A comparison of the mechanical strength and stiffness of MWNT-PMMA and MWNT-epoxy nanocomposites

Similar documents
The stress transfer efficiency of a single-walled carbon nanotube in epoxy matrix

RHEOLOGICAL AND MORPHOLOGICAL PROPERTIES OF NANOCOMPOSITES BASED ON PA66/PA6/MULTI-WALLED CARBON NANOTUBE PREPARED BY MELT MIXING

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

The effect of interfacial bonding on the damping behavior of nanocomposites

Improvement of the chemical, thermal, mechanical and morphological properties of polyethylene terephthalate graphene particle composites

Thermal degradation kinetics of Arylamine-based Polybenzoxazines

Defense Technical Information Center Compilation Part Notice

Fibrillated Cellulose and Block Copolymers as a Modifiers of Unsaturated Polyester Nanocomposites

Assessment of carbon nanotube dispersion and mechanical property of epoxy nanocomposites by curing reaction heat measurement

Three-dimensional flexible and conductive interconnected graphene networks grown by chemical vapour deposition

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

Thermal Stability of Trifunctional Epoxy Resins Modified with Nanosized Calcium Carbonate

POLYURETHANE SURFACE TREATMENT ON TWO KINDS OF BASALT FIBER COMPOSITE AND MECHANICAL PROPERTIES COMPARISON

GRAPHENE BASED POLY(VINYL ALCOHOL) NANOCOMPOSITES: EFFECT OF HUMIDITY CONTENT

EFFECT OF SONICATION AND HIGH SHEAR MIXING PARAMETERS ON NANOCLAY DISPERSION IN EPOXY

Supplementary Figures

The experimental determination of the onset of electrical and thermal conductivity percolation thresholds in carbon nanotube-polymer composites

Supporting Information for. Biobased, Self-Healable, High Strength Rubber with Tunicate. Cellulose Nanocrystals

NITRILE RUBBER (NBR) NANOCOMPOSITES BASED ON DIFFERENT FILLER GEOMETRIES (Nanocalcium carbonate, Carbon nanotube and Nanoclay)

Thermal Properties of Carbon Nanotube (CNT) Reinforced Polyvinyl Alcohol (PVA) Composites

Halloysite. Nanotubes Epoxy. NanoComposites

Mechanical properties of multi-walled carbon nanotube/polyester nanocomposites

IMPROVEMENT IN MECHANICAL PROPERTIES OF MODIFIED GRAPHENE/EPOXY NANOCOMPOSITES

The morphology of PVDF/1Gra and PVDF/1IL/1Gra was investigated by field emission scanning

RHEOLOGICAL CHARACTERIZATION OF THERMOSETTING RES- IN SYSTEM WITH THERMOPLASTIC FUNCTIONAL LAYER

Nano-materials in Polymer Composites for High-Volume Applications

Research Article Effects of CNT Diameter on the Uniaxial Stress-Strain Behavior of CNT/Epoxy Composites

[83] RMUTP Research Journal: Special Issue 2014 The 4 th RMUTP International conference: Textiles and Fashion

DETERMINATION OF GLASS TRANSITION TEMPERATURE FOR POLYESTER / GRAPHENE OXIDE AND POLYESTER / GRAPHITE COMPOSITE BY TMA AND DSC

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

CHEM-C2410: Materials Science from Microstructures to Properties Composites: basic principles

ENHANCED THERMAL CONDUCTIVITY OF EPOXY BASED COMPOSITES WITH SELF-ASSEMBLED GRAPHENE-PA HYBRIDS

Materials Science and Engineering A

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

MICROMECHANICAL DEFORMATIONS IN PARTICULATE FILLED POLYMERS: THE EFFECT OF ADHESION

Influence of graphene nanoplatelets on curing and mechanical properties of graphene/epoxy nanocomposites

EFFECTS OF MOLECULAR STRUCTURE ON MACROSCOPIC MECHANICAL PROPERTIES OF AN ADVANCED POLYMER (LARC -SI)

A Hydrophilic/Hydrophobic Janus Inverse-Opal

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

POLYBENZOXAZINE BASED NANOCOMPOSITES REINFORCED WITH MODIFIED GRAPHENE OXIDE

Change in physico-mechanical and thermal properties of polyamide / silica nanocomposite film

