Anda di halaman 1dari 12

Research Article

Received: 18 June 2009 Revised: 8 October 2009 Accepted: 2 November 2009 Published online in Wiley Interscience: 7 April 2010

(www.interscience.wiley.com) DOI 10.1002/pi.2817

Surface morphology, thermomechanical and barrier properties of poly(ether sulfone)-toughened epoxy clay ternary nanocomposites
Abdul Azeez Asif, Bibin John, Vattikuty Lakshmana Rao and Kovoor Ninan Ninan
Abstract
Poly(ether sulfone) (PES)-toughened epoxy clay ternary nanocomposites were prepared by melt blending of PES with diglycidyl ether of bisphenol A epoxy resin along with Cloisite 30B followed by curing with 4,4 -diaminodiphenylsulfone. The effect of organoclay and thermoplastic on the fracture toughness, permeability, viscoelasticity and thermomechanical properties of the epoxy system was investigated. A signicant improvement in fracture toughness and modulus with reduced coefcient of thermal expansion (CTE) and gas permeability were observed with the addition of thermoplastic and clay to the epoxy system. Scanning electron microscopy of fracture-failed specimens revealed crack path deection and ductile fracture without phase separation. Oxygen gas permeability was reduced by 57% and fracture toughness was increased by 66% with the incorporation of 5 phr clay and 5 phr thermoplastic into the epoxy system. Optical transparency was retained even with high clay content. The addition of thermoplastic and organoclay to the epoxy system had a synergic effect on fracture toughness, modulus, CTE and barrier properties. Planetary ball-milled samples gave exfoliated morphology with better thermomechanical properties compared to ultrasonicated samples with intercalated morphology. c 2010 Society of Chemical Industry Keywords: fracture toughness; gas permeability; organoclay; rheology

INTRODUCTION
In the recent past, extensive research has been carried out on layered silicate epoxy nanocomposites including the processing, preparation and exfoliation mechanisms1 12 and the incorporation of these epoxy nanocomposites as a matrix into traditional carbon bre-reinforced composites.13,14 Numerous reports in the literature describe the improvement in barrier properties, tensile strength, modulus, glass transition temperature and toughness, and a decrease in ammability and residual stress by the incorporation of 210 wt% of organoclay into epoxy resins.1 12 Fully exfoliated morphologies, where individual silicate layers are homogeneously and uniformly dispersed throughout the polymer matrix, are most preferred for the enhancement of such properties. The main reasons for the improved properties of nanocomposites is the very large interfacial interaction between matrix and layered silicates, the high aspect ratio of the dispersed clay particles and the hindered tortuous diffusing paths created by the dispersed silicate nanolayers. Layered silicate nanocomposites are becoming attractive in cryogenic storage tanks based on the results reported by the NASA Glenn Research Center15 after the failure of X-33 composite fuel tanks due to microcracking.16 There are many factors that inuence the morphology of nanocomposites; whether phase-separated, intercalated or exfoliated depends upon the type of resin,17 curing agent (curative) used,1,18 rate of curing, methods of preparation (melt mixing, solution mixing or in situ polymerization)9 and processing method.11,12

The addition of clay particles can also have an effect on the kinetics of the cure process. The addition of clay often leads to an increased rate of gelation.19,20 Zunjarrao et al.11 reported that high-shear mixing resulted in better dispersion of clay particles and better mechanical properties over ultrasonication, even if both methods gave an exfoliated morphology. It was reported that 4,4 -diaminodiphenylsulfone (DDS)-cured epoxy nanocomposites showed an exfoliated morphology with improved mechanical properties when dispersed by exerting shear force on epoxy montmorillonite solution using ball milling.12 Liu et al.21 reported an increase of 65% in fracture toughness for epoxy/clay nanocomposites with the incorporation of 4 wt% Cloisite 93A (d001 = 25.8 ). Also, an increase in fracture toughness of more than 100% was observed by us22,23 and others24,25 for thermoplastic-toughened epoxy systems. The addition of thermoplastic into epoxy clay nanocomposites provides substantial improvement in fracture toughness. Not

Correspondence to: Vattikuty Lakshmana Rao, Adhesives and Advanced Matrix Resin Section, Propellants and Special Chemicals Group, Propellants, Polymers, Chemicals and Materials Entity, Vikram Sarabhai Space Centre, Trivandrum695022, India. E-mail: v lakshmanarao@vssc.gov.in Adhesives and Advanced Matrix Resin Section, Propellants and Special ChemicalsGroup,Propellants,Polymers,ChemicalsandMaterialsEntity,Vikram Sarabhai Space Centre, Trivandrum-695022, India

986

Polym Int 2010; 59: 986997

www.soci.org

c 2010 Society of Chemical Industry

PES-toughened epoxy clay ternary nanocomposites many reports are available on thermoplastic-toughened epoxy clay nanocomposites in the open literature. Frohlich et al.26 reported an improved toughness in rubber-toughened hybrid epoxy nanocomposites compared with nanocomposites without rubber, but the glass transition temperature (Tg ) was lowered drastically. Balakrishnan et al.27 reported that the addition of both elastomers and nanoclay to an epoxy resin resulted in an enhanced ductility and improved toughness without compromising strength and modulus. Isik et al.28 studied the mechanical properties of epoxypolyether polyol organically treated montmorillonite nanocomposites with respect to polyether polyol and clay content. They observed an increased Tg and Youngs modulus with respect to clay content. Peng et al.29 reported the inuence of organically modied clay on the reaction-induced phase separation behaviour of epoxy/poly(ether imide) ternary nanocomposites. Recently, we have reported the toughening by a hydroxyl-terminated poly(ether ether ketone) with pendant methyl groups (PEEKMOH) of epoxy clay ternary nanocomposites based on quaternary alkyl ammonium-modied montmorillonite without polar functional groups (Cloisite 25A)30 and montmorillonite with functional alkyl ammonium modier (Cloisite 30B).31 In the present study, we report poly(ether sulfone) (PES)toughened epoxy clay ternary nanocomposites based on montmorillonite with functional alkyl ammonium modier (Cloisite 30B). We studied the effect of PES and processing techniques on phase morphology, surface morphology, fracture toughness, barrier properties and thermomechanical properties of the epoxy clay ternary nanocomposites.

www.soci.org

Epoxy resin

1. PES (5phr) at 180C

Epoxy-PES blend 1. Cloisite 30B (1-8 phr) 2. 80C, mechanical stirring 2 hours 3. 80C ultrasonication / Planetary ball milling Epoxy-PES-clay ternary mixture 1. DDS at 180C, mechanical stirring 2. Evacuation Epoxy-PES-clay-DDS ternary mixture (Uncured) 1. 180C for 3 hours 2. 200C for 2 hours Epoxy-PES-clay ternary nanocomposite
Scheme 1. Flow chart for the preparation of PES-toughened epoxy clay ternary nanocomposite.

