Microcellular PP vs. Microcellular PP/MMT Nanocomposites: A Comparative Study of Their Mechanical Behavior

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

REGULAR CONTRIBUTED ARTICLES

S. J. A. Rizvi1,2, N. Bhatnagar2*
1 2

Department of Petroleum Studies, Faculty of Engineering & Technology, AMU, Aligarh, India Department of Mechanical Engineering, Indian Institute of Technology, Delhi, New Delhi, India

Microcellular PP vs. Microcellular PP/MMT Nanocomposites: A Comparative Study of their Mechanical Behavior
Microcellular foams have attracted both scientific and industrial community. Injection molding is perhaps most versatile industrial method of polymer processing. The microcellular injection molding is a newcomer variant in this class. It is capable to produce microcellular structure. The mechanical behavior of a microcellular polymer is different than conventional foams. Although they exhibit much improved mechanical properties than conventional foams but still they are inferior to solid polymer. Incorporation of nanoparticles in matrix may affect the nucleation and bubble growth process. Further, nanoparticles will enhance the performance (stiffness, strength etc.) of matrix. In present study an effort has been made to quantitatively analyze the improvements in mechanical properties of microcellular foam with the addition of nanoclay (MMT). Nanocomposite of polypropylene and nanoclay was prepared in a co-rotating twin screw extruder. Maleic anhydride grafted polypropylene (PP-g-MA) was used as compatibilizer. PP/nanocomposites were foamed in a microcellular injection molding machine, where supercritical nitrogen (SCFN2) was mixed with plasticized PP/ nanocomposite inside the screw barrel assembly. The molded foamed samples were tested for various mechanical properties. Experimental results show that a considerable improvement in the property of microcellular polypropylene by the addition of 2 wt.% nanoclay (MMT). the bubble nucleation process. Plate-like nanoparticles can also reduce gas diffusivity in the polymer matrix. In addition, the presence of nanoparticles may enhance mechanical and physical properties, the heat distortion temperature, and fire resistance of polymer foams (Lee et al., 2005). The high aspect ratio and large surface area of nanoparticles offer the potential for high reinforcing efficiency, good barrier properties, and improved dimensional and thermal stability. The nanometer dimension is especially beneficial for reinforcing foam materials, considering the thickness of cell walls is in the micron and submicron regime. It is therefore ideal to use nanoparticles to reinforce microstructures in order to achieve macroscale property improvement of the final products. Montmorillonite (MMT) is the clay most frequently used in the preparation of polymer nanocomposites (PNCs) (Utracki, 2004; Okamoto, 2007). The considerable attention received by that this filler derives from its potential to exfoliate in the polymer matrix and from its structure, which exhibits the required stiffness, strength and dimensional stability. MMT is a layered silicate with large surface-to-volume ratio (~ 700 m2/g) that exists naturally in a tactoid structure comprised of several tens of stacked layers. These layers have a typical lateral dimension of 0.1 0.5 lm and layer thickness plus interlayer spacing of about 1 nm (Rohlmann et al., 2008). Typically, two types of polymer/silicate nanocomposites may be prepared: the intercalated and the exfoliated or delaminated PNCs. The former are those in which the polymer molecules infiltrate the interlayer spacing of the tactoids preserving the stacked structure. The later are the ones in which the dispersal of nanolayers of MMT is achieved. One of the dimensions of the dispersed particles is in the nanometer range in both cases, even though in the intercalated PNCs is several times larger than in the exfoliated ones. Although both these structures often coexist in the PNCs, it is believed that the outstanding mechanical, flammability and barrier properties of these materials derive from the large aspect ratio of the exfoliated nanolayers (Utracki, 2004; Pinnavaia and Beall, 2000; Leszczynska et al., 2007). In spite of many attractive improvements in mechanical and physical properties of the polymer/(intercalated or exfoliated) clay nanocomposites, a significant drawback low fracture toughness has greatly limited their engineering application (Chan et al., 2002). 1

1 Introduction Polymeric foams are representative of a group of lightweight materials that have been widely used in a variety of industries with a market value of US $2 billion in 2000. However, the foam applications are limited by their inferior mechanical strength, poor surface quality, and low thermal and dimensional stability. A small amount of well-dispersed nanoparticles in the polymer may serve as nucleation sites to facilitate
* Mail address: &

