic gallery surface for intercalation, while they react

    with organic materials [1,2]. Nanocomposites, which possess high

    stiffness and dimensional stability can also be used as flame

    retardant materials [3,4].

    For structural adhesive applications, the development of

    epoxy-based nanocomposites has drawn considerable attention.

    Pinnavaia et al. [510] used layered silicate nanolayers as alter-

    native inorganic components for the construction of epoxy-based

    nanostructured composites. They reported that clay silicate

    nanolayers possess high particle aspect ratios comparable to

    conventional fibers and they modified the nanocomposite inter-

    layer surface by ion-exchange reaction, intercalating the nanoclay

    galleries by organic polymer precursors for the formation of

    organicinorganic nanocomposites with stable SiO bonds.

    They reported that the platy morphology of the silicate layers,

    exfoliated to form clay nanocomposites, resulted in drama-tically improved properties such as barrier and mechanical

    properties that are not available with conventional composite

    materials [510].

    The synthesis of amine-cured epoxy nanocomposites through

    the polymerization of monomers in the galleries of protonated

    onium ion exchanged forms of smectic clays depends on two

    crucial factors: firstly, the ability of the onium ion to serve as a

    compatibilizing agent, which allows for cointercalation of the

    resin and curing agent, and secondly, the ability of the onium ion

    to acid-catalyze the intragallery ring opening polymerization

    reaction. The catalytic function of the onium ion is important

    because, as noted above, it allows the intragallery polymerization

    Contents lists available at ScienceDirect

    journal homepage: www.elsevier.com/locate/ijadhadh

    International Journal of Adhesion & Adhesives

    0143-7496/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.ijadhadh.2010.09.006

    n Corresponding author. Tel.: +1 330 972 5508.

    E-mail address: [email protected] (E. Sancaktar).

    International Journal of Adhesion & Adhesives 31 (2011) 286300

    http://-/?-http://www.elsevier.com/locate/ijadhadhhttp://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ijadhadh.2010.09.006mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ijadhadh.2010.09.006http://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ijadhadh.2010.09.006mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ijadhadh.2010.09.006http://www.elsevier.com/locate/ijadhadhhttp://-/?-
  • 7/28/2019 1-s2.0-S0143749611000194-main

    2/15

    rate to be competitive with the extragallery polymerization rate.

    Consequently, the nanolayers can be completely dispersed in the

    polymer matrix, where they can fully contribute to the reinforce-

    ment mechanism.

    The exfoliation and intercalation of montmorillonite nano-

    layers in an epoxy matrix is readily achieved with acidic organo-

    philic alkylammonium ions such as C18H37NH3+ abbreviated as

    C18A. According to Pinnavaia et al., it is necessary for the onium

    ion to serve as a compatibilizing agent to allow cointercalation ofthe resin and curing agent, and to acid-catalyze the intragallery

    ring opening polymerization reaction. This is necessary for the

    synthesis of amine-cured epoxy nanocomposites through the

    polymerization of monomers in the galleries of protonated onium

    ion exchanged forms of smectic clays [810].

    Pinnavaia et al. utilized layered silicate magadiite with a layer

    thickness of 1.12 nm, exfoliated in an epoxy matrix for the

    formation of polymerinorganic nanolayer composites. Shell Epon

    828 elastomeric epoxy nanocomposite reinforced by magadiite in

    this fashion exhibited solvent uptake reduced by 2.5-fold in

    comparison to the neat resin, as illustrated by an immersion in

    propanol for 50 days. Pinnavaia et al. report that after 50 days

    immersion in propanol the pristine polymer absorbs 2.5 times

    more than the nanocomposite, and at this time the pristine

    sample began to crack and break up, whereas the shape and

    texture of the nanocomposite sample appeared unchanged [8].

    We have to be careful, however, in differentiating between

    changes in the diffusion coefficient versus the equilibrium moist-

    ure content of nanocomposite resins as a result of nanoclay

    additions. For example, Rana [11] and Rana et al. [12] studied

    moisture diffusion through neat vinyl ester resin containing

    nanoclay by performing transient and steady-state diffusion

    experiments and illustrated that montmorillonite clay was

    effective in reducing the diffusion coefficient of water through

    the vinyl ester resin. Rana went on to report that the solubility

    of water in the nanocomposites appeared to increase with the

    amount of clay, as reflected in the equilibrium moisture content.

    This was due to water adsorption onto the clay. Rana hypothe-

    sized that the clay acted as a moisture scavenger by trapping

    water and prevented it from reaching any further. His transient

    diffusion experiments revealed that the diffusion coefficient, D, of

    water at room temperature dropped by approximately 92% after

    an incorporation of 5 wt% of Cloisite 10As clay. His results

    showed an increase from 0.66% to 2.08% in equilibrium moisture

    content during his transient diffusion experiments with nano-

    composite films in distilled water at 25 1C, while D went down

    from 7.36109 to 0.609109 cm2/s. The advantage in incor-

    porating the clay was even more dramatic for diffusion at

    elevated temperatures as for the diffusion of distilled water at

    42.5 1C, with D going down from 16.43109 to

    0.93109 cm2/s. The excess moisture acquired due to the

    presence of reinforcing (standard) glass fibers was reduced from

    4.360% with no-clay to 3.221% at room temperature, with theaddition of 5% clay [11].

    Results by Rana revealed that water uptake with clay rein-

    forced epoxy nanocomposite samples should not be measured

    using direct (thickness-normalized) water absorption experi-

    ments, but rather using diffusion through fully saturated samples,

    as well as an excess moisture content in epoxy bonded samples

    in-situ. Such measurements should be made by normalizing with

    equilibrium (saturation) absorption levels. Since the clay particles

    scavenge water, the nanocomposite samples initially absorb

    slightly higher amounts of water in comparison to the no-clay

    samples, with the water molecules congregating around the clay

    particles. On the other hand, the presence of these clay particles

    still hinders diffusion of water through the sample, thus protect-

    ing the structural interfaces.

    Rana also reported increases in the polymer glass transition

    temperature and mechanical properties with nanoclay addition

    (Cloisite 10A); however, these increases were obtained by only up

    to 1% nanoclay addition by weight, except for the tensile modulus.

    Addition of higher amounts of nanoclay brought these material

    properties back to the levels obtained with the neat resin. The

    tensile modulus was found to increase monotonically, up to

    $13%, when Cloisite 10A was added up to 5% by weight. With

    1 wt% nanoclay addition, the notched impact strength increased$10%, the tensile strength increased $26% and strain at failure

    increased $17%. The torsional properties were similarly

    enhanced [11].

    Since epoxy applications may exist in areas of high moisture

    content and under mechanically induced stress, the effect of such

    stressing on water uptake by epoxyclay nanocomposites is also

    of interest. Sancaktar and Kuznicki [13] used low viscosity liquid

    aromatic diglycidyl ether of bisphenol A (DGEBA) epoxy resin

    Epon 815C mixed with 0.5% nanoclay (Cloisite 30B) by weight to

    produce an exfoliated clayepoxy resin nanocomposite system.

    These samples were immersed in water in stressed condition

    (flexural stress) to assess the effect of stress on nanocomposite

    epoxy system for its water uptake behavior. Application of the

    flexural stress affected the water uptake barrier properties for

    nanoclay/epoxy nanocomposites, with the flexural stress acting to

    decrease the rate of absorption as well as to decrease the

    equilibrium moisture content in the nanocomposite. The results

    revealed up to 33% reduction in water uptake for the stressed

    samples [13].