EXPERIMENTAL
Materials Diglycidyl ether of bisphenol A epoxy resin (LY 556, Ciba Geigy) with an epoxide equivalent weight of 188.68, DDS (Purity: 99%) (Merck) and PES having a Tg value of 201 C were used as received. Cloisite 30B (Southern Clay Product; methyl tallow(bis-2hydroxyethyl) quaternary ammonium-modied montmorillonite, d001 = 1.85 nm, Cation Exchange Capacity (CEC) = 95 meq (100 g clay)1 , specic gravity = 1.98 g cm3 ) was dried under vacuum at 70 C before use. Preparation of PES-toughened epoxy clay ternary nanocomposites PES-toughened epoxy clay ternary nanocomposites were prepared as shown in Scheme 1. For all compositions, PES concentration was maintained at 5 phr. The preparation methods included two processing techniques: planetary ball milling and ultrasonication. A Fritsch Planetary Mono Mill Pulverisette 6 with 10 mm diameter balls (2.98 g each) of zirconium dioxide material at 500 rpm was used for planetary ball milling. Ultrasonication was performed using an Elma ultrasonic bath (Elma Transonic, TI-H-5) at a frequency of 35 kHz and a power of 100 W. Characterization Fourier transform infrared (FTIR) spectra of cured and uncured samples in KBr pellets were recorded using a PerkinElmer Spectrum GXA FTIR spectrometer. The curing behaviour of PES-toughened epoxy clay DDS ternary mixtures was studied using DSC operated in standard mode with a TA Instruments model DSC-2920 from room temperature to 350 C at 10 C min1 in nitrogen atmosphere. Isothermal rheological experiments were carried
Polym Int 2010; 59: 986997

out at 180 C using a Rheologica Stresstech rheometer, model Rheologic Viscotech QC, with a parallel plate assembly operated in oscillation mode at a frequency of 1 Hz and controlled strain of 0.01. Wide-angle XRD data were obtained with powder samples of the Cloisite 30B and nanocomposites using a PANalytical model X-ray diffractometer with Xpert pro software and nickel-ltered Cu K radiation at 30 kV and 20 mA. The morphology of cryogenically fractured tensile and fracture-failed samples was analysed using a Philips XL 20 SEM instrument. The failed surfaces were etched with dimethylsulfoxide for 24 h to remove the thermoplastic phase. The specimens were dried in a vacuum oven at 100 C overnight to remove the solvent. All the specimens were sputter-coated with gold before obtaining the micrographs. A thermomechanical analyser (PerkinElmer TMA-7) was used to determine the coefcient of thermal expansion (CTE) and Tg . Scans were run at a heating rate of 10 C min1 from room temperature to 250 C under nitrogen atmosphere. The thermal stability of the nanocomposites was analysed using TGA with a TA Instruments model SDT 2960 thermal analyser. The samples were heated from room temperature to 900 C at a heating rate of 10 C min1 in nitrogen atmosphere. The viscoelastic properties of the nanocomposites were measured using a TA Instruments DMA 2980 dynamic mechanical thermal analysis (DMTA) instrument. DMTA specimens of size 50 10 3 mm3 were cut from a 3 mm thick cured laminate of nanocomposite using a diamond cutter. The analysis was done in three-point bending mode at a frequency of 1 Hz from room temperature to 300 C at a heating rate of 2 C min1 . Experiments on the permeation of oxygen gas were carried out using a LYSSY (model 1005000) manometer gas permeability tester with measuring range 110 000 mL m2 day1 and ow rate of 10 L h1 at room temperature. Fracture toughness and tensile and exural properties were determined as per ASTM 5045, ASTM D638 (Type V) and ASTM D790, respectively. The measurements were done using a universal

987

c 2010 Society of Chemical Industry

www.interscience.wiley.com/journal/pi

www.soci.org

A Asif et al.

Initial

914

180C / 1 h

%T

180C / 3 h

180C / 3 h + 200C / 2 h

4000.0

3000

2000 cm-1

1500

1000

500 400.0

Figure 1. FTIR spectra of uncured sample and various cured samples.

testing machine (model TNE 5000) at a crosshead speed of 10 mm min1 . Rectangular specimens of 100 10 3 mm3 were used for determining exural strength. Flexural modulus was determined from the slope of the initial portion of the exural stressstrain curve. Flexural strength was calculated using32 Flexural strength = 3Pl 2bd2 (1)

Table 1. Tg values of the epoxy/PES/DDS/clay mixture cured at various temperatures and for various times Cure temperature and time 180 C/1 h 180 C/3 h 180 C/3 h + 200 C/2 h Tg ( C) 151 192 208

where P is the load at break, l is the span length and b and d are the breadth and thickness of the specimen, respectively. Dogbone-shaped specimens were used for the evaluation of tensile properties. Single edge notch specimens of 46 6 3 mm3 (span length = 24 mm) were used to measure the fracture toughness of the epoxy clay ternary nanocomposites. A notch of 2.7 mm was made at one edge of the specimen. A natural crack was made by pressing a fresh razor blade into the notch. The fracture toughness was expressed as stress intensity factor (KIC ) calculated using33 L KIC = f (x) (2) BW 1/2 where 0 < x < 1 and f (x) = 6x 1/2 1.99 (1 x)(2.15 3.93x 2.7x )
2

0.4 0.3 0.2 Heat flow (W g 1) 0.1 0.0 -0.1 -0.2 -0.3

(1 + 2x)(1 x)3/2

-0.4 0 50 100 150 200 250 300 350 Temperature (C)

and L is the load at crack initiation, B is the specimen thickness, W is the specimen width, a is the crack length and x = a/W.