Intern. Polymer Processing XXVI (2011) 4

Carl Hanser Verlag, Munich


IPP_ipp-2375 27.5.11/stm media kthen

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites In this paper, the effect of microcellular foaming on mechanical properties of neat polypropylene and PP/MMT (2 % by wt.) nanocomposites was studied. 2 Experimental 2.1 Materials Polypropylene (REPOL H110MA, Reliance Industries Ltd., [6]) having a MFI of 11g/10 min, as per ASTM D-1238 standard is used as matrix material. Maleic anhydride grafted polypropylene (MA-g-PP) is used as the compatibilizer. The MFI of MA-g-PP is 52.5 g/10 min. Montmorillonite nanoclay (Closite 15A) was obtained from Southern Clay Products Inc., USA. This was used as the nano-filler for PP-nanocomposite preparation. As informed by the manufacturing company, Closite15A was organically modified with dimethyl dyhyrogenated to allow quaternary ammonium ion, and the tallow composition is about 65 % C18, 30 % C16 and 5 % C14. The cation exchange capacity (CEC) of Closite 15A is 125 mEq/ 100 g; the corresponding organic modifier content is about 40 wt.%. 2.2 Preparation of Test Specimens For comparative study of the effect of nanoclay and microcellular foaming on the mechanical properties of polypropylene (PP), the following sets of specimens were prepared; 1. Neat solid polypropylene. 2. Microcellular foamed polypropylene. 3. Solid Polypropylene/MMT nanocomposite (2 % by wt.). 4. Microcellular foamed polypropylene/ MMT nanocomposite. 2.2.1 Compounding of PP/MMT Nanocomposites The melt intercalation offers a simple way of preparing nanocomposites. However, care has to be taken to fine tune the layered silicates surface chemistry in order to increase the silicate compatibility with the polymer matrix. Many studies have shown that the polar interactions of polymer and clay surface play a critical role in achieving particle dispersion. For non-polar polymers, e. g., PP, a polar compatibilizer such as maleic anhydride modified PP (PP-g-MA) is commonly added to improve the compatibility of PP and clay and thus the clay nanoparticle dispersion. All reported studies on PP nanocomposite foams were synthesized in this manner. Processing conditions such as shear rate and mixing have profound effects on the structure evolution of polymer nanocomposites by melt intercalation and these effects are still not well understood. In present study PP/nanocomposites were prepared via melt blending route in a twin-screw extruder from Thermo Fischer. A co-rotating twin-screw extruder with the characteristics; screw diameter 16 mm, barrel length 1.08 m, L/D 40, number of heating cylinders 6 and maximum screw speed 1 000 min 1. Pellets of PP and MA-g-PP were tumble mixed with the nanoclay (Closites15A) and simultaneously introduced into the hopper. Screw speed of 300 min 1 was kept constant for manufacture of nanocomposite material. The concentration of MMT nanoclay in nanocomposite was maintained at approximately 2.0 wt.% level. Ratio between the nanoclay and MAg-PP was 1 : 1.The extrudate was cut into pellets and then oven dried before being injection molding. 2.2.2 Injection Molding of Test Specimens Injection molding of test specimens were carried out on microcellular injection molding machine manufactured by Battenfeld Austria (model HM 40/ 210), see Fig. 1. Temperature profile in the barrel was maintained at 40, 165, 190, 220, and 240 8C from hopper to the nozzle end. The actual melt temperature was 215 8C. A cycle time of 60 s including cooling time of 45 s was common for all molding operation. This machine is capable of producing parts with microcellular foamed structure as well as solid parts. A family mold was used for preparation of tensile, flexural and impact test specimen as per ASTM D638-03 (Type-I), ASTM D790 and D256 standards. Physical blowing agents N2 gas (above its critical point) was used in microcellular injection molding process. Super critical N2 is termed as super critical fluid (SCF), which

B)

A)

C)

Fig. 1. Microcellular injection molding machine (A), shut-off nozzle (B) and ASTM test specimen mold (C)