    Park and Jana [1416] reported on conditions for complete

    exfoliation of nanoclay particles in epoxynanoclay composite

    systems. These authors developed master curves from the values

    of storage modulus (G0) of curing epoxy inside the clay galleries

    and the complex viscosity (Zn) of curing epoxy outside the

    galleries to show that complete exfoliation is produced if the

    ratio, G0/Zn is greater than approximately 2 (1/s); otherwise,

    intercalated structures are obtained. They argued that the values

    of G0 are strong functions of organic treatment of clay, the

    chemical structure of curing agents, and the chemical structures

    of epoxy resin and provided a comprehensive list of cure pro-

    cesses for epoxy/nanoclay systems [14].

    Property enhancement in nanoparticle reinforced polymers

    depends on the degree of nanoparticle dispersion [1719], which

    can be achieved by dispersion of nanoparticles into low viscosity

    components (for example liquid hardeners for thermoset resins)

    followed by in-situ polymerization [1820]. The stages of nano-

    particles dispersion in polymeric matrices involve: (a) intercalation

    of particles by polymer chains (a nanoscopic process); (b) flow-

    induced exfoliation of individual particles (a mesoscopic process),

    and; (c) homogeneous dispersion of exfoliated particles into the

    matrix (a macroscopic process). The first of these processes, the

    intercalation process is governed solely by diffusion [21], while the

    exfoliation and homogenization processes are controlled by mix-ing. Therefore, step (a) involves clay treatment, wet-out and

    intercalation. In clay treatment, functionalized onium ions

    appropriate for the desired matrix polymer replace the inorganic

    cationic counterions adsorbed to the inner surfaces of the clay

    galleries to balance the negative charges resulting from an iso-

    morphic substitution of lower valence atoms in the clay lattice. The

    clay aggregates are infiltrated with polymer or precursor during

    the wet-out stage. Intercalation occurs when the polymer or

    precursor molecules enter, and swell the galleries. Subsequently,

    layers separate and become individual during the exfoliation

    step (b). The nano-platelets created in this manner still need to be

    dispersed homogeneously in the matrix with step (c) [22].

    As mentioned above, low amounts of montmorillonite rein-

    forcement typically result in large increases in stiffness,

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 287

  • 7/28/2019 1-s2.0-S0143749611000194-main

    3/15

    permeability and heat resistance along with improvements in

    tensile and impact strength for the resulting nanocomposites. The

    degree of delamination of the clay tactoids to result in large

    aspect ratio of the dispersed clay laminas increase permeation

    resistance of the nanocomposite. The resulting nanoclay layer

    spacing gives an idea about the degree of intercalation towards

    exfoliation. This (d001) spacing of the clay crystal lattice is

    typically measured by the wide-angle X-ray scattering. Interpre-

    tation of the spectra can also provide information on the dis-tribution of stacking sequences [19,23,24].

    The results on the mechanical behaviors of clay reinforced

    nanocomposite polymers reported above reveal the necessity of

    optimization with simultaneous considerations of water diffusion

    and strength/strain properties. The important mechanical proper-

    ties: tensile strength, impact strength, and failure strain seem to

    reach their maximum values at rather lower weight percentages

    of clay loading, approximately 1% by weight, and seem to go back

    to and, more or less, remain near the no-clay values at nanoclay

    loadings above approximately 1% by weight. The diffusion coeffi-

    cient, however, decreases monotonically with clay loading. Con-

    sequently, if increasing the mechanical properties of the

    nanocomposite adhesive is not the main aim, then one can load

    58% clay in epoxy adhesives to obtain significant reductions in

    the coefficient of diffusion, thus improving the durability and

    environmental resistance of structures adhesively bonded by

    these nanocomposite adhesives.

    Another method of reinforcing epoxy adhesives for improve-

    ments in mechanical properties along with the creation of elec-

    trical conduction function for them is the use of electrospun

    polyacrylonitrile (PAN) fiber precursor-based carbon nanofiber

    (CNF) mats incorporated in epoxy resin as shown by Sancaktar

    and Aussawasathien [2527]. Mechanical properties of electro-

    spun carbon nanofibers (ECNF)/epoxy nanocomposites were

    obtained using tensile and flexural tests [25]. The tensile modulus

    of the pure epoxy matrix was found to be 5.7 GPa and increased by

    46% to 8.4 GPa when reinforced by 6 g wt% ECNF mat. The

    maximum flexural tensile stress value at the outer surfaces of

    the neat epoxy samples was measured to be approximately

    330 MPa, which was approximately twice that of the model epoxy

    tensile strength (approximately 150 MPa). When increasing

    amounts of the ECNF mat were added to this neat resin, however,

    the maximum flexural tensile stress at the outer surface decreased

    as a result of the embrittlement effect of the ECNF mat in the

    composites. Note that this embrittlement effect, which reduced

    the flexural strength of the nanocomposite, was not observed with

    the tensile behavior, where the ultimate tensile strength increased

    by as much as 27% with 2.06% by weight ECNF mat addition.

    Our previous unpublished works using Epon 815C/Epicure

    3140 curing agent system revealed the following results regard-

    ing the effects of synthetic graphite nanoparticles or Multiwall

    Carbon Nanotubes (MWCNT) reinforced nanocomposites: condi-

    tions resulting in 1.31% water absorption in the neat resinresulted in 1.22% and 1.09% water absorption along with 20%

    and 30% increase in tensile moduli values, respectively, when

    synthetic graphite nanoparticles or MWCNT were added in 1% by

    weight proportions, respectively.

    The primary objectives of this current study were as follows:

    Investigate epoxynanoclay composite formation, using two

    common epoxy resins with differing viscosities.

    Determine the role of viscosity on the intercalation and

    exfoliation of the clay tactiods.

    Explore the effect of clay loading on mechanical properties of

    resins.

    Explore effects of mechanical interlocking and crack propaga-

    tion with epoxynanoclay composites using fracture tests on

    porous granite substrates bonded with the nanocomposite

    adhesive and approximate fracture toughness of the nanocom-

    posite adhesive systems using the load at fracture.

    Thus, we hope to improve the performance of epoxynanoclay

    composites in practical applications such as in bonding porous

    stones (granite), which are used extensively in construction.

    Many such applications will involve epoxy systems under stress,

    in humid areas, or under potential fracture. This study attempts tofirst discover basic mechanical benefits of adding different

    amounts of nanoclay to adhesive systems. Also, the ability of

    the nanoclay phase to work against crack propagation will be

    explored, specifically in a mechanical interlocking adhesion situa-

    tion, with adhesive having diffused into the porous granite

    subsequently failing during fracture testing.

    2. Materials and experimental procedures

    2.1. Epoxies and curing agents

    Two epoxy materials, one type of curing agent, and one type of

    adherend material were studied in this work. Epon 815C and Epon830 are DGEBPA resins manufactured by Resolution Performance

    Products in Houston, Texas, USA. Epon 815C is a low viscosity

    epoxy resin that reacts with a wide range of curing agents. It is

    made up of a mixture of 86.4% DGEBPA epoxy combined with 13.6%

    monofunctional butyl glycidyl ether called Heloxy Modifier 61

    (Resolution Performance Products, Houston, TX, USA). Heloxy

    modifiers are reactive diluents that are either monofunctional to

    minimize cost, or polyfunctional to maximize properties. Epon

    815C is a good choice for adhesion applications as well as

    encapsulating and coating processes. Epon 830 is made up of

    100% DGEBPA resin, resulting in an above average viscosity, but

    low crystallization tendency. This resins chemical resistance makes

    it a good choice for industrial floor coatings and grouting and it is

    typically the base resin for Heloxy-modified resins. The resin andmodifier properties are given in Table 1. The curing agent used was

    Epi-cure 3223, an unmodified aliphatic amine diethylenetriamine

    (DETA, Miller-Stephenson Chemical Co., Danbury, CT, USA).