0h 180 / 1 h 180 / 3 h 180 / 3 h + 200 / 2 h


Figure 2. DSC traces of epoxy/DDS/PES/clay mixture.

RESULTS AND DISCUSSION


FTIR studies The FTIR spectra of the nanocomposites (before curing and after curing at various time intervals) are shown in Fig. 1. The epoxy peak at 914 cm1 reduces in intensity for the samples cured at 180 C for 1 h and it totally disappears for samples cured for 3 h and for samples post-cured at 200 C for 2 h conrming the

completion of the cure reaction. Post-curing is performed in order to ensure full curing, as evidenced from Tg values obtained from DSC (Table 1).
Polym Int 2010; 59: 986997

988

www.interscience.wiley.com/journal/pi

c 2010 Society of Chemical Industry

PES-toughened epoxy clay ternary nanocomposites

www.soci.org

Table 2. CTE and Tg values of the ternary nanocomposites Coefcient of thermal expansion (106 C1 ) Composition Epoxy Epoxy/5 phr PES Epoxy/5 phr PES/1 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B Epoxy/5 phr PES/5 phr Cloisite 30B Epoxy/5 phr PES/8 phr Cloisite 30B 30200 C 102 3 100 2 98 2 96 1 91 3 96 2 TMA 199 197 198 199 199 199 Tg ( C) DMTA (from tan ) 221 216 214 214 209 205 DSC 212 209 208 208 205 204

Storage Modulus (MPa)

0.05

0.04

0.03

0.02

0.01

0.00

-0.01 1400

1600

1800

2000

2200

2400

2600

5000 Time (s)

10000

Epoxy Epoxy/5 phr PES Epoxy/3 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B
Figure 3. Storage and loss moduli as a function of time at 180 C for the nanocomposites.

DSC studies The cure behaviour of PES-toughened epoxy clay DDS mixture (before curing and after curing at various time intervals) is shown in Fig. 2. A sharp exotherm is observed for uncured samples with initial cure temperature at 160 C and maximum cure temperature at 224 C indicates the epoxy amine reaction. As curing proceeds at 180 C for 3 h, the exothermic peak totally disappears. However, post-curing at 200 C for another 2 h results in high Tg (Table 1). Thus, increasing the temperature in the post-curing stage makes the reaction more favourable and allows the polymerization reaction to be completed since the curing process becomes diffusion controlled at a high degree of conversion of epoxy. Tg of the epoxy system (Table 2) is found to decrease with an increase in clay concentration, which may be due to the plasticizing effect of organic moieties of clay particles thereby decreasing the crosslink density of the epoxy network.
Polym Int 2010; 59: 986997

Rheological characterization The application of thermoset-based nanocomposites is highly dependent upon their processproperty relationships. The type of resin, curing agent, nanoparticles, cure kinetics and dispersion techniques will play a major role during gelation and vitrication stages in the nal morphology of the crosslinked structure.34 Thermoset polymers are usually cured using isothermal cure cycles. The effects of isothermal curing on the intergallery spacings of epoxy/layered silicate nanocomposites investigated using XRD have been reported by several authors. These reports revealed that, once the clay is swollen with the epoxy-curing agent prepolymer, no further expansion occurs until the temperature of the cure is increased to the onset of cure.1,34 The effect of thermoplastic and clay on the rheology of epoxy clay nanocomposites isothermally cured at 180 C is shown in Fig. 3. There is a rapid increase in storage modulus at gelation and a levelling off as vitrication occurs irrespective of the composition. It is further observed that storage modulus is lower for the clay-lled

Loss Modulus (MPa)

989

c 2010 Society of Chemical Industry

www.interscience.wiley.com/journal/pi

www.soci.org
Epoxy/5phr PES

A Asif et al.

40 35 30 Time (min) 25 20 15 10 5 0

Epoxy

Epoxy/ 5phr PES/ 3phr Cloisite 30B

400

Intensity

Epoxy/3phr Cloisite 30B

200

0 0 10 2 () Cloisite 30B Epoxy/3 phr Cloisite 30B Epoxy/5 phr PES /1 phr Cloisite 30B Epoxy/5 phr PES /3 phr Cloisite 30B Epoxy/5 phr PES /3 phr Cloisite 30B (Planetary ball milling)
Figure 5. XRD patterns of Cloisite 30B and various nanocomposites.

Composition

20

30

Figure 4. Gel time versus composition of epoxy systems.

system compared to neat resin and thermoplastic-toughened epoxy resin at all time intervals. Similar observations are noticed for loss modulus versus time for the epoxy clay nanocomposites (Fig. 3). The effect of the thermoplastic/clay on the gel time can be studied by plotting the storage modulus (G ) and loss modulus (G ) as function of time at 180 C for the epoxy system. The G /G crossover in Fig. 3 is taken as the gel point for these systems and is shown as a function of thermoplastic and clay in Fig. 4. The gel time of the epoxy system decreases from 36 to 24 min on addition of 3 phr clay. The reduction in gel time is due to the catalytic effect of clay on the curing reaction. Similar observations were made by several authors for epoxy and cyanate ester systems.34,35 During the curing process the diffusion of the unreacted prepolymer into the intergalleries of the clay layers followed by polymerization play a vital role in the ultimate morphology of the system. It is generally believed that the relative rate of extragallery and intergallery polymerization must be balanced in order to obtain an exfoliated structure.36 38 This balance depends on several factors, including prepolymer viscosity and the nature of silicates (i.e. CEC, intergallery polarity).1,34,39 The gel time of the epoxy system increases from 36 to 41 min on addition of 5 phr PES, which may be due to retardation of the cure reaction of the epoxy by the high molecular weight thermoplastic. However, the addition of both thermoplastic and clay to the epoxy system has a synergic effect on gel time. Wide-angle XRD To further investigate the morphology of the clay inside the ternary nanocomposites, the diffraction pattern of pristine Cloisite 30B was compared with those of nanocomposites having various clay fractions and nanocomposites processed with various processing techniques. Figure 5 reveals that the [001] diffraction peak of Closite 30B clay appears at 2 = 4.81 with an interlamellar spacing (d001 ) of 1.81 nm, as calculated from Braggs law: n = 2d sin (3)