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

Department of Mechanical Engineering, Imperial College London, London SW7 2AZ, UK

CHARACTERIZATION OF POLY(METHYL METHACRYLATE BASED NANOCOMPOSITES ENHANCED WITH CARBON NANOTUBES

Teorie netkaných textilií

Fracture of vacancy-defected carbon nanotubes and their embedded nanocomposites

Supporting Information

Supporting Information

Preparation, Characterization and comparative Temperature dependent Electrical Properties of Polythiophene & its Nanocomposites using Carbon Nanotubes

Composite Materials Research Laboratory, State University of New York at Buffalo, Buffalo, NY , USA

Supporting Information

Characteristics of Thermosonic Anisotropic Conductive Adhesives (ACFs) Flip-Chip Bonding

Effect of Cross-Link Density on Interphase Creation in Polymer Nanocomposites

Interface Toughness of Carbon Nanotube Reinforced Epoxy Composites

Electronic Supplementary Information. Three-Dimensional Multilayer Graphene Web for Polymer Nanocomposites with

Mechanical Behaviors Study of Epoxy Resins Reinforced with Nano-SiO 2 Particles Jian Zheng, Xiong Chen, Deng Jia, Xin Tong

Introduction. Theoretical Model and Calculation of Thermal Conductivity

Re-agglomeration of Carbon nanotubes in two-part epoxy system; Influence of the concentration

Multi-Wall Carbon Nanotubes/Styrene Butadiene Rubber (SBR) Nanocomposite

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

Molecular Dynamics Study of the Effect of Chemical Functionalization on the Elastic Properties of Graphene Sheets

Functionalized flexible MOF as filler in mixed matrix membranes for highly selective separation of CO 2 from CH 4 at elevated pressures

MODELLING INTERACTION EFFECT OF NANOSILICA PARTICLES ON NANOSILICA/EPOXY COMPOSITE STIFFNESS

Circuit elements in carbon nanotube-polymer composites

Tuskegee-Cornell PREM A Research and Educational Partnership in Nanomaterials between Tuskegee University and Cornell University

CURE DEPENDENT CREEP COMPLIANCE OF AN EPOXY RESIN

SWCNT COMPOSITES, INTERFACIAL STRENGTH AND MECHANICAL PROPERTIES

Novel Dispersion and Self-Assembly

Carbon nanotube coated snowman-like particles and their electro-responsive characteristics. Ke Zhang, Ying Dan Liu and Hyoung Jin Choi

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

Effect of Carbon Nanotubes Combined with Hexaphenoxycyclotriphosphazene on the Flame Retardancy of Epoxy Resin

Mechanical and Thermoviscoelastic Behavior of Clay/Epoxy Nanocomposites

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

Improvement of Carbon Nanotubes Dispersivity in Poly(Styrene/Methacrylate) Composites by Chemical Functionalization

Improvement of Thermal and Mechanical Properties of Carbon Nanotube Composites through Chemical Functionalization

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

Effect of Functionalization on the Crystallization Behavior of MWNT-PBT Nanocomposites

APPLICATION OF THERMAL METHODS IN THE CHEMISTRY OF CEMENT: KINETIC ANALYSIS OF PORTLANDITE FROM NON- ISOTHERMAL THERMOGRAVIMETRC DATA

EPOXY/SINGLE WALLED CARBON NANOTUBE NANOCOMPOSITE THIN FILMS FOR COMPOSITES REINFORCEMENT

Interfacial Load Transfer in Polymer/Carbon Nanotube Nanocomposites with a Nanohybrid Shish Kebab Modification

Effect of different crosslink densities on the thermomechanical properties of polymer nanocomposites

Aminopropyltrimethoxysilane-Functionalized Boron Nitride. Nanotube Based Epoxy Nanocomposites with Simultaneous High

New Developments in Structure/Property Relationships: Relating Fluid Ingress to mechanical and thermal properties

Charpy Impact Response of Glass Fiber Reinforced Composite with Nano Graphene Enhanced Epoxy

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

Direction sensitive deformation measurement with epoxy/cnt nanocomposites

Surface Modification of Carbon Fibres for Interface Improvement in Textile Composites