    2.2. Organically treated clay

    The organically modified Clay was supplied by Southern Clay

    Products in Gonzales, Texas. It is Montmorillonite clay modified

    with methyl, tallow, bis-2-hydroxyethyl, quaternary ammonium

    (Fig. 1). The clay is made up of approximately 1 nm thick layers

    electrostatically held together in the form of micron sized

    tactiods. Table 2 shows the properties of this clay.

    2.3. Intercalation and curing procedures

    The procedure for the formation of epoxyclay nanocompo-

    sites using Cloisites 30B was based on the previous research [14].

    Table 1

    Properties of Epon series resin and modifiers.

    Resin/

    modifier

    Epoxide equivalent

    weight (g/eq)

    Viscosity at

    25 1C (P)

    Density

    (g/cm3)

    Number average

    molecular weight

    815C 180195 57 1.13 700

    830 190198 170225 1.16 700

    Modifier

    61

    145155 12 0.91

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300288

  • 7/28/2019 1-s2.0-S0143749611000194-main

    4/15

    The following procedure was used:

    1. The epoxy resin and clay were stirred at 1000 rpm by magnetic

    stirrer for 6 h at 90 1C.

    2. The mixture was placed in a vacuum oven for 30 min at 80 1C.

    This was found to optimize intercalation.

    3. The mixture was cooled to room temperature and held for 15 min.

    4. A stoichiometric amount of curing agent (see Tables 3 and 4)

    was added to the resin and clay mixture and mixed by handfor 5 min at room temperature.

    5. The resin-clay-curing agent mixture was degassed in a room

    temperature vacuum oven for 510 min to remove air bubbles.

    6. The mixture was placed in a dog bone cast mold, film mold, or

    applied to granite fracture samples.

    7. The system was cured at the temperature and time suggested

    by the manufacturer for the application.

    The ratios and formulations for a 20 g sample for both epoxy

    systems at various clay loadings are given in Tables 3 and 4. Cure

    conditions and procedures are shown in Tables 5 and 6. The film

    samples were gelled at the desired thickness at room temperature

    in a compression mold as a first step. All other samples were

    cured in one step.

    2.4. X-ray characterization

    One-dimensional wide-angle X-ray diffraction measurements

    of the epoxyclay nanocomposites were carried out on Rigaku

    X-ray diffractometer (Tokyo, Japan) with 1.54 A (Cu) wavelength,

    50 kV tube voltage, and 150 mA tube current. The 2-theta scan

    range 1.581 at an interval of 0.0501 at a rate of 1.001/minute.

    Both transmittance and reflectance spectrums were collected.

    Samples of cured epoxyclay nanocomposites at various loadings

    were tested to determine the interlayer spacing of clay layers and

    the extent of intercalation or exfoliation.

    2.5. TEM characterization

    The transmission electron microscope (TEM) samples were

    prepared by microtoming in a Reichert Ultracut Microtome

    (Depew, NY, USA) at room temperature. The samples were then

    analyzed in an FEI model TANCNAI-12 TEM (Hillsboro, OR, USA),

    using 300 mesh copper grids. TEM images were used to confirm

    findings from X-ray diffraction on the degree of intercalation and

    exfoliation, as well as the morphology of the dispersed clay phase.

    2.6. Tensile testing

    Bulk tensile test specimens were prepared at clay loadings of

    0%, 0.5%, 1%, 2%, and 4% to determine the mechanical properties.

    The epoxy systems were cast in aluminum dog bone tensile

    specimen molds and cured at the cure conditions shown

    in Tables 5 and 6. The mold was coated with Teflon-based mold

    release agent to aid in demolding. The samples were removed and

    cooled for 1 h and then sanded. The samples were then tested onan Instron 5567 machine (Canton, MA, USA), using a 1 kN load cell

    and a crosshead speed of 5 mm/min. Fig. 2 shows the dimensions

    of the dog bone sample. The sample thickness was around

    1.5 mm, but some variations occurred in thickness and width

    due to sanding of the samples to remove flash and other edge

    effects. The sanding procedure started with 150 grit power

    sanding, followed by 300 and 400 grit hand sanding.

    2.7. Adherend preparation and adhesive bonding

    The natural solid granite adherend samples were prepared for

    fracture testing by first cutting them with a Work Force CTC550

    wet saw (Atlanta, GA, USA) into rectangular specimens of 50 mm

    length, 30 mm width, and 10 mm thickness. The samples wererinsed with water, cleaned with compressed air, then rinsed with

    acetone and allowed to dry. Double lap joint specimens (Fig. 3)

    were prepared using an overlap length of 12.0 mm with an initial

    crack length of 7.0 mm. These dimensions provided adhesive

    failure rather than failure of the adherend during initial trials.

    The samples were clamped under a pressure of 0.550 MPa during

    CH2CH2OH

    CH3 N+ (tallow)

    CH2CH2OH

    Fig. 1. Chemical structure of methyl, tallow, bis-2-hydroxyethyl, and quaternary

    ammonium.

    Table 2

    Properties of Cloisites 30B clay.

    Clay type Modifier

    Concentration

    (micro- equivalent/

    10 g clay)

    Specific gravity X-ray diffraction

    results (A)

    Typical dry particle sizes (mm)

    10% Less than 50% Less than 90% Less than

    Cloisites 30B 90 1.98 18.5 2 6 13

    Table 3

    Formulation of Epon 815C, DETA, and 30B.

    Weight percentages (%) 20 g Sample formulation (g)

    30B

    clay

    815C

    Resin

    DETA curing

    agent

    30B

    clay

    815C

    Resin

    DETA curing

    agent

    0.0 89.00 11.00 0.00 17.80 2.20

    0.5 88.56 10.95 0.10 17.71 2.19

    1.0 88.11 10.89 0.20 17.62 2.18

    2.0 87.22 10.78 0.40 17.44 2.16

    4.0 85.44 10.56 0.80 17.09 2.11

    Table 4

    Formulation of Epon 830, DETA, and 30B.

    Weight percentages (%) 20 g Sample formulation (g)

    30B

    clay

    830

    Resin

    DETA curing

    agent

    30B

    clay

    830

    Resin

    DETA curing

    agent

    0.0 90.17 9.83 0.00 18.03 1.97

    0.5 89.72 9.78 0.10 17.94 1.961.0 89.27 9.73 0.20 17.85 1.95

    2.0 88.37 9.63 0.40 17.67 1.93

    4.0 86.56 9.44 0.80 17.31 1.89

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 289

  • 7/28/2019 1-s2.0-S0143749611000194-main

    5/15

    cure to ensure uniform adhesive thickness of around 0.08 mm and

    adequate bonding of the substrates. The various formulations of

    epoxy and clay were prepared and applied to the areas to be

    bonded. The initial crack was formed using a double-edged razor

    blade coated with a UV-cured polysiloxane coating, and the

    starter blade was removed after curing. The initial crack length

    was determined through experimentation and served to initiate

    fracture failures.

    2.8. Fracture testing

    The specimens for fracture tests were prepared as described

    in Section 2.7. A wedge type fracture method was developed and

    employed using a wedge with a vertex angle of 46.161. The wedge

    was placed in between the double side of the specimen and a

    compressive load was applied until failure. The use of a 50 kN

    load cell was found necessary based on the greater of the crushing

    force of the granite substrate and shear strength of the epoxy. The

    fracture toughness was approximated by load at fracture. After

    considerable testing, the load cell required for dimensions in thiswork was found to be 30 kN. The crosshead speed of 5 mm/min

    was used based on the ASTM D3931 method [28]. This test (Fig. 4)

    was used to find the load at the catastrophic crack propagation in

    an attempt to assess the fracture properties of the epoxy systems

    on granite.