nanocomposites. They found that intergallery spacing increases with an increase in cure temperature due to lower prepolymer viscosity and faster intergallery curing. Here we examined the effect of clay concentration and thermoplastic on the d spacing for the epoxy/curing agent/silicate systems. The diffraction peak of epoxy/3phr Cloisite 30B without thermoplastic modier shifts to 2 = 2.49 and the d spacing increases to 3.54 nm, since the epoxy/hardener matrix enters the clay galleries. The diffraction peak of epoxy/3phr Cloisite 30B nanocomposite with 5 phr PES modier is further shifted to 2 = 2.35 with an increase of d spacing to 3.75 nm. The increase in d spacing on incorporation of PES is not substantial. This may be due to the fact that high molecular weight PES (inherent viscosity = 0.38 dL g1 ) is expected to diffuse less in between the clay layers compared with low molecular weight constituents of epoxy (188 g mol1 ) and DDS (248 g mol1 ). Similar observations were made by us30,31 for PEEKMOH-toughened epoxy clay ternary nanocomposites and by Isik et al.28 for epoxypolyether polyol organically treated montmorillonite ternary nanocomposites. As the clay concentration increases from 1 to 3 phr in PES-toughened epoxy clay ternary nanocomposites, the diffraction angle peak shifts from 2 = 2.31 to 2.35 and the d spacing marginally decreases from 3.81 to 3.75 nm. However, the planetary ballmill dispersion technique facilitates an exfoliated morphology with total disappearance of the peak in the low-angle region for PES-toughened epoxy clay ternary nanocomposites. Haijun et al.12 also observed an exfoliated morphology for epoxy clay nanocomposites with better mechanical properties when dispersed by high-shear mixing using ball milling. Tensile, exural and fracture toughness properties Tensile, exural and fracture toughness properties of the PES-toughened epoxy and PES-toughened epoxy clay ternary nanocomposites are given in Table 3. The data reveal that the
Polym Int 2010; 59: 986997

where n is an integer, the wavelength, the glancing angle of incidence and d the interplanar spacing of clay layers. Dean et al.35 have reported the effect of clay concentration and curing temperature on the intergallery spacing for epoxy/silicate

990

www.interscience.wiley.com/journal/pi

c 2010 Society of Chemical Industry

PES-toughened epoxy clay ternary nanocomposites

www.soci.org

Table 3. Tensile, exural and fracture toughness properties of ternary nanocomposites Tensile strength (MPa) 60 4 80 6 64 3 61 4 60 5 65 2 58 4 50 5 Tensile modulus (GPa) 1.67 0.02 1.55 0.04 1.77 0.02 1.84 0.06 1.79 0.03 1.92 0.04 1.77 0.06 1.74 0.05 Elongation (%) 3.4 0.3 8 0.4 5.3 0.1 4.5 0.3 4.2 0.4 4.1 0.3 4.0 0.4 3.8 0.5 Flexural strength (MPa) 122 6 135 6 134 8 115 7 111 10 132 9 109 7 101 5 Flexural modulus (GPa) 2.95 0.11 2.96 0.12 3.03 0.03 3.09 0.02 3.20 0.01 3.59 0.05 3.24 0.06 3.47 0.10 Fracture toughness, KIC (MN m3/2 ) 1.16 0.16 1.62 0.19 1.53 0.07 1.54 0.11 1.65 0.10 1.93 0.14 1.45 0.12 1.32 0.08

Composition Epoxy Epoxy/5 phr PES Epoxy/5 phr PES/1 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B (1 h ultrasonication) Epoxy/5 phr PES/3 phr Cloisite 30B (1 h planetary ball milling) Epoxy/5 phr PES/5 phr Cloisite 30B Epoxy/5 phr PES/8 phr Cloisite 30B

tensile strength of the PES-toughened epoxy system is higher than that of the neat epoxy systems by 33%, which may be due to the plasticization of PES. But the tensile strength decreases with an increase in clay content at constant PES content (5 phr). This is because of the stress concentration effect of clay agglomerates at higher clay contents and thus the claypolymer surface interaction decreases. Another reason is that as the clay content increases, the viscosity of the system increases resulting in heterogeneity and nanovoid formation due to the entrapment of air bubbles during sample preparation.40 Similar observations have been reported for polyether polyol-modied epoxy montmorillonite ternary nanocomposites.28 The tensile and exural moduli of the ternary nanocomposites are found to increase with the incorporation of Cloisite 30B. An increase of 10 and 18.7% in the tensile modulus is observed for PES-toughened epoxy clay ternary nanocomposites with respect to neat epoxy and PES-toughened epoxy system, respectively, with 3 phr Cloisite 30B incorporation. Surprisingly the increase in tensile modulus is decreased to 5 and 12% with 8 phr Cloisite 30B incorporation. This is due to the fact that, as the clay concentration increases, the constraining effect of clay agglomerates inhibits the plastic deformation of the matrix as reported by Wang et al.41 Qi et al.42 observed a 15.1% improvement in tensile modulus on addition of 10% montmorillonite modied with methyl tallow (bis-2-hydroxyethyl) quaternary ammonium salt to diglycidyl ether of bisphenol A epoxy system. Similarly Basara et al.43 found a 17.2% enhancement in tensile modulus on addition of 7 wt% Cloisite 30B to the epoxy system. Isik et al.28 observed an increase in tensile modulus with respect to clay concentration for polyether polyol-modied epoxy montmorillonite ternary nanocomposites. The exural modulus is increased by 17 and 21% for 8 phr Cloisite 30B incorporation with respect to epoxy and PES-toughened epoxy matrix, respectively. The fracture toughness values of the nanocomposites are given in Table 3, which are expressed in terms of the stress intensity factor (KIC ). Fracture toughness values are found to be higher for PES-toughened epoxy clay ternary nanocomposites compared to neat epoxy resin. The enhancement in toughness of the nanocomposites can be attributed to the good dispersion of the clay particles, efcient stress transfer through the clay layers as well as increase in interfacial interaction between clay particles and epoxy. It has been reported that the increase of fracture surface area due to crack path deection is the major toughening mechanism in the epoxy clay nanocomposites.41 However, as the clay concentration increases the fracture toughness decreases steadily for ternary nanocomposites revealing that the
Polym Int 2010; 59: 986997