MOLECULAR MODELING OF THERMOSETTING POLYMERS: EFFECTS OF DEGREE OF CURING AND CHAIN LENGTH ON THERMO-MECHANICAL PROPERTIES

A NEW GENERATION OF CONSTRUCTION MATERIALS: CARBON NANOTUBES INCORPORATED TO CONCRETE AND POLYMERIC MATRIX

Non-conventional Glass fiber NCF composites with thermoset and thermoplastic matrices. F Talence, France Le Cheylard, France

Supporting information for

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

CARBON NANOTUBE-POLYMER COMPOSITES: AN OVERVIEW Brian Grady University of Oklahoma

A Study of Post Heat Treatment Temperature Variables on Mechanical Properties of Nanopolymer Composites

The Pennsylvania State University. The Graduate School ANALYSIS OF DAMPING CHARACTERISTICS OF POLYMERIC COMPOSITES CONTAINING CARBON NANOTUBES

International Journal of Pure and Applied Sciences and Technology

ECHNICAL-SCIENTIFICPAPERS-

DEVELOP WEAR-RESISTANT POLYMERIC COMPOSITES BY USING NANOPARTICLES

IMPROVEMENT IN THE MECHANICAL PROPERTIES OF B-STAGED CARBON NANOTUBE/EPOXY BASED THIN FILM SYSTEMS

The effect of surface functional groups of nanosilica on the properties of polyamide 6/SiO 2 nanocomposite

Transcription:

Composite Interfaces, Vol. 14, No. 4, pp. 285 297 (2007) VSP 2007. Also available online - www.brill.nl/ci A comparison of the mechanical strength and stiffness of MWNT-PMMA and MWNT-epoxy nanocomposites LU-QI LIU and H. DANIEL WAGNER Department of Materials and Interfaces, Weizmann Institute of Science, Rehovot 76100, Israel Received 5 May 2006; accepted 18 August 2006 Abstract The surface of multi-wall carbon nanotubes (MWNTs) was functionalized by covalent linking of long alkyl chains. Such functionalization led to a much better tube dispersion in organic solvents than pristine nanotubes, favored the formation of homogenous nanocomposite films, and yielded good interfacial bonding between the nanotubes and two polymer matrices: a thermoset (Epon 828/T-403) and a thermoplastic (PMMA). Tensile tests indicated, however, that the reinforcement was greatly affected by the type of polymer matrix used. Relative to pure PMMA, a 32% improvement in tensile modulus and a 28% increase in tensile strength were observed in PMMA-based nanocomposites using 1.0 wt% nanotube filler. Contrasting with this, no improvement in mechanical properties was observed in epoxy-based nanocomposites. The poorer mechanical performance of the latter system can be explained by a decrease of the crosslinking density of the epoxy matrix in the nanocomposites, relative to pure epoxy. Indeed we demonstrate that the presence of nanotubes promotes an increase in the activation energy of the curing reaction in epoxy, and a decrease of the degree of curing. Keywords: Multi-wall carbon nanotube; functionalized carbon nanotube; nanocomposites; mechanical properties; activation energy. 1. INTRODUCTION An increasing number of studies reveal that to significantly improve the mechanical properties of polymer-based nanocomposites, carbon nanotube (CNT) dispersion and nanotube-polymer interfacial interaction must be adequately controlled and understood [1 10]. However, the effect of these parameters appears to vary depending on whether the polymer matrix is a thermosetting or a thermoplastic material. Generally speaking, improvements of the mechanical properties seem to be more easily obtained with thermoplastic polymer matrices than with thermosetting matrices. For example, significant nanocomposite stiffness increases arise with nylon-6 To whom correspondence should be addressed. E-mail: Daniel.wagner@weizmann.ac.il