    2.9. Film sample preparation for TEM analysis

    The film samples were cast using Teflon and PET compression

    mold fixture, illustrated in Fig. 5. The prepared uncured filled and

    neat epoxy systems were poured into the compression mold and

    allowed to gel, while under 34.5 MPa pressure to ensure an

    average thickness of 0.15 mm. The materials gelled at room

    temperature were then removed and cured at 1201

    C for 3 h for

    Epon 815C, and at 100 1C for 1 h for Epon 830. Since certain

    volatiles may escape, especially during elevated temperature

    curing, efforts were made to make the film casting process

    comparable to the fracture and tensile test specimens. Thus, thecuring was done after the removal of the top Teflon layer, so that

    the gelled film was stable enough not to adversely affect the film

    thickness uniformity.

    Once the film samples were cured they were subjected to

    X-ray diffraction and TEM analysis.

    3. Results and discussion

    3.1. Interpretation of X-ray and TEM findings

    Various samples were tested in an attempt to determine the

    interlayer distances of the dispersed clay tactiods and subse-

    quently the degree of intercalation or exfoliation in the 815C-

    DETA system. First, clays that were treated by the manufacturer

    and clays treated in our laboratory were compared in their

    powdered form. Fig. 6 shows clay 30B to have peak near 4.91representing a d-spacing of 1.80 nm (Fig. 6b). This shows good

    agreement with what was reported by the manufacturer

    (Table 2). We also treated the clay, which was supplied by

    Southern Clay Products Co. in an untreated form, using hexade-

    cylamine, and obtained d-spacing of 1.84 nm (Fig. 6a). We expect

    little difference in performance of nanocomposites filled with clay

    30B or the clay we treated using hexadecylamine, due to small

    difference in interlayer distances achieved. Consequently the

    pretreated 30B clay was used for the remainder of the work.

    Fig. 7 shows the X-ray results for various mixing and curing

    procedures tested to confirm their effectiveness. The variousconditions used are outlined in Table 7, with all having 4% weight

    Cloisites 30B. Condition 7a is described in [14]. With condition

    7b, the mixing of the epoxy with organoclay was doubled to 12 h

    with rest of the procedure being the same as that for condition 7a.

    Condition 7c represents a system using a two-stage curing

    procedure recommended by the manufacturer of the resin and

    the curing agent. For condition 7d, the resin and clay were mixed

    for the standard time, but extensive hand mixing and subsequent

    ultrasonic agitation were used after the addition of the curing

    agent. The final condition (7e) depicted is a system using 24%

    weight acetone added to the clay as an initial swelling agent prior

    to magnetic mixing.

    Fig. 8 shows more details of the effect of the acetone on the

    clay, showing curves for the clay only (Fig. 8b), clay swelled with

    Table 5

    Cure conditions and procedures for Epon 815C, DETA, and 30B.

    815C/DETA/30B Step 1 Step 2

    Specimen type Time (h) Temperature (1C) Pressure (MPa) Time (h) Temperature (1C) Pressure (MPa)

    Dog bone 3 120 0.101

    Fracture 3 120 0.550

    Film 4 26 34.5 3 120 0.101

    Table 6

    Cure conditions and procedures for Epon 830, DETA, and 30B.

    830/DETA/30B Step 1 Step 2

    Specimen type Time (h) Temperature (1C) Pressure (MPa) Time (h) Temperature (1C) Pressure (MPa)

    Dog bone 1 100 0.101

    Fracture 1 100 0.550

    12.7 mm

    47.0 mm

    13.0 mm5.0 mm

    Fig. 2. Dimensions of tensile specimen.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300290

  • 7/28/2019 1-s2.0-S0143749611000194-main

    6/15

    acetone (8c), and the nanocomposite system formed (8a). Exam-

    ination of Figs. 7 and 8 leads to the following observations:

    The first four curing and mixing conditions (7a, b, c, and d)

    showed small differences in the degree of intercalation, as

    shown by the interlayer spacings measured.

    Blade is removed to form

    initial crack of 7.0 mmEpoxy thickness

    of 0.08 mm

    12.0 mm

    50.0 mm

    30.0 mm

    10.0 mm

    5.0 mm

    Fig. 3. Diagram of double lap joint.

    2 = 46.16

    FMode I = (F/2)*tan ()

    F

    FModeII = F/2

    Fig. 4. Force diagram of wedge and double lap sample used in fracture testing

    showing the force components acting on the cover plates to induce fracture.

    Teflon

    Teflon

    PET

    2.0 mm

    0.15 mm

    Fig. 5. Diagram of mold used for film casting.

    1.5

    2 Theta

    RelativeIntensity(a.u.)

    4.10 nm

    4.20 nm

    4.20 nm

    (a)

    (b)

    (c)

    (d)

    (e)

    4.41 nm

    2.5 3.5 4.5 5.5 6.5 7.5

    Fig. 7. X-ray diffraction patterns for 815C-DETA-4% 30B epoxy system at various

    curing and mixing conditions: (a) standard procedure; (b) 2 xs epoxy-clay mixing

    time; (c) 2-step cure; (d) 7 xs epoxy-clay-curing agent mixing time; and

    (e) acetone added.

    1.5

    2 Theta (Degree)

    Rea

    ltive

    Intens

    ity

    (a.u.

    )1.84 nm

    1.80 nm

    (a)

    (b)

    2.5 3.5 4.5 5.5 6.5 7.5

    Fig. 6. X-ray diffraction patterns for (a) Cloisites Na+ treated with hexadecyla-

    mine and (b) Cloisites 30B.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 291

  • 7/28/2019 1-s2.0-S0143749611000194-main

    7/15

    The additional 6 h of mixing in condition 7b did not show

    substantial benefit, only decreasing the interlayer distance

    from 4.41 to 4.20 nm.

    The two-stage curing (7c) using a room temperature stage and

    an elevated temperature stage resulted in a decrease in an

    interlayer distance (4.10 nm) compared to the system cured inone stage at an elevated temperature (4.41 nm). This is

    probably because, in the single stage, the temperature

    increases early in the curing process, thus reducing viscosity

    of the system and lowering the extra-gallery viscous forces

    that work against the exfoliation process (see Fig. 9) [14].

    The extensive mixing and ultrasonic agitation of the system as it

    gelled, also did not provide intercalating benefit (condition 7d). A

    similar effect is seen as in condition 7c; the increase of viscosity

    as the system gelled actually worked to hinder the effectiveness

    of the exfoliating elastic forces inside the clay galleries, despite

    the additional mixing. Although condition 7d showed less inter-

    layer spacing than the standard system 7a (4.20 compared to

    4.41 nm), it had higher gallery height compared to the two-stage

    curing process of condition 7c (4.20 compared to 4.10 nm). This

    is probably because the system cured (or rather gelled) for

    35 min rather than 3 h as in the two-stage cure, and the

    additional shear mixing perhaps aided in intercalation.

    Due to the increase of viscosity as the system gelled, mixing

    past the gel point in condition 7d increased the likelihood ofair bubbles being trapped in the system.