2
IPP_ipp-2375 27.5.11/stm media kthen

Intern. Polymer Processing XXVI (2011) 4

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites

A)

B)

C)

D) Parameters

Fig. 2. Foam Microstructure: PP foam (A, B); PP/nanocomposite foam (C, D) Neat PP foam PP/MMT foam 52.7 6.55 109

is mixed with molten polymer in the machine barrel and injected into the mold. Super critical fluid is continuously fed to the barrel of injection molding machine with the help of two stage high pressure rotary compressor and high capacity gas accumulator. The screw design is also optimized for the single phase polymer/gas solution. Shut off nozzle is also incorporated to hold the polymer/ gas solution after plasticization and to avoid any unwanted pressure quench causing phase separation. 3 Results and Discussion 3.1 Foam Morphology Foamed tensile specimens (ASTM D-638, Type I) of where cryogenically fractured at the center of gauge length and examined under scanning electron microscope. Fig. 2 shows the image of fractured surface. The lower magnification (< 100) micrographs show the variation of cell population across the thickness of samples whereas higher magnification (500) was used to obtain micrographs for image analysis. Image analysis was carried out on image-J package. The results of image analysis are tabulated in Table 1. The effect of incorporation of nanoclay is reflected on morphological parameters of foam. In PP/MMT nanocomposite, the heterogeneous nucleation caused by the presence of nanoclay, is responsible for smaller cell size and higher cell density. From the image analysis of micrographs shown in Fig. 2B and D, the pore size distribution is plotted in Fig. 3. The PP/ MMT nanocomposite foam shows majority of cells lie in the range of 10 to 50 microns. In case of neat PP foam, bimodality is observed and the overall pore size distribution is much broad than PP/MMT nanocomposite foam. Intern. Polymer Processing XXVI (2011) 4

Average cell size (lm) Cell population density (cell per cm3) Table 1. Foam parameters

55.5 6.25 109

Fig. 3. Pore size distribution plot for PP foam and PP/nanocomposite foam

The micrograph shown in Fig. 4 was obtained from transmission electron microscopy (TEM) of PP/MMT nanocomposite along the flow direction. Since the PP/MMT nanocomposite compatibilized with PP-g-MA, was melt processed in a co rotating TSE (Thermo Fisher), a reasonably good intercalated structure was obtained. 3

IPP_ipp-2375 27.5.11/stm media kthen

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites

Fig. 4. TEM of PP/MMT nanocomposite

Fig. 6. Extension vs. load plot for PP and PP/nanoclay in foamed and solid conditions

3.2.1 Tensile Modulus Polymer nanocomposite foams exhibit substantially improved properties compared to their neat polymer foam counterparts. The tensile modulus of PP/clay nanocomposite foams has been measured and compared to neat PP foam. As shown in Fig. 8, the nanocomposite foams exhibit a much higher modulus as compared to neat PP foam and moderately higher than solid neat PP. A high tensile modulus means that the material is rigid more stress is required to produce a given amount of strain. In polymers, the tensile modulus and compressive modulus can be in a close or wide range. This variation may be 50 % or more, depending on resin type, reinforcing agents, and processing methods. The tensile and compressive moduli are often very close for metals. Fig. 7 shows the slope of load extension plots for PP and PP/nanoclay in foamed and solid conditions. Highest modulus is shown by PP/nanoclay in solid condition. Foamed PP shows the lowest modulus. It can be seen from Fig. 8 that foaming has adverse effect on tensile modulus. Addition of nano particles result into enhancement in tensile modulus however foamed PP/ nanoclay samples show higher modulus than neat polypropylene by 5 %.

Fig. 5. XRD of solid and foamed PP/MMT nanocomposite

The X-ray diffraction (XRD) was carried for the solid and foamed specimen of PP/MMT nanocomposite. The diffractograms (I vs. 2h) for solid and foamed specimen of PP/MMT nanocomposite are shown in Fig. 5. The very small peak at 2.73nm in solid PP/MMT nanocomposite followed by peak at 1.34 nm justifies the good dispersion of nanoclay in compatibilized polypropylene matrix. In case of foamed PP/MMT nanocomposite, absence of peak around 2.7 nm indicates better dispersion on nanoclay. Further the shift of peak to left (at 1.35 nm) in foamed PP/MMT nanocomposite, confirms the increase in d-spacing. During the melt processing (microcellular injection molding), the reduced viscosity of PP matrix, caused by plasticization effect of super critical nitrogen, helps the dispersion of nanoclay and leads to exfoliated nanostructure.