286 L.-Q. Liu and H. D. Wagner as a matrix [11, 12] whereas only marginal improvement or even a decrease in nanocomposite tensile modulus is observed with standard epoxy matrices [13 16]. The quasi-absence of reinforcing ability of nanotubes in epoxy is attributed to the poor dispersion and interfacial adhesion of nanotubes in the matrix. An effective solution to this double problem is to suitably functionalize the nanotube surface [7, 10, 17, 18]. In this paper, the surface of multi-wall carbon nanotubes was functionalized by covalent linking of long alkyl chains. Such functionalization led to a much better tube dispersion in organic solvents than the pristine nanotubes, and favored the formation of homogenous composite films [19, 20]. Moreover, the presence of chemical groups on the nanotubes surface was hypothesized to yield improved interfacial bonding between the nanotubes and two polymer matrices: a thermoset (Epon 828/T-403) and a thermoplastic (PMMA). As will be seen, unambiguous differences in reinforcing ability were observed following the introduction of nanotube in the two polymer matrix types. The poorer mechanical properties of the epoxy-based nanocomposite were linked to the increased activation energy of the curing reaction of epoxy, leading to a relatively lower crosslinking density. 2. EXPERIMENTAL 2.1. Materials The nanotubes used in the present work were multiwall carbon nanotubes (MWNT) from Sun Nanotech (Nanchang, China). They were functionalized (f -MWNT) according to a previously published procedure [19]. The nanotube diameter ranged between 30 and 70 nm. Previous elementary analysis (EA) and thermogravimetric analysis (TGA) measurements have indicated that the nanotube content in functionalized carbon nanotube samples is around 70 wt%. The epoxide resin used in this work was diglycidyl ether of bisphenol A (Epon 828) provided by Kidron Co. (Israel). Glycolitic polypropyleneoxide triamine (Jeffamine T-403) with a molecular weight of 440 was used as curing agent in the preparation of epoxy matrices. PMMA was purchased from Aldrich (M w = 350,000). 2.2. Epoxy-based nanocomposites The pure epoxy system (Epon 828/T-403, weight ratio: 100/42) was prepared by stirring the liquid mixture for 15 min at room temperature to make it homogenous, followed by degassing in vacuum for 2 h. The mixture was then poured into a dog-bone shaped RTV mold, cured at 80 C for 2 h and postcured at 125 C for 3 h. Since nanocomposites based on this epoxy system must be prepared by dissolving the nanotubes in CH 2 Cl 2, we prepared pure epoxy control specimens by mixing the curing agent T-403 with this solvent, then evaporating it completely before mixing with Epon 828. We ensured that solvent evaporation was indeed complete

Strength of MWNT nanocomposites 287 by measuring the glass transition temperature (T g ) and mechanical properties. These showed no significant change, an indication that CH 2 Cl 2 had indeed been completely removed. To prepare 1.0 wt% nanotube-epoxy resin composites, a given amount of f -MWNT samples was first dispersed in CH 2 Cl 2 ; the T-403 curing agent was then added to the solution which was mildly sonicated for 1 h using a 50 60 Hz bath sonicator. CH 2 Cl 2 was slowly evaporated by continuous stirring, and the mixture was put under vacuum at 40 C for several hours to remove all the residual solvent. A stoichiometric amount of Epon 828 was then added into the mixture, followed by the same steps as previously described to prepare the unreinforced Epon 828/T-403 matrix. The overall weight ratio of f -MWNT/Epon828/ T-403 was 1/25/75, which resulted in 1.0 wt% nanotube-epoxy composites. A similar procedure was used to prepare 0.1 wt% and 0.5 wt% nanotube based Epon 828/T-403 nanocomposites. 2.3. PMMA-based nanocomposites To prepare nanotube-pmma composites, a given amount of f -MWNT was first dispersed in CH 2 Cl 2, then mixed with a volume of PMMA solution in CH 2 Cl 2 and mildly sonicated for 1 h using a 50 60 Hz bath sonicator. After slowly removing most of the solvent at high temperature using magnetic stirring, the concentrated composites solution was poured into a dish, where the solvent evaporated to leave a film. The film was heated at 40 C under vacuum to let its weight reach an equilibrium value. Pure PMMA specimens were also prepared using this procedure. 2.4. Morphological characterization The morphology of the nanotubes/epoxy composites was investigated using a high resolution scanning electron microscope (SEM, Philips XL-30, The Netherlands) at an accelerating voltage of 5 10 kv. Prior to observation, the fractured surfaces (following the tensile tests, see below) were sputtered-coated with chromium. 2.5. Quasi-static tensile tests Tensile tests of the polymer and nanocomposites were performed with a Rheometric Scientific Minimat tensile tester using a 200 N load cell at a testing speed of 1.0 mm/min. The typical size of the epoxy-based nanocomposites was 1 mm (thickness) 2 mm (width) 20 mm (gauge length). The dimensions of the PMMA-based nanocomposite films were 130 180 µm (thickness) 2mm (width) 20 mm (gauge length). To ensure data accuracy and repeatability, seven samples of each type were tested. 2.6. Thermal analysis The curing behavior of Epon 828 with T-403 was monitored by differential scanning calorimetry (DSC) in the temperature range of 30 290 C using a Mettler differential