    The greatest interlayer spacing for 4% weight clay loading was

    found in condition 7e. The addition of acetone as a swelling

    agent showed good X-ray results compared to all of the other

    conditions in Table 7 and Fig. 7. The reason is best seen

    in Fig. 8. The addition of the organic solvent swells the

    organoclay from 1.80 to 4.77 nm (8b8c). This mixture, mag-

    netically-stirred into the resin, results in a nanocomposite

    with a peak shifting far to the left indicating a d-spacing of

    greater than 5.88 nm (8a and 7e).

    While the addition of acetone to the system proved to increase

    the intercalation height of the 4% clay loaded nanocomposites,

    subsequent mechanical testing proved that residual acetonemolecules in the epoxy system acted as inhibitors to crosslinking,

    resulting in a decrease in crosslink density and consequently

    decrease in mechanical properties. This decrease outweighed the

    benefits of the increased d-spacing.

    The standard mixing procedure in condition 7a was thus

    chosen as the optimal procedure among those considered. Next,

    X-ray measurements were taken of samples with the clay loading

    varying from 0% to 4% as shown in Fig. 10:

    The results showed larger d-spacing for the nanocomposites

    with lower clay loadings. This is most likely because having

    less clay particles increases the chance that resin molecules

    will intercalate well into the clay galleries. TEM images

    (discussed later) show that the smaller-sized aggregates of

    Table 7

    Various curing and mixing conditions used on 815C-DETA-4% 30B epoxy system.

    Sample Mixing condition (with 4% 30B) Curing condition X-ray result (nm)

    Resin/clay Resin/clay/curing

    4.1.2a 6 h 90 1C (magnetic stirrer) 5 min 25 1C (hand mix) 3 h, 120 1C 4.41

    4.1.2b 12 h 90 1C (magnetic stirrer) 5 min 25 1C (hand mix) 3 h, 120 1C 4.20

    4.1.2c 6 h 90 1C (magnetic stirrer) 5 min 25 1C (hand mix) 3 h, 25 1C + 4.10

    2 h, 1501C

    4.1.2d 6 h 90 1C (magnetic stirrer) 35 min 25 1C (hand mix+ultras onic) 3 h , 12 0 1C 4.01

    4.1.2e 6 h 90 1C with acetone (magnetic stirrer) 5 min 25 1C (hand mix) 3 h, 120 1C 45.88

    1.5

    2 Theta (degree)

    RelativeIntensity(a.u.)

    1.80 nm

    (a)

    (b)

    (c)

    4.77 nm

    2.5 3.5 4.5 5.5 6.5 7.5

    Fig. 8. X-ray diffraction patterns for (a) 815C-DETA-4% 30B-acetone nanocompo-

    site, (b) 30B clay, and (c) 30B clay swelled with acetone.

    van der Waals

    Electrostatic forceElastic force

    Extra-gallery viscous force

    Extra-gallery viscous force

    (Clay layer)

    Fig. 9. Schematic representation of forces acting on a pair of clay layers during

    intercalation and exfoliation.

    1.5

    2 Theta (degree)

    Rela

    tiveIntensity(a.u.) 4.64 nm

    5.19 nm

    (a)

    (b)

    (c)

    (d)

    (e)

    (f)

    4.41 nm

    2.5 3.5 4.5 5.5 6.5 7.5

    Fig. 10. X-ray diffraction patterns for 815C-DETA-30B epoxy system at various

    clay loadings: (a) 4%; (b) 2%; (c) 1%; (d) 0.5%; (e) 0%; and (f) air.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300292

  • 7/28/2019 1-s2.0-S0143749611000194-main

    8/15

    clay in the lower weight percent samples also seem to have

    larger d-spacing. The near featureless curve of the 0.5% nanocomposite

    (Fig. 10d) is a good indication of the exfoliation of the clay

    tactiods to a distance greater than 5.88 nm. Thus, best barrier

    and mechanical properties should be expected at this loading.

    The curves are also compared to air (10f) and neat epoxy (10e)

    as baselines.

    Some Epon 815C systems prepared using different curing and

    mixing procedures were imaged with TEM to further investigate

    the morphology of the clay and epoxy already investigated by the

    X-ray diffraction. Figs. 11 and 12 show TEM images of the 2% clay

    loaded system at different magnifications. Figs. 13 and 14 show

    the system with 0.5% weight clay at the same magnification for

    comparison. The TEM samples were collected by microtoming

    Fig. 11. Low magnification TEM image of 815C/DETA with 2% weight clay loading,

    showing typical size of a clay aggregate (18,500 magnification).

    Fig. 12. High magnification TEM image of 815C/DETA with 2% weight clay loading,

    showing mostly intercalated nanocomposite structure (195,000 magnification).

    Fig. 13. Low magnification TEM image of 815C/DETA with 0.5% weight clay

    loading, showing typical size of a clay aggregate (18,500 magnification).

    Fig. 14. High magnification TEM image of 815C/DETA with 0.5% weight clay

    loading, showing exfoliated nanocomposite structure (195,000 magnification).

    1.5

    2 Theta (degree)

    RelativeIntensity(a.u.)

    5.04 nm

    5.04 nm

    (a)

    (b)

    (c)

    (d)

    (e)

    5.04 nm

    5.04 nm

    2.5 3.5 4.5 5.5 6.5 7.5

    Fig. 15. X-ray diffraction patterns for 830-DETA-30B epoxy system at various clay

    loadings: (a) 4%; (b) 2%; (c) 1%; (d) 0.5%; and (e) air.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 293

  • 7/28/2019 1-s2.0-S0143749611000194-main

    9/15

    along the cross-section of the film samples created in the manner

    stated in Section 2.9. The TEM results showed as follows:

    The TEM images seem to agree relatively well with the

    interlayer spacing found in X-ray diffraction results. TEM

    shows areas of both intercalation and exfoliation, with exfo-liated and intercalated clay aggregate structures similar to

    those found by Park and Jana and others (Figs. 11 and

    13) [14,29].

    As suggested earlier by the X-ray diffraction, it appears that

    there is a slightly better exfoliation of the clay layers in 815C

    samples with less clay loading. TEM images showed that

    typically, 0.5% weight clay content samples had smaller

    aggregates compared to the 2% loaded samples (Fig. 11 com-

    pared to Fig. 13). This smaller aggregate size results in less

    interference of clay tactiods, and consequently better exfolia-

    tion as seen in Fig. 14 compared to 12.

    For the most part, the clay layers seemed to remain parallel

    even for the samples with large d-spacing. This morphology

    has been shown to be beneficial for creating zones of shear

    Fig. 16. Low magnification TEM image of 830/DETA with 2% weight clay loading,

    showing typical size of a clay aggregate (18,500 magnification).

    Fig. 17. High magnification TEM image of 830/DETA with 2% weight clay loading,

    showing mostly intercalated nanocomposite structure (195,000 magnification).

    Fig. 18. Low magnification TEM image of 830/DETA with 0.5% weight clay loading,

    showing typical size of a clay aggregate (18,500 magnification).

    Fig. 19. High magnification TEM image of 830/DETA with 0.5% weight clay

    loading, showing mostly an intercalated nanocomposite structure (195,000

    magnification).

    0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    0

    Clay Loading (wt %)

    Young'sModulus(MPa)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 20. Variation of Youngs modulus for 815C-DETA at different clay loadings.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300294

  • 7/28/2019 1-s2.0-S0143749611000194-main

    10/15

    yielding, and consequently increased rupture strain and

    toughness [29,30].

    Varying degrees of intercalated and exfoliated nanocompo-

    site areas seem to form in both the samples with 0.5% and

    2% weight clay. It is reasonable to assume that this

    trend occurs for systems at other clay loadings as well, so

    various results for mechanical and barrier properties are

    expected.