90 80 70 Tensile stress (MPa) 60 50 40 30 20 10 0 0 1 2 3 4 5 6 7 8 Tensile Strain (%)


Epoxy Epoxy/5 phr PES Epoxy/5 phr PES/1 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B Epoxy/5 phr PES/5 phr Cloisite 30B

Figure 6. Tensile stressstrain behaviour of the ternary nanocomposites.

thermoplastic is the main contributing factor to the increase in toughness of the nanocomposites rather than the clay alone. At higher clay concentration there is a possibility of isolated agglomeration of clay particles. These isolated clay agglomerates can act as failure sites from where the crack initiates and this reduces the stress intensity factor. Nanocomposites for which the clay particles are dispersed by planetary ball milling show a greater improvement in tensile, exural and fracture toughness properties compared to ultrasonicated systems. This may be due to the shear force exerted by ball milling being more effective than ultrasonication in dispersing the clay layers in the polymer matrix resulting in higher surface area of interaction between clay layers and matrix. Zunjarrao et al.11 reported that the fracture toughness of an epoxy/clay system was increased by 35 and 20% with 2% volume fraction of montmorillonite modied with octadecyl ammonium salt incorporated by shear mixing and ultrasonication, respectively. Tensile and exural stressstrain behaviour of the ternary nanocomposites is shown in Figs 6 and 7, respectively. It can be seen that the area under the stressstrain curve decreases with an increase in clay concentration for the PES-toughened epoxy

991

c 2010 Society of Chemical Industry

www.interscience.wiley.com/journal/pi

www.soci.org

A Asif et al.

140 120 Flexural Strength (MPa) 100

0.6

0.4 Tan 0.2 0.0


0 1 2 3 4 5 6 7 8 Flexural Strain (%) Epoxy Epoxy/5 phr PES Epoxy/5 phr PES/1 phr Cloisite 30B Epoxy/5 phr PES/3 phr Cloisite 30B Epoxy/5 phr PES/5 phr Cloisite 30B Epoxy/5 phr PES/8 phr Cloisite 30B

80 60 40 20 0

90

180

270

Temperature (C) Epoxy Epoxy/5phr PES Epoxy/5phr PES/1 phr Cloisite 30B Epoxy/5phr PES/3 phr Cloisite 30B Epoxy/5phr PES/5 phr Cloisite 30B Epoxy/5phr PES/3 phr Cloisite 30B (Planetary ball milling)
Figure 9. Tan versus temperature for PES-toughened epoxy clay ternary nanocomposites.

Figure 7. Flexural stressstrain behaviour of the ternary nanocomposites.

100 80 % weight 60 40 20 0 0 200 400 600 800 1000 Temperature (C)


Epoxy/5 phr PES Epoxy/5 phr PES/3 phr Cloisite 30B Epoxy/5 phr PES/5 phr Cloisite 30B Epoxy/5 phr PES/8 phr Cloisite 30B

Thermomechanical analysis (TMA) TMA was used to determine the CTE and Tg of the nanocomposites. CTE and Tg of the ternary nanocomposites are given in Table 2. The data reveal that CTE values decrease marginally with an increase in clay content up to 5 phr. This decrease may be due to the uniform and nanolevel distribution of clay particles in the epoxy matrix resulting in a restriction of mobility of the epoxy chains and also due to efcient stress transfer to clay layers with large surface area. The reduction in CTE with the incorporation of clay platelets was also reported by us30,31 and several other authors.15 No appreciable change in Tg is observed with the incorporation of clay into the epoxy matrix. However, DMTA and DSC analyses reveal a considerable decrease in Tg on incorporation of clay (Table 2). Dynamic mechanical thermal analysis The viscoelastic properties of the epoxy resin, PES-modied epoxy resin and PES-modied epoxy resin clay ternary nanocomposites were investigated using DMTA. The tan curves for the PESmodied epoxy clay ternary nanocomposites are shown in Fig. 9. The curves for the nanocomposite show only a single Tg value. Although two Tg values corresponding to epoxy-rich and thermoplastic-rich regions are expected, only one Tg value is observed because of the close proximity of Tg of the PES and Tg of the DDS-cured epoxy resin, and also due to homogeneity without phase separation. Tg of the nanocomposites decreases by about 7 C (from 216 C for epoxy to 209 C) on incorporation of 5 phr clay. During intercalation/exfoliation, the interfacial surface area of the clay particles and the interaction with polymer can drastically alter the chain kinetics in the regions surrounding them and lead to a lower crosslink density as suggested by Yasmin et al.44 Another possible reason for the decrease of Tg of the nanocomposites with increasing clay loading is due to the plasticizing effect of the
Polym Int 2010; 59: 986997

Figure 8. TGA curves of the nanocomposites.

system resulting in a gradual decrease in fracture toughness (Table 3). It is also observed that, as the clay content increases, the slope of the stressstrain curve also increases, indicating an increase in modulus (Table 3). Thermogravimetric analysis Thermograms of PES-toughened epoxy clay ternary nanocomposites are shown in Fig. 8. No appreciable change is observed in the initial degradation temperature and overall thermal stability except a slight increase in percentage char residue at higher temperature.