288 L.-Q. Liu and H. D. Wagner scanning calorimeter (DSC 30 connected to a TC11 processor). A sample weight of 11 ± 2 mg was used. To determine the activation energy of the curing reaction of neat epoxy resin and nanotube based composites, dynamic DSC scans were recorded at four different heating rates (i.e. 2.5, 5, 7.5, 10 C/min). Moreover, DSC tests in isothermal mode were conducted at 100 C. Non-isothermal scans were then performed on similar samples to obtain the heat of reaction necessary to complete the cure of the reactive system. 3. RESULTS AND DISCUSSION 3.1. Mechanical properties Typical tensile stress strain curves for epoxy-based and PMMA-based nanocomposites are presented in Fig. 1. The Young s modulus and ultimate tensile strength of both types of nanocomposites are presented in Fig. 2, which demonstrate that (i) the mechanical properties of epoxy-based nanocomposites are not significantly affected, positively or negatively, by the presence of f -MWNT, except possibly for a slight decrease in the failure strain; (ii) the mechanical properties of PMMAbased nanocomposites are significantly and increasingly improved by the presence of f -MWNT. The behavior of epoxy-based nanocomposites containing functionalized nanotubes is in fact not different from previous work with epoxy-based nanocomposites with pristine nanotubes [13, 21]. The immediate conclusion is that the better dispersion of f -MWNT (in various organic solvents) has no effect on the mechanical properties of the present epoxy-based nanocomposites, and the presence of alkyl chains attached on the nanotube surface, which in principle induces improved interfacial bonding, is ineffective in these nanocomposites. This drastically contrasts with the results for PMMA-based nanocomposites, as seen, for which both the tensile modulus and the tensile strength significantly increase upon adding small amounts of f-mwnt, thus confirming earlier data [7, 10 12]. Cooper et al. [22] reported a 25% increase in Young s modulus in PMMA reinforced with 4 wt% (aligned) carbon nanofibril. The f -MWNT based composite studied here exhibits better mechanical performance, which may be attributed to the quality of the nanotube dispersion in the matrix and the better tube-matrix interfacial adhesion. What is the intrinsic cause of such differences in the mechanical properties of epoxy-based and PMMA-based nanocomposites? Why would a given batch of nanotubes, with good dispersion and adhesion capability, give rise to much improved properties in one type of polymer and have no effect on another? Most published works to date indeed attribute this often observed weaker reinforcement effect in thermosetting polymers (such as epoxy) to poor nanotube dispersion and weak interfacial adhesion, leading to tube aggregation and easier tube pull out under an external force [4, 23]. We postulate here that another possibly determining cause of the lack of reinforcing effect in thermosets has its roots in a modification of the macromolecular network structure of epoxy resulting from the addition of

Strength of MWNT nanocomposites 289 Figure 1. Typical stress strain curves for (a) pure epoxy and epoxy-based nanocomposites; (b) PMMA and PMMA-based nanocomposites. MWNT into the matrix. Recent work with SWNT and nanofibers [24 26] reveals that the presence of nanotubes may alter the curing process, and a link between this effect and the mechanical behavior of SWNT-based epoxy composites has been seen [26]. 3.2. Differential scanning calorimetry We investigate possible changes in epoxy structure due to the introduction of f - MWNT by means of differential scanning calorimetry (DSC) in the dynamic and isothermal modes, and study the curing reaction of epoxy matrix with and without nanotubes. In the following, the discussion is based on the assumption that the heat evolved during the curing reaction is proportional to the degree of reaction.