3.2 Tensile Properties The tensile tests were carried out at room temperature. Instron 5582 testing machine equipped with a 100 kN load cell was used. Tensile testing was conducted according to ASTM D638-03 (Type-I) with a crosshead speed of 5 mm/min for all the samples. Five samples were tested for each type of sample. Foamed samples of polypropylene and PP/nanoclay were stored for at least 10 to 15 days before testing so that dissolved nitrogen gas should escape from polymer matrix. This step is important because tensile testing with freshly molded microcellular specimen may lead to error in computation of % elongation at break and toughness because of internal plasticization effect of physical blowing agent N2. This step is also valid for testing of flexural and impact properties. Fig. 6 shows the tensile plot for PP and PP/nanoclay in foamed and solid conditions. 4
IPP_ipp-2375 27.5.11/stm media kthen

Fig. 7. Effect of foaming on tensile modulus

Intern. Polymer Processing XXVI (2011) 4

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites duced toughness value for both solid and foamed PP/MMT nanocomposite in comparison to their neat PP counterparts respectively.

3.2.3 Energy to Yield Point and Break Point Energy required for deformation of tensile specimen within elastic limit and at the rupture point is called Energy to Yield Point and Energy to Break Point. Fig. 11 and 12 shows these values for PP and PP/ nanoclay in foamed and solid conditions. Energy at yield point and break point decreases with foaming. Since foaming increases the flexibility therefore energy requirement to cause rupture is low. Similar effect can be observed for PP/ nanoclay nanocomposite. But very low reduction in yield energy value for foamed and solid PP/nanoclay composite material is observed. The foamed samples of neat PP and PP/MMT nanocomposite show contradictory observation in Fig. 11 and 12. The energy requirement for the deformation upto yield point, for foamed PP/MMT is much higher (0.539 J) than that of foamed neat PP (0.371 J). But when these samples are deformed until break point, the neat PP foam dissipates more energy (9.566 J) as compared to PP/MMT nanocomposite (7.97 J). But when this contradictory information is analyzed in conjunction with stress vs. strain curve (Fig. 9) and toughness data (Fig. 10), it can be explained that the foamed PP/MMT nanocomposite offers a stiff resistance to deformation (until yield point) due to enhancement in modulus by the incorporation of nanoclay into PP matrix. Beyond the yield point, the plastic deformation sets in and more ductile porous matrix of neat PP foam exhibits larger elongation (Fig. 15) hence higher toughness (Fig. 10) than foamed PP/MMT nanocomposite. Addition of nanoclay in PP

Fig. 8. Tensile modulus values for PP and PP/nanoclay in foamed and solid conditions

3.2.2 Toughness Toughness can be defined as ability of material to absorb mechanical (or kinetic) energy up to failure. Toughness can be found by taking the area under the stress-strain curve. Fig. 9 shows the stress-strain curve for PP and PP/nanoclay under foamed and solid conditions. Toughness was calculated and values are plotted in Fig. 10. Highest toughness is found in neat polypropylene. It can be concluded from the Fig. 10 that microcellular foaming leads to flexibility and hence reduction in toughness. Further it can be noted that nano particles enhances the modulus of matrix and reduces the maximum strain as shown in Fig. 15, therefore the area under the stress-strain curve reduces leading to re-

Fig. 9. Effect of foaming on toughness

Fig. 11. Energy at yield point

Fig. 10. Toughness values for PP and PP/nanoclay in foamed and solid conditions

Fig. 12. Energy at break point

Intern. Polymer Processing XXVI (2011) 4


IPP_ipp-2375 27.5.11/stm media kthen

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites

Fig. 13. Ultimate tensile strength (MPa) values of PP and PP/nanoclay under foamed and solid conditions