992

www.interscience.wiley.com/journal/pi

c 2010 Society of Chemical Industry

PES-toughened epoxy clay ternary nanocomposites

www.soci.org enhancement effect from the addition of rigid inorganic clay particles. Ternary nanocomposites processed using the planetary ball mill dispersion technique with 3 phr clay show an increase of 21% in storage modulus, whereas the ultrasonication dispersion technique leads to a 12% increase for the neat or PES-modied epoxy resin. This clearly indicates an enhanced interfacial interaction between clay layers and the matrix due to exfoliated morphology (Fig. 5). However, the increase in storage modulus is much less in the present system compared to PEEKMOH-toughened epoxy clay ternary nanocomposites reported earlier31 due to the absence of any chemical reaction/phase separation between PES and epoxy. Scanning electron microscopy The phase morphology of the PES-toughened epoxy and PEStoughened epoxy clay ternary nanocomposites were examined using SEM. SEM micrographs of tensile-failed surfaces of epoxy and epoxy/5 phr PES/3 phr Closite 30B and fracture toughnessfailed surfaces of epoxy, epoxy/5 phr PES, epoxy/3phr Cloisite 30B and epoxy/5 phr PES/3 phr Cloisite 30B are shown in Figs 11(a) and (b) and 12(a)(d), respectively. All the blends and nanocomposites are found to be homogeneous. This is due to the intermolecular hydrogen bonding between sulfonyl groups of PES and hydroxyl groups of epoxy along with the intramolecular hydrogen bonding between sulfonyl groups of DDS and hydroxyl groups of epoxy, as shown in Scheme 2, restricting phase separation. Similarly, Ni and Zheng also reported the effect of inter- and intramolecular hydrogen bonding between sulfonyl groups of diaryl sulfone and hydroxyl groups of epoxy on phase separation in epoxy/poly(caprolactone) blend systems.46 The SEM micrographs of tensile-failed specimens (Fig. 11) reveal that the fracture surface of the neat epoxy system shows typical characteristic of brittle fracture. The surface is smooth with uninterrupted crack propagation. The failure surface of epoxy/5 phr PES/3 phr Closite 30B is rough with ridge patterns, and river markings can also be seen on the fracture surface. The roughness of the fracture surface is for two reasons. First, it is an indication of crack path deection. Second, the roughness indicates the ductile nature of the crack. The matrix becomes less brittle in comparison with the unmodied epoxy resin because of a decrease in crosslink density. The polar sulfone groups of PES enhance the interfacial adhesion through the hydrogen bonding interaction between epoxy matrix and PES. Similarly a strong interfacial/polar interaction between functional hydroxyl groups of the organic moiety of clay/epoxy and PES of ternary nanocomposites inuences the crack path deection. Another factor responsible for the increase in the fracture toughness is the local plastic deformation of the matrix. According to Hedrick et al.,47 river markings on the fracture surface are an indication
Secondary hydroxyl group of epoxy/ hydroxyl group of clay/ hydroxyl group of clay modifier

3000 Storage modulus (MPa)

2000

1000

0 0 50 100 150 200 250

Temperature (C)
Epoxy Epoxy/5phr PES Epoxy/5phr PES/1 phr Cloisite 30B Epoxy/5phr PES/3 phr Cloisite 30B Epoxy/5phr PES/5 phr Cloisite 30B Epoxy/5phr PES/3 phr Cloisite 30B Planetary ball milling)

Figure 10. Storage modulus versus temperature for PES-toughened epoxy clay ternary nanocomposites.

organic modier of the clay layers as reported by Chen et al.36 They also reported a lower Tg with higher d spacing. In the present system, exfoliated nanocomposites show a lower Tg compared to intercalated nanocomposites (Fig. 9). The decrease in Tg may also be due to the formation of matrix interface between the silicate layers36 where properties are different from those of the bulk of the matrix. A lower Tg with higher clay content was also reported by us and by many other investigators.30,31,36,45 The storage modulus (E ) values, recorded as a function of temperature, for PES-toughened epoxy clay ternary nanocomposites are shown in Fig. 10. The storage moduli of the PES-modied epoxy resin and PES-modied epoxy resin clay nanocomposites are found to be higher than that of neat resin below Tg . With an increase in temperature, the storage modulus decreases, with a sharp decrease being observed near Tg . The storage moduli of the PES-modied epoxy and PES-modied epoxy clay ternary nanocomposites are slightly lower than that of the neat epoxy resin in the rubbery plateau region, indicating lower crosslink density of the modied systems. The storage modulus of the nanocomposites increases with an increase in clay concentration irrespective of the processing techniques because of an

O H O S O

Sulfone group of PES/ sulfone group of DDS

993

Scheme 2. Chemical interactions of hydroxyl groups with sulfone groups.

Polym Int 2010; 59: 986997

c 2010 Society of Chemical Industry

www.interscience.wiley.com/journal/pi

www.soci.org

A Asif et al.

(a)

(b)

Figure 11. SEM images of tensile-failed surfaces of (a) epoxy and (b) epoxy/5 phr PES/3 phr Cloisite 30B.

(a)

(b)

(c)

(d)

Figure 12. SEM images of fracture-failed surfaces of (a) epoxy, (b) epoxy/5 phr PES, (c) epoxy/3 phr Cloisite 30B and (d) epoxy/5 phr PES/3 phr Cloisite 30B.