290 L.-Q. Liu and H. D. Wagner Figure 2. Comparison of mechanical properties of epoxy-based and PMMA-based nanocomposites: (a) Tensile modulus (b) tensile strength. Dynamic DSC scans of pure Epon 828/T-403 at four different heating rates are presented in Fig. 3. The overall heat evolved in the reaction is taken as the average heat of reaction calculated from each thermogram shown in the figure. As seen, the peak exothermal temperature (T p ) shifts to higher temperatures as the heating rate is increased. Similar thermograms were obtained for all epoxy-based nanocomposites. From the data presented in Table 1, the heat of polymerization ( H 0 ) does not vary much with the heating rate and the coefficients of variation of the average heats of polymerization are below 9% in all cases. This justifies the validity of this method to measure the heat of polymerization [27]. As seen, the total heat of reaction H 0 strongly decreases with an increase in the nanotube content, an indication that the reaction mechanism has possibly been

Strength of MWNT nanocomposites 291 Figure 3. Dynamic DSC scans of Epon 828/T-403 at four heating rates. Table 1. Heat of polymerization for pure epoxy and epoxy-based nanocomposites, under different heating rates q ( Cmin 1 ) Heat of polymerization H 0 (J g 1 ) Pure epoxy 0.1 wt% f -MWNT 0.5 wt% f -MWNT 1.0 wt% f -MWNT nanocomposites nanocomposites nanocomposites 2.5 316 302 265 250 5 296 280 260 261 7.5 294 264 241 227 10 286 281 235 217 Average value 298 ± 12.8 281.8 ± 15.6 250.3 ± 14.5 238.8 ± 20.3 affected by the presence of the nanotubes. A similar decrease was also reported for poly(etherimide) (PEI) mixed with epoxy [27], for carbon fiber in epoxy resin [28], as well as for in-situ polymerization PMMA in the presence of carbon nanotube [29]. In the latter study, the average molecular weight and the glass transition temperature of the nanocomposite were found to be lower in comparison with pure PMMA prepared under the same condition. As a further step, dynamic mode DSC analysis was used to measure the activation energy of the reaction. Two kinetic models the Kissinger and the Flynn Wall Ozawa methods [30 33] were used to analyze the relation between the activation energy (E a ), the heating rate (q), and the peak exotherm (T p ). According to the Kissinger method, the activation energy is obtained from the maximum reaction rate, where d(dα/dt)/dt is zero at a constant-heating rate. The parameter α is the degree of cure, and is proportional to the heat generated during reaction. The key

292 L.-Q. Liu and H. D. Wagner expression resulting from the Kissinger method is as follows: d[ln(q/tp 2)] = E a d(1/t p ) R, (1) where R is the ideal gas constant. A plot of ln(q/tp 2) versus 1/T p provides the activation energy E a. The Flynn Wall Ozawa model provides the following equation: [ ] AEa log(q) = log 2.315 0.457E a, (2) g(a)r RT p where g(a) is a function of the degree of cure α. Applying these two models results in the linear plots presented in Fig. 4. The slopes of the lines provide the activation energies presented in Table 2. Compared to pure epoxy, the activation energy increases significantly with the weight content of nanotube using both models (the Flynn Wall Ozawa method provides slightly higher values than the Kissinger method). It may be concluded that the reaction between the epoxy and the curing agent is probably hindered by the presence of the nanotubes. This conclusion is further supported by the isothermal DSC analysis presented below. The degree of cure, α, up to time t in the isothermal reaction process can be obtained from isothermal DSC kinetics [25 27, 34]: α = H t /(H + H S ),where H t denotes the evolved heat up to time t,h is the total heat under the isothermal curve, and H s is the residual heat. As shown in Fig. 5, the degree of cure decreases noticeably as a result of the introduction of nanotubes into epoxy. This decrease in the degree of cure is consistent with the data of Table 2: indeed, the reported increase in activation energy does not favor crosslinking of the epoxy network. Furthermore, the glass transition temperature was measured to add to the above argument: the value of T g for pure epoxy is 81 C, and the addition of 0.1 wt% f -MWNT decreases the value down to 74 C. The likely conclusion, as above, is that the crosslink density of epoxy is reduced compared to that of pure epoxy, due to the presence of nanotubes. However, at higher nanotube content, the nanotubes increasingly prevent the free movement of polymer chains [35], an effect that is opposite to the lower crosslinking effect: this is reflected in the slight increase of T g to 77 C for 0.5 wt% nanotube content and to 78 C at 1 wt% nanotube content, respectively. Since a direct correlation exists between the Young s modulus of epoxy and the degree of crosslinking [36], the above result is reflected in the behavior of the modulus (and strength) observed in Fig. 2. The fracture surfaces of the nanocomposites were investigated by SEM. Fig. 6(a) provides evidence for good nanotube dispersion within matrices and strong interfacial adhesion between the nanotubes and epoxy, and the same is observed in the PMMA-based nanocomposites, see Fig. 6(b). In view of this, we therefore conclude that the lower mechanical properties observed in the nanotube-epoxy system has its origin in the lower degree of molecular crosslinking.