    50

    55

    60

    65

    70

    75

    0

    Clay Loading (wt %)

    Max

    Stress

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 21. Variation of maximum stress for 815C-DETA at different clay loadings.

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0.14

    0.16

    Clay Loading (wt %)

    MaxStrain(m/m)

    0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 22. Variation of maximum strain for 815C-DETA at different clay loadings.

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0.14

    0

    Clay Loading (wt %)

    Strain

    atMaximumS

    tress(m/m)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 23. Variation of strain at maximum stress for 815C-DETA at different clay

    loadings.

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    45.0

    50.0

    0

    Clay Loading (wt %)

    YieldS

    tress

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 24. Variation of offset yield stress for 815C-DETA at different clay loadings.

    0

    1

    2

    3

    4

    5

    6

    7

    8

    0

    Clay Loading (wt %)

    Toug

    hness

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 25. Variation of toughness, measured as an area under the stressstrain curve

    for 815C-DETA, at different clay loadings.

    Fig. 26. Digital photograph of 815C-DETA-30B tensile tests specimens at various

    clay loadings. The samples were stressed to failure, with the 0% sample showing

    some necking and the 1% sample showing stress whitening and some necking, as

    observed at upper failure areas in the inset photos.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 295

  • 7/28/2019 1-s2.0-S0143749611000194-main

    11/15

    The higher viscosity 830-DETA epoxy system was also ana-

    lyzed with X-ray diffraction and TEM and the results are shown in

    Figs. 1519. Examination of these figures reveals the following

    findings:

    Compared to the 815C X-ray findings, the 830 system showed

    less dramatic of a leftward shift in peaks as the weight

    percentage was decreased as seen in Fig. 15. The best results

    seem to be from the 0.5% clay loaded system as it has a more

    featureless curve (Fig. 15d), but an indication of a peak near

    5.04 nm suggests that the degree of exfoliation is not as good

    as in the 0.5% loaded 815C system (Fig. 10d). Although the

    0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    0

    Clay Loading (wt %)

    Young's

    Mo

    du

    lus

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 27. Variation of Youngs modulus for 830-DETA at different clay loadings.

    50

    55

    60

    65

    70

    75

    80

    85

    0

    Clay Loading (wt %)

    Max

    Stress

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 28. Variation of maximum stress for 830-DETA at different clay loadings.

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0.14

    0.16

    0

    Clay Loading (wt %)

    Max

    Stra

    in(m/m)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 29. Variation of maximum strain for 830-DETA at different clay loadings.

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0.14

    0.16

    0

    Clay Loading (wt %)

    StrainatMa

    ximumS

    tress(m/m)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 30. Variation of strain at maximum stress for 830-DETA at different clay

    loadings.

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    45.0

    0

    Clay Loading (wt %)

    YieldStress

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 31. Variation of offset yield stress for 830-DETA at different clay loadings.

    0

    1

    2

    3

    4

    5

    6

    7

    8

    9

    0

    Clay Loading (wt %)

    Toug

    hness

    (MPa

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 32. Variation of toughness, measured as an area under the stressstrain curve

    for 830-DETA, at different clay loadings.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300296

  • 7/28/2019 1-s2.0-S0143749611000194-main

    12/15

    system does appear to intercalate well, increasing the clay

    layer distance up to 5.04 nm, curves would also seem to

    indicate that exfoliation is less likely or at least more difficult

    in the higher viscosity system.

    The 830 system followed the same trend as the 815C system,

    with the lower loading of clay resulting in a smaller aggregate

    size as seen by comparing Figs. 18 and 16.

    Higher magnification TEM images showed a lesser degree of

    exfoliation and more areas of intercalated nanocomposite

    formation at both 0.5% and 2% clay loading levels in the Epon

    830 system. This was expected for the much higher viscosity

    resin. Magnified images of typical intercalated structures can

    be seen in Fig. 17 for the system with 2% weight clay and

    in Fig. 19 for the 0.5% loaded system. In the 815C system,

    having less clay particles in smaller aggregates worked to an

    Displacement (mm)

    CrackP

    ropagationLoad(kN)

    vertical displacement

    during crack propagation

    Fracture toughness

    approximated by the

    load at fracture

    Fig. 33. Typical load versus displacement plot showing values for fracture load

    and vertical displacement during crack propagation.

    Failure location

    Failure location

    Fig. 34. Digital photograph of a bonded granite double lap sample showing typical

    shearing failure of the granite substrate (a) and typical failure primarily in the

    adhesive layer (b).

    Table 8

    Load and displacement values for 815C-DETA-30B fracture samples (standard

    deviations shown in parentheses).

    Clay

    loading

    (weight

    %)

    Fracture load

    (kN)

    FMode I(kN)

    FMode

    II (kN)

    Vertical

    displacement

    during crack

    propagation

    (mm)

    Mode I

    displacement

    during crack

    propagation

    (mm)

    0 0.1787 (0.0095) 0.0381 0.0894 0.7833 (0.1259) 0.33380.5 0.2462 (0.0013) 0.0525 0.1231 1.1416 (0.0877) 0.4865

    1 0.2009 (0.0026) 0.0428 0.1005 0.9275 (0.0293) 0.3952

    2 0.1995 (0.0084) 0.0425 0.0997 0.8721 (0.1550) 0.3716

    4 0.1846 (0.0124) 0.0393 0.0923 0.9748 (0.0413) 0.4154

    Table 9

    Load and displacement values for 830-DETA-30B fracture samples (standard

    deviations shown in parentheses).

    Clay

    loading

    (weight

    %)

    Fracture load

    (kN)

    FMode I(kN)

    FMode

    II (kN)

    Vertical

    displacement

    during crack

    propagation

    (mm)

    Mode I

    displacement

    during crack

    propagation

    (mm)

    0 0.1843 (0.0052) 0.0393 0.0921 0.5915 (0.0506) 0.2521

    0.5 0.1926 (0.0229) 0.0410 0.0963 0.7527 (0.0622) 0.3207

    1 0.2127 (0.0177) 0.0453 0.1064 0.8474 (0.0375) 0.3611

    2 0.2349 (0.0273) 0.0500 0.1174 0.8695 (0.0335) 0.3705

    4 0.1256 (0.0131) 0.0268 0.0628 0.6691 (0.1836) 0.2851

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0

    Clay Loading (wt %)

    Forcea

    tca

    tas

    trop

    hiccrac

    kprop

    aga

    tion

    (kN)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 35. Variation of load at catastrophic crack propagation for various clayloadings in the 815C-DETA epoxy system.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 297

  • 7/28/2019 1-s2.0-S0143749611000194-main

    13/15

    improved exfoliation, but in this case the impact of aggregate

    size on exfoliation seemed to be less.

    3.2. Analysis of tensile properties

    Figs. 2025 show key mechanical properties for the Epon

    815C-DETA system loaded with various weight percentages of

    30B clay. All were cured at 120 1C for 3 h.X-ray and TEM seemed to indicate the best nanocomposite

    formation at 0.5% clay loading. The mechanical properties seem to

    agree with this; 0.5% organoclay loading on the 815C-DETA

    system had the following effects:

    The addition of clay increases Youngs modulus from

    1184 MPa for the neat resin to 1317 MPa ( Fig. 20).

    The maximum stress also increases from 67.49 to 73.19 MPa

    (Fig. 21).

    The addition of 0.5% weight clay embrittles the system. This

    decreases the maximum strain from 0.1148 to 0.1078 mm/mm

    (Fig. 22), but increases strain at maximum stress from 0.0823

    to 0.1038 mm/mm (Fig. 23).