of the plastic deformation of the matrix. On careful observation of Fig. 11(b), there are more crack path deection sites with thermoplastic inclusion. The fracture surface of the neat epoxy system (Fig. 12(a)) exhibits an uninterrupted and featureless crack propagation resulting from the brittle failure of epoxy. In contrast, the fracture surface of 5 phr PES-toughened epoxy (Fig. 12(b)) shows an improved toughness characterized by crack path deection, rough and ridge patterns with river marking lines. The fracture surface of the epoxy incorporating 3 phr clay (Fig. 12(c)) shows only a slight improvement in fracture toughness as evidenced from the agglomerated and debonded clay particles around the onset of crack deection. Liu et al.48 also observed similar fractographic appearance for nanoclay-lled epoxy systems. The micrograph of 3 phr clay and 5 phr PES-reinforced epoxy ternary nanocomposites (Fig. 12(d)) shows a better dispersion of clay particles as well as the presence of PES in the ternary nanocomposites. The fracture toughness of the PES-toughened epoxy clay system is improved

compared with epoxy clay ternary nanocomposites, but inferior to that of the PES-toughened epoxy blend. This may be because of the presence of debonded clay particles. However, the addition of clay particles to the PES-toughened epoxy blend reduces the toughness, but increases the modulus. SEM images of fracture-failed surface of epoxy/5 phr PES/3 phr clay ternary nanocomposite processed using ultrasonication and planetary ball mill clay dispersion techniques are shown in Figs 13(a) and (b), respectively. The fracture surface of the ternary nanocomposite processed using the planetary ball mill to disperse the clay shows multiple fracture with rougher surface resulting in high fracture toughness (Table 3) compared to samples with clay dispersed using ultrasonication (Fig. 13(a)). Optical transparency Photographs of neat epoxy, PES-toughened epoxy and PEStoughened epoxy clay ternary nanocomposites are shown
Polym Int 2010; 59: 986997

994

www.interscience.wiley.com/journal/pi

c 2010 Society of Chemical Industry

PES-toughened epoxy clay ternary nanocomposites

www.soci.org

(a)

(b)

Figure 13. SEM images of fracture-failed surface of epoxy/5 phr PES/3phr clay ternary nanocomposite processed using (a) ultrasonication and (b) planetary ball mill clay dispersion techniques.

(a)

(b)

(c)

(d)

Figure 14. Photographs of (a) epoxy, (b) epoxy/5 phr PES, (c) epoxy/5 phr PES/3 phr Cloisite 30B and (d) epoxy/5 phr PES/5 phr Cloisite 30B.

in Fig. 14. As the clay concentration increases, the optical transparency of the PES-toughened epoxy decreases due to the occasional presence of agglomerated clay particles in intercalated morphology of the ternary nanocomposites. However, optical transparency is retained even with 5 phr clay concentration unlike conventional microlled composites. The retention of optical transparency is due to the nanolevel distribution of clay platelets. Another plausible reason is that the dimensions of the clay platelets are less than the wavelength of the light, allowing the passage of light without hindrance. Oxygen gas permeability Oxygen gas permeability data for neat epoxy, epoxy clay and PES-toughened epoxy clay ternary nanocomposites are shown in Fig. 15. The oxygen gas permeability increases to 26% on addition
Polym Int 2010; 59: 986997

of 5 phr PES to the neat epoxy system, indicating the amorphous nature of the thermoplastic. Espuche et al.49 also observed that the permeability of thermoplastic/thermoset blends increases with an increase in thermoplastic content. But the permeability is reduced to 70% on incorporation of 3 phr clay to the neat epoxy system, indicating the layered nature of the clay particles, which act as a barrier to penetrants. When PES is added to the epoxy clay nanocomposites, the reduction in oxygen gas permeability is only 30% of that of the neat epoxy system. It is interesting to note that as the clay concentration increases from 3 to 5 phr in PES-toughened epoxy clay ternary nanocomposite, the reduction in permeability increases from 30 to 57% of that of the neat epoxy system. It is important to state that gas permeability is controlled by the clay morphology of the nanocomposite50 and the interaction between polymer matrix and organic moieties of clay layers. The reduction in gas permeability is due to tortuous

995

c 2010 Society of Chemical Industry

www.interscience.wiley.com/journal/pi

www.soci.org

A Asif et al.

140 Permeability mL m2 day1 120 100 80 60 40 20 0


a b c d e

markings increased with nanoclay incorporation conrming that the nanoclay not only takes part in improving the stiffness, but also in improving the toughness of the nanocomposites along with the thermoplastic. In addition, planetary ball-milled samples showed multi-plane fracture compared to ultrasonicated ones resulting in high fracture toughness. A marginal improvement in char residue was observed with an increase in clay content. The optical transparency of PEStoughened epoxy blends was retained even with 5 phr loading of clay particles.

ACKNOWLEDGEMENTS
The authors thank the Director, Vikram Sarabhai Space Centre (VSSC) for giving permission to publish the article. One of the authors (A.Asif) is thankful to Chairman, Indian Space Research Organisation and Director, VSSC for providing ISRO research fellowship. Thanks are due to all colleagues of the Analytical and Spectroscopy Division and Material Characterisation Division of VSSC, Sree Chitra Tirunal Institute for Medical Sciences and Technology, Biomedical Technology Wing, Hindustan Latex Ltd, Thiruvananthapuram for analytical support.

Composition
Figure 15. Oxygen gas permeability of the nanocomposites: (a) epoxy; (b) epoxy/5 phr PES; (c) epoxy/3 phr Cloisite 30B; (d) epoxy/5 phr PES/3 phr Cloisite 30B; (e) epoxy/5 phr PES/5 phr Cloisite 30B.

paths; in other words, a longer diffusion path through which gas penetrants travel in the presence of layered silicates, as shown in Fig. 15.