Strength of MWNT nanocomposites 293 Figure 4. Activation energies obtained by the Kissinger and Flynn Wall Ozawa methods: (a) pure epoxy system; (b) 0.1 wt%; (c) 0.5 wt%; (d) 1.0 wt% f -MWNT epoxy-based nanocomposites. 4. CONCLUSIONS Small amounts of functionalized multi-wall carbon nanotubes were well dispersed in an epoxy resin and a PMMA matrix, respectively. Contrasting properties were observed in these systems: (i) no significant improvement in modulus or strength was obtained for epoxy-based nanocomposites; (ii) a systematic improvement of modulus and strength was observed in the PMMA-based nanocomposites, as a function of the amount of nanotubes, up to 32% increase in tensile modulus and 28%

294 L.-Q. Liu and H. D. Wagner Figure 4. (Continued.) Table 2. Activation energies for epoxy and epoxy-based nanocomposites Kissinger (kj/mol) Epoxy 44.9 ± 2.0 48.8 ± 1.2 0.1 wt% f -MWNT nanocomposites 48.7 ± 0.6 52.6 ± 0.5 0.5 wt% f -MWNT nanocomposites 52.4 ± 1.2 56 ± 1.3 1.0 wt% f -MWNT nanocomposites 57.2 ± 2.7 60.6 ± 2.5 Flynn Wall Ozawa (kj/mol)

Strength of MWNT nanocomposites 295 Figure 5. Degree of cure of the reaction, α, against time at 100 C isotherm for pure epoxy and 1.0 wt% f -MWNT epoxy-based nanocomposites. Figure 6. SEM fracture surfaces of (a) 1.0 wt% f -MWNT epoxy-based nanocomposites, and (b) 1.0 wt% f -MWNT PMMA-based composites. increase in strength, with 1.0 wt% nanotube content. The presence of nanotubes promotes an increase in the activation energy of the curing reaction in epoxy, and a decrease of the degree of curing. As a result, the crosslinking density of the

296 L.-Q. Liu and H. D. Wagner Figure 6. (Continued.) epoxy matrix in the nanocomposites is lower than in pure epoxy, which we postulate is the main reason for the poorer mechanical performance of the epoxy-based nanocomposites. Future work should focus on improving the mechanical properties of epoxy-based nanocomposites by optimizing the stoichiometry and kinetics of matrix curing. Acknowledgements This project was supported by the NOESIS European project on Aerospace Nanotube Hybrid Composite Structures with Sensing and Actuating Capabilities, the G. M. J. Schmidt Minerva Centre of Supramolecular Architectures, and the Israeli Academy of Science. H. D. Wagner is the recipient of the Livio Norzi Professorial Chair and wishes to acknowledge the inspiring aid of L. Hampton and B. Goodman. REFERENCES 1. P. J. F. Harris, Carbon Nanotubes and Related Structures: New Materials for the Twenty-first Century, pp. 186 206. Cambridge University Press, UK (1999). 2. R. Saito, G. Dresselhaus and M. S. Dresselhaus, Physical Properties of Carbon Nanotubes, pp. 207 225. Imperial College Press, London, UK (1998). 3. T. W. Ebbesen (Ed.), Carbon Nanotubes: Preparation and Properties, pp. 225 246, CRC Press, Boca Raton, Florida (1997). 4. P. M. Ajayan, L. S. Schadler, C. Giannaris and A. Rubio, Adv. Mater. 12, 750 753 (2000). 5. D. Qian, E. C. Dickey, R. Andrews and T. Rantell, Appl. Phys. Lett. 76, 2868 2870 (2000).