    As expected, the more brittle system shows a higher yieldstress (45.5 MPa, Fig. 24) but a slightly lower toughness

    (5.25 MPa, Fig. 25), measured as area under the stressstrain

    curve.

    The addition of 1% weight 30B clay seems to cause some

    interesting property changes in the epoxy system:

    An increase in rupture strain from 0.1148 to 0.1405 mm/mm

    and corresponding increase in toughness from 5.38 to

    7.30 MPa accompanies this nanocomposite (Fig. 25). The

    reason for this is attributed to the shear yielding and toughen-

    ing sometimes observed in clay nanocomposites, retaining

    some orientation of the clay layers [29,30]. This is further

    confirmed by the presence of stress whitening and somenecking in the 1% sample as seen in Fig. 26.

    When the clay serves to aid in shear yielding, it cannot act to

    embrittle the system as in the 0.5% loaded sample. Further-

    more, no significant increase is observed in the modulus or

    yield stress for the 1% sample (Figs. 20 and 24, respectively),

    and a less dramatic increase in maximum stress from 67.49 to

    70.94 MPa (Fig. 21).

    The addition of clay greater than 1% seems to have no positive

    effect on mechanical properties. In fact, all key properties except

    for Youngs modulus seem to steadily decrease with the increased

    addition of clay. The Youngs modulus values show an initial

    increase at 0.5% clay loading, but they drop down to a level

    approximately equal to that for the neat resin at higher clayloadings (Fig. 20). It is possible that adding high amounts of

    nanoclay increases the organic content to the level, where it acts

    to plasticize the epoxy network and decrease the crosslink

    density. Thus, an optimal point of clay loading is observed for

    mechanical properties, and in the 815C-DETA-30B system this

    appears to be around 0.5% clay loading.

    3.2.1. The effect of resin viscosity

    The effectiveness of nanocomposite formation was explored

    on an epoxy system of much higher viscosity (Epon 830-DETA).

    Essentially, the two epoxy systems are chemically very similar,

    with the viscosity reduction of 815C achieved by the addition

    of reactive diluent of butyl glycidyl ether. This small reactive

    molecule effectively reduces the viscosity from around 170225 P

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    0

    Clay Loading (wt %)

    Verticaldisplacementduringcrack

    propagation(mm)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 36. Variation of vertical wedge displacement during crack propagation for

    various clay loadings in the 815C-DETA epoxy system.

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0

    Clay Loading (wt %)

    For

    ceatcatastrophiccrackpropagation(kN)

    0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

    Fig. 37. Variation of load at catastrophic crack propagation for various clay

    loadings in the 830-DETA epoxy system.

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    0

    Clay Loading (wt %)

    Vert

    ica

    ldisp

    lacemen

    tdu

    ringcrac

    k

    propaga

    tion

    (mm

    )

    0.5 1 1.5 2 2.5 3 3.5 4 4.5

    Fig. 38. Variation of vertical wedge displacement during crack propagation for

    various clay loadings in the 830-DETA epoxy system.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300298

  • 7/28/2019 1-s2.0-S0143749611000194-main

    14/15

    (Epon 830) to 57 P (Epon 815C) and seems to aid in the

    intercalation process as the nanocomposites elastic behavior

    seems to have been extended with the lower viscosity resin

    system. The mechanical properties of the higher viscosity

    830-DETA-30B system are plotted in Figs. 2732. Examination

    of these figures reveals the following findings:

    The higher viscosity 830-DETA system showed less elastic

    behavior compared to the 815C-DETA system at every clayloading level with lower yield stress for the Epon 830 system

    compared to the Epon 815C.

    The higher viscosity system showed higher modulus, higher

    maximum stress, and higher rupture strain at nearly every clay

    loading level.

    The higher viscosity Epon 830 system had the best modulus at

    a clay loading of 0.5% and the best yield stress at 2% clay

    loading (Figs. 27 and 31, respectively).

    In comparison to the neat sample, toughness (measured as an

    area under the stressstrain curve) decreased at all clay levels

    with only the 2% weight sample providing toughness value

    close to that for the neat sample (Fig. 32).

    The overall impact of the clay on key mechanical properties of

    the 830-DETA-30B systems was less than the impact it had on

    the key properties of the 815C-DETA-30B system. This is

    attributed to the less effective intercalation and exfoliation

    (and thus less effective nanocomposite formation) in the 830

    system, due to higher viscosity and a higher molecular weight.

    3.3. Assessment of fracture results

    Different configurations of the double lap fracture test speci-

    men (Fig. 4) were investigated to ensure failure primarily through

    the adhesive layer. A crack length of 7 mm was used and the

    overlap length was varied from 15 to 12 mm, resulting in load

    versus displacement plots of the type shown in Fig. 33. At larger

    overlap lengths, shear failure of the granite stone substrate was

    observed (see Fig. 34a), so the overlap length was reduced until

    failure through epoxy was observed (Fig. 34b). The optimal

    overlap length of 12 mm was then used on the two epoxy systems

    (815C and 830) at various clay loadings to observe the effect of

    clay on the fracture toughness of the epoxy bonded to granite

    (approximated by the load at fracture). The fracture test results

    are tabulated in Tables 8 and 9. The tables include the fracture

    load and vertical wedge displacement during crack propagation

    (see Fig. 33) as well as the Mode I displacement perpendicular to

    the crack propagation and the force components (FM ode I and

    FMode II as seen in Fig. 4) on the granite substrate cover plates. The

    load at catastrophic crack propagation and the vertical wedge

    displacement during crack propagation for the 815C-DETA system

    are plotted in Figs. 35 and 36 and in Figs. 37 and 38 for the 830-

    DETA system. Based on these figures the following observationscan be made:

    All systems showed brittle failure through the epoxy layer.

    The fracture tests seem to reveal results analogous to the

    tensile tests. With the higher viscosity 830-DETA resin system,

    the advantage of clay addition seems to be less in comparison

    to the similarly filled lower viscosity system (Figs. 35 and 37).

    The load at failure and total vertical displacement during crack

    propagation was highest for the 815C-DETA system (Figs. 35

    and 36) at 0.5% 30B clay loading. This clay loading also

    provided the largest modulus, maximum stress, and yield

    stress in the tensile testing (Figs. 20, 21, and 24), but not the

    highest toughness (Fig. 25). This also indicates that although

    the clay addition provided maximum toughening properties in

    the tensile tests with 1% by weight in 815C-DETA adhesive, it

    may not be effective as a crack stopping toughening agent at

    this loading level when bonding porous stone substrates.

    The addition of clay to the 830 system again seems to show

    less of an effect in the higher viscosity resin. With the best

    failure load and maximum vertical displacement being

    observed around 2% clay loading. This correlates the tensile

    test data, with the clay loading levels for high values of

    modulus, maximum stress, yield stress and toughness(Figs. 27, 28, 31 and 32) matching the clay loading levels for

    maximums in the fracture data (Figs. 37 and 38).

    The fracture results reveal the 815C system with larger values

    for vertical displacement at nearly every clay loading level

    compared to the 830 system.

    The difference in viscosity between the two systems seemed to

    have minimal effects on the macro scale. Penetration into the

    granite pours seem to be similar, probably due to the high

    temperature cure causing a reduction in viscosity to a similar

    level for both systems at early stages of curing.