REFERENCES
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 Kornmann X, Lindberg H and Berglun LA, Polymer 42:4493 (2001). Utracki LA, Sepehr M and Boccaleri E, Polym Adv Technol 18:1 (2007). Okada A and Usuki A, Macromol Mater Sci Eng 291:1449 (2006). Paul DR and Robeson LM, Polymer 49:3187 (2008). Hussain F, Hojjati M, Okamoto M and Russell EG, J Compos Mater 40:1511 (2006). Utracki LA, Clay-Containing Polymeric Nanocomposites, vol. 2. Rapra Technology, Shawbury (2004). Pinnavaia TJ and Beall GW, Polymer Clay Nanocomposites. John Wiley, Chichester (2001). Ray SS and Okamoto M, Prog Polym Sci 28:1539 (2003). Alexandre M and Dubois , Mater Sci Eng 28:1 (2000). Ogasawara , Ishida Y, Ishikawa T, Aoki T and Ogura T, Composites A 37:2236 (2006). Zunjarrao SC, Sriraman R and Singh RP, J Mater Sci 41:2219 (2006). Hai-jun L, Guo-Zheng L, Xiao-yan M, Bao-yan Z and Xiang-bao C, Polym Int 53:1545 (2004). Rice BP, Chen C, Cloos L and Curliss D, SAMPE J 37:2 (2001). Immerman JF, Hayes BS and Seferis JC, Compos Sci Technol 62:1249 (2002). Campbell SG and Johnston C, NASA Glenns Research and Technology Reports. NASA (2004). Final report of the X-33 Liquid Hydrogen Tank Test Investigation Team, George C Marshall Space Flight Center, Huntsville, NASA report, May (2000). Lan T and Pinnavaia JJ, Chem Mater 6:2216 (1994). Becker O, Varley R and Simon G, Polymer 43:4365 (2002). Harsch M, Karger-Kocsis J and Holst M, Eur Polym J 43:1168 (2007). Xu WB, Zhou ZF, He PS and Pan WP, J Thermal Anal Calorim 78:113 (2004). Liu T, Tiju WC, Tong Y, He C, Goh SS and Chung TS, J Appl Polym Sci 94:1236 (2004). Francis B, Thomas S, Jose J, Ramaswamy R and Rao VL, Polymer 46:12372 (2005). Francis B, Thomas S, Asari GV, Ramaswamy R, Jose S and Rao VL, J Polym Sci B: Polym Phys 44:541 (2006). Mimura K, Ito H and Fujioka H, Polymer 41:4451 (2000). Pasquale GD, Motta O, Rocca A, Carter JT, McGrail PT and Acierno D, Polymer 38:4345 (1997). Frohlich J, Thomann R and Mulhaupt R, Macromolecules 36:7205 (2003). Balakrishnan S, Start PR, Raghavan D and Hudson SD, Polymer 46:11255 (2005). Isik I, Yilmazer U and Bayram G, Polymer 44:6371 (2003). Peng M, Li H, Wu L, Chen Y, Zheng Q and Gu W, Polymer 46:7612 (2005). Asif A, Leena K, Rao VL and Ninan KN, J Appl Polym Sci 106:2936 (2007).

CONCLUSIONS
PES-toughened epoxy clay ternary nanocomposites were processed by melt mixing of PES with epoxy along with clay by mechanical stirring followed by ultrasonication/planetary ball milling. Storage modulus was increased rapidly at gelation point and levelled off as vitrication occurs. Similarly loss modulus was also increased to a maximum at a particular time and thereafter decreased drastically in the rheological plots of the blends. The gel time of epoxy was increased on addition of PES and decreased on addition of clay. XRD revealed that the ultrasonication technique gave intercalated morphology while the planetary ball milling technique gave exfoliated morphology. Tensile, exural and storage moduli and fracture toughness were increased on addition of clay platelets. However, there is a marginal decrease in tensile and exural strength as the clay content increases. The increase in fracture toughness and storage modulus, and improvement in tensile and exural properties were found to be much greater for planetary ball-milled samples compared to ultrasonicated samples. The CTE values decreased with an increase in clay content up to 5 phr; thereafter they increased. Oxygen gas permeability was decreased considerably on incorporation of clay into the epoxy matrix. But with the incorporation of PES the permeability of epoxy clay nanocomposites was increased. However, with an increase in clay concentration, the permeability of PES-toughened epoxy clay ternary nanocomposites was further decreased showing a synergic effect. SEM revealed no phase separation due to strong polar interactions and hydrogen bonding between PES and epoxy matrix. However, the area of crack path deection and river

996

www.interscience.wiley.com/journal/pi

c 2010 Society of Chemical Industry

Polym Int 2010; 59: 986997

PES-toughened epoxy clay ternary nanocomposites


31 Asif A, Rao VL, Saseendran V and Ninan KN, Polym Eng Sci 49:756 (2009). 32 Standard test methods for exural properties of unreinforced and reinforced plastics and electrical insulating materials, ASTM D790. 33 Standard test methods for plane-strain fracture toughness and strain energy release rate of plastic materials, ASTM D 5045. 34 Ganguli S, Dean D, Jordan K, Price G and Vaia R, Polymer 44:6901 (2003). 35 Dean D, Walker R, Theodore M, Hampton E and Nyairo E, Polymer 46:3014 (2005). 36 Chen JS, Ober CK, Zhang Y, Ulrich W and Giannelis E, Polymer 43:4895 (2002). 37 Lan T, Kaviratna PD and Pinnavaia TJ, J Phys Chem Solids 57:6 (1996). 38 Wang J, Lan T and Pinnavaia JT, Chem Mater 8:2200 (1996). 39 Tolle TB and Anderson D, Compos Sci Technol 62:1033 (2002). 40 Zilg C, Mulhaupt R and Finter , J Macromol Chem Phys 200:661 (1999).

www.soci.org
41 Wang K, Chen L, Wu J, Toh ML, He C and Yee AF, Macromolecules 38:788 (2005). 42 Qi B, Zhang QX, Bannister M and Mai YW, Compos Struct 75:514 (2006). 43 Basara C, Yilmazer U and Bayram G, J Appl Polym Sci 98:1081 (2005). 44 Yasmin A, Luo JJ, Abot JL and Daniel IM, Compos Sci Technol 66:2415 (2006). 45 Lee L and Lichtenhan JD, J Appl Polym Sci 73:1993 (1996). 46 Ni Y and Zheng S, Polymer 46:5828 (2005). 47 Hedrick JL, Jurek MJ, Yilgor I and McGrath JE, Polym Prepr 26:293 (1985). 48 Liu T, Tjiu WC, Tong Y, He C, Goh SS and Chung TS, J Appl Polym Sci 94:1236 (2004). 49 Espuche E, Escoubes M, Pascault JP and Taha M, J Polym Sci B: Polym Phys 37:473 (1999). 50 Bharadwaj RK, Macromolecules 34:9189 (2001).

997
www.interscience.wiley.com/journal/pi

Polym Int 2010; 59: 986997

c 2010 Society of Chemical Industry

Anda mungkin juga menyukai