Strength of MWNT nanocomposites 297 6. M. Cadek, J. N. Coleman, V. Barron, K. Hedicke and W. J. Blau, Appl. Phys. Lett. 81, 5123 5125 (2002). 7. H. Z. Geng, R. Rosen, B. Zheng, H. Shimoda, L. Fleming, J. Liu and O. Zhou, Adv. Mater. 14, 1387 1390 (2002). 8. M. D. Frogley, D. Ravich and H. D. Wagner, Compos. Sci. Technol. 63, 1647 1654 (2003). 9. A. H. Barber, S. Cohen and H. D. Wagner, Appl. Phys. Lett. 82, 4140 4142 (2003). 10. L. Q. Liu, A. H. Barber, S. Nuriel and H. D. Wagner, Adv. Function. Mater. 15, 975 980 (2005). 11. W. D. Zhang, L. Shen, I. Y. Phang and T. Liu, Macromol. 37, 256 259 (2004). 12. T. Liu, I. Y. Phang, L. Shen, S. Y. Chow and W. D. Zhang, Macromol. 37, 7214 7222 (2004). 13. K. T. Lau and D. Hui, Carbon 40, 1605 1606 (2002). 14. D. Penumadu, A. Dutta, G. M. Pharr and B. Files, J. Mater. Res. 18, 1849 1853 (2003). 15. L. Q. Liu and H. D. Wagner, Compos. Sci. Technol. 65, 1861 1868 (2005). 16. F. H. Gojny, M. H. G. Wichmann, U. Köpke, B. Fiedler and K. Schulte, Compos. Sci. Technol. 64, 2363 2371 (2004). 17. M. C. Paiva, B. Zhou, K. A. S. Fernando, Y. Lin, J. M. Kennedy and Y. P. Sun, Carbon 42, 2849 2854 (2004). 18. J.W.Yang, J.H.Hu, C.C.Wang, Y.J.Qin and Z.X.Guo,Macromol. Mater. Eng. 289, 828 832 (2004). 19. L. Q. Liu, S. Zhang, T. Hu, Z. X. Guo and C. Ye, Chem. Phys. Lett. 359, 191 195 (2002). 20. Y. J. Qin, L. Q. Liu, J. H. Shi and Z. X. Guo, Chem. Mater. 15, 3256 3260 (2003). 21. J.Zhu, J.D.Kim, H.Q.Peng, J.L.Margraveand E.V.Barrera, Nano Lett. 3, 1107 1113 (2003). 22. C. A. Cooper, D. Ravich, D. Lips, J. Mayer and H. D. Wagner, Compos. Sci. Technol. 62, 1105 1112 (2002). 23. L. S. Schadler, S. C. Giannaris and P. M. Ajayan, Appl. Phys. Lett. 73, 3842 3844 (1998). 24. D. Puglia, L. Valentini and J. M. Kenny, J. Appl. Poly. Sci. 88, 452 458 (2003). 25. J. Bae, J. Jang and S. Yoon, Macromol. Chem. Phys. 203, 2196 2204 (2002). 26. H. Miyagawa and L. T. Drzal, Polymer 45, 5163 5170 (2004). 27. L. Barral, J. Cano and J. Lopez, Polymer 41, 2657 2666 (2000). 28. M. Yin, J. A. Koutsky and T. L. Barr, Chem. Mater. 5, 1024 1031 (1993). 29. N. R. Raravikar, L. S. Schadler and A. Vijayaraghavan, Chem. Mater. 17, 974 983 (2005). 30. H. E. Kissinger, Anal. Chem. 29, 1702 1706 (1957). 31. T. Ozawa, Bull. Chem. Soc. Jpn. 38, 1881 1886 (1965). 32. C. D. Doyle, J. Appl. Poly. Sci. 24, 639 642 (1962). 33. J. H. Flynn and L. A. Wall, J. Res. Natl. Bur. Stand A Phys. Chem. 70A, 487 523 (1966). 34. W. Xu, S. Bao and S. Shen, J. Poly. Sci. B 41, 378 386 (2003). 35. Z. Yao, N. Braidy, G. A. Botton and A. Adronov, J. Am. Chem. Soc. 125, 16015 16024 (2003). 36. L. E. Nielsen and R. F. Landel, Mechanical Properties of Polymers and Composites, 2nd edn, p. 49. Marcel Dekker, New York, NY (1994).