    4. Conclusions

    Between the higher and lower viscosity resins, the betterlayered silicate nanocomposite performance was seen with the

    lower viscosity 815C resin. X-ray and TEM images confirm the

    levels of exfoliation for each resin and at each clay loading. The

    optimal level of nanoclay loading seems to be around 0.5% for

    Epon 815C and 2% for Epon 830 for mechanical properties. The

    most noticeable improvements in mechanical properties for the

    815C system were found to be maximum stress, elastic modulus,

    and yield stress. Increased toughness and stress whitening at 1%

    loading are evidence that the clay can act as a shear-yielding

    toughening agent in this epoxy system. Nanocomposite formation

    in the 830 resin appears to be hindered by its higher viscosity

    compared to the 815C, showing dramatic improvements in only

    yield stress and a slight improvement in maximum stress.

    Fracture analysis of the samples seemed to reveal results

    similar to those from tensile testing. The addition of clay

    improved the fracture load for the samples similar to the way

    maximum stress was increased in tensile properties for both Epon

    815C and 830 resin-based nanocomposites. One inconsistency

    was that for the lower viscosity resin Epon 815C-based nano-

    composite, the fracture toughness, which we approximated by

    load at fracture, did not show a significant increase at the 1% clay

    level as in the case of tensile toughness. This suggests that, when

    the lower viscosity resin is used, while the clay may be an

    effective toughening agent in regard to shear yielding with 1%

    loading, its ability to stop cracks in Mode I situations is less

    effective at the same level of clay loading. TEM imaging reinforced

    this idea revealing areas of intercalation, which may be sites of

    shear yielding. The fracture results showed failure of the

    mechanically interlocked adhesive layer. Therefore, we believethat the increase in the bulk adhesive strength provided by the

    addition of clay resulted in improved adhesive joints of brittle,

    porous (granite) substrates.

    References

    [1] Vaia RA, Giannelis EP. Lattice model of polymer melt intercalation inorganically-modified layered silicates. Macromolecules 1997;30:79909.

    [2] Vaia RA, Giannelis EP. Polymer melt intercalation in organically-modi-fied layered silicates: model predictions and experiment. Macromolecules1997;30:80009.

    [3] Lan T, Pinnavaia TJ. Clay-reinforced epoxy nanocomposites. Chem Mater1994;6:22169.

    [4] Gilman JW, Kashiwagi T. Nanocomposites: a revolutionary new flame

    retardant approach. SAMPE J 1997;33:406.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300 299

  • 7/28/2019 1-s2.0-S0143749611000194-main

    15/15

    [5] Lan T, Kaviratna PD, Pinnavaia TJ. Mechanism of clay tactoid exfoliation inepoxy-clay nanocomposites. Chem Mater 1995:214450.

    [6] Lan T, Wang Z, Shi H, Pinnavaia TJ. Clayepoxy nanocomposites: relationshipsbetween reinforcement properties and the extent of clay layer exfoliation.Polym Mater Sci Eng 1995;73:2967.

    [7] Shi H, Lan T, Pinnavaia TJ. Interfacial effects on the reinforcement propertiesof polymerorganoclay nanocomposites. Chem Mater 1996;8:15847.

    [8] Wang Z, Massam J, Pinnavaia TJ. Epoxyclay nanocomposites. In: PinnavaiaTJ, Beall GW, editors. Polymerclay nanocomposites. Hoboken, New Jersey:

    John Wiley & Sons Ltd.; 2000. p. 12749.[9] Wang Z, Pinnavaia TJ. Hybrid organicinorganic nanocomposites: exfoliation

    of magadiite nanolayers in an elastomeric epoxy polymer. Chem Mater1998;10:18206.

    [10] Wang Z, Lan T, Pinnavaia TJ. Hybrid organicinorganic nanocompositesformed from an epoxy polymer and a layered silicic acid (magadiite). ChemMater 1996;8:22004.

    [11] Rana HT. Moisture diffusion through neat and glass-fiber reinforced vinylester resin containing nanoclay. M.S. Thesis Department of Chemical Engi-neering, West Virginia University 2003.

    [12] Rana HT, Gupta RK, Gangarao HVS, Sridhar LN. Measurement of moisturediffusivity through layered-silicate nanocomposites. AICHE J 2005;51:324956.

    [13] Sancaktar E, Kuznicki J. Stress-induced reduction of water uptake in clay-reinforced epoxy nanocomposites. Curr Nanosci 2006;2:3517.

    [14] Park JH, Jana SC. Mechanism of exfoliation of nanoclay particles in epoxyclay nanocomposites. Macromolecules 2003;36:275868.

    [15] Park JH, Jana SC. A case study on the effects of plasticization of epoxynetworks by organic treatment on exfoliation of nanoclay. Macromolecules2003;36:83917.

    [16] Park JH, Jana SC. The relationship between nano- and micro-structures andmechanical properties in PMMAepoxynanoclay composites. Polymer2003;44:2091100.

    [17] Giannelis EP. Polymer layered silicate nanocomposites. Adv Mater 1996;8:2935.

    [18] Messersmith PB, Giannelis EP. Synthesis and characterization of layered

    silicateepoxy nanocomposites. Chem Mater 1994;6:171925.[19] Messersmith PB, Giannelis EP. Synthesis and barrier properties of poly(e-

    caprolactone)-layered silicate nanocomposites. J Polym Sci, Part A: Polym

    Chem 1995;33:104757.[20] Messersmith PB, Giannelis EP. Polymer-layered silicate nanocomposites: in-

    situ intercalative polymerization of e-caprolactone in layered silicates. Chem

    Mater 1993;5:10646.[21] Lee JY, Baljon ARC, Loring RF, Panagiotopoulos AZ. Modeling intercalation

    kinetics of polymer silicate nanocomposites. Mater Res Soc Symp Proc

    1999;543:13135.[22] Krishnamoorti R, Giannelis EP. Rheology of end-tethered polymer layered

    silicate nanocomposites. Macromolecules 1997;30:4097102.[23] Vaia RA, Ishii H, Giannelis EP. Synthesis and properties of two-dimensional

    nanostructure by direct intercalation of polymer melts in layered silicates.

    Chem Mater 1993;5:16946.[24] Vaia RA, Teukolsky RK, Giannelis EP. Interlayer structure and molecular

    environment of alkylammonium layered silicates. Chem Mater 1994;6:

    101722.[25] Sancaktar E, Aussawasathien D. Nanocomposites of epoxy with electrospun

    carbon nanofibers: mechanical behavior. J Adhes 2009;85:16079.[26] Sancaktar E, Aussawasathien D. Electrospun polyacrylonitrile-based carbon

    nanofibers and their silver modifications: surface morphologies and proper-

    ties. Curr Nanosci 2008;4:1307.[27] E. Sancaktar, D. Aussawasathien, Effect of Non-Woven Carbon Nanofiber Mat

    Presence on Cure Kinetics of Epoxy Nanocomposites, Macromol Symp 264

    (2008) 2633.[28] /http://webstore.ansi.org/RecordDetail.aspx?sku=ASTM+ D3931-08S; 2010.[29] Liu T, Tjiu WC, Tong Y, He C, Goh SS. Morphology and fracture behavior of

    intercalated epoxy/clay nanocomposites. J Appl Polym Sci 2004;94:123644.[30] Ratna D, Becker O, Krishnamurthy R, Simon GP, Varley RJ. Nanocomposites

    based on a combination of epoxy resin, hyperbranched epoxy and a layered

    silicate. Polymer 2003;44:744957.

    E. Sancaktar, J. Kuznicki / International Journal of Adhesion & Adhesives 31 (2011) 286300300

    http://webstore.ansi.org/RecordDetail.aspx?sku=ASTM+D3931-08http://webstore.ansi.org/RecordDetail.aspx?sku=ASTM+D3931-08http://webstore.ansi.org/RecordDetail.aspx?sku=ASTM+D3931-08http://webstore.ansi.org/RecordDetail.aspx?sku=ASTM+D3931-08