Fig. 15. Maximum % Strain for PP and PP/nanoclay in foamed and solid conditions

3.3 Flexural Properties The flexural strength of a material is defined as its ability to resist deformation under lateral loading. For materials that deform significantly but do not break, the load at yield, typically measured at 5 % deformation/strain of the outer surface, is reported as the flexural strength or flexural yield strength. The test beam is under compressive stress at the concave surface and tensile stress at the convex surface. The value represents the highest stress experienced within the material at its moment of rupture. In a bending test, the highest stress is reached on the surface of the sample. For the measurement of flexural properties of PP and PP/nanoclay (foamed and solid samples) ASTM D790 standard was followed. Flexural testing was carried out on a three point flexural testing attachment supplied by Zwicks universal testing machine. All the tests were performed at room temperature (25 8C maintained by air conditioner). Cross head speed was 50 mm/min and the support separation was 70 mm for all flexural tests. Fig. 16 and 17 shows the flexural modulus and

Fig. 14. Strength at break (MPa) values of PP and PP/nanoclay under foamed and solid conditions

matrix improves the tensile modulus that reflects in term of enhanced ultimate tensile strength (Fig. 13) and strength at break point (Fig. 14) for PP/MMT nanocomposites under solid and foamed conditions. On the other hand the ductility of matrix is sacrificed for improvement in stiffness due to the addition of nanoclay in PP matrix. Since toughness is total energy required for the rupture that can be quantitatively estimated as area under the stress/strain curve, a ductile matrix will exhibit higher toughness than a stiff matrix with lesser elongation at break point. 3.2.4 Ultimate Tensile Strength The maximum stress a material can withstand when subjected to tension. It is the maximum stress on the stress strain curve. Fig. 13 and 14 show the ultimate tensile strength (MPa) and strength at break values for PP and PP/nanoclay in foamed and solid conditions. It has been observed that ultimate tensile strength and strength at break decreases with foaming. Similar findings were published by Hwang et al. (2009). The respective values of ultimate tensile strength and strength at break increases with incorporation of nanoclay for both foamed and solid samples. 3.2.5 Maximum Percentage Strain The ability of material to undergo elongation can be expressed by the term maximum % strain. Higher value of max. % strain reflects the ability of material to elongate more before final rupture or break up. Fig. 15 shows the maximum % strain for PP and PP/nanoclay in foamed and solid conditions. 6

Fig. 16. Flexural Modulus for PP and PP/nanoclay in foamed and solid conditions

Fig. 17. Flexural Strength for PP and PP/nanoclay in foamed and solid conditions

Intern. Polymer Processing XXVI (2011) 4


IPP_ipp-2375 27.5.11/stm media kthen

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites strength for PP and PP/ nanoclay in foamed and solid conditions. Decrease in flexural modulus and strength can be observed in foamed samples in both the figures. As expected there is remarkable increase in flexural modulus and strength when nanoclay is added in small quantity i. e. 2 % by wt. By the incorporation of nanoclay in polypropylene, enhancement in flexural strength as shown in Fig. 16 is much higher than the increase in flexural modulus as shown in Fig. 16. It has been found that the flexural properties are relatively unaffected by changes in weight reduction due to foaming of sample. This relative insensitivity of flexural modulus to weight reduction is a primary reason why the process is effective for structural applications. The other benefit in structural parts is the fact that relatively small changes in flexural modulus have little effect on overall part performance. more uniform bubble nucleation with higher cell density and smaller cell diameter. The adverse effect of foaming on mechanical properties is less in PP/nanocomposite than neat PP. Further incorporation of the nanoclay in PP matrix also strengthen matrix and improves the impact strength. It is obvious from the above graph that impact strength decreases with foaming for neat polypropylene and PP/nanoclay. As expected addition of nanoclay 2 % by wt. increases the impact strength of microcellular polypropylene. 4 Conclusion It has been observed that most of the mechanical properties including impact strength degrade with foaming. Firstly incorporation of very low amount of nanoclay significantly improves the mechanical properties of the foamed polymer. This is because of the high nucleation efficiency, nanoparticles provide a powerful way to increase cell density and reduce cell size. The image analysis of micrographs, see Table 1, confirm that presence of nanoparticles increases the cell density and reduces the average size of cell. Secondly melt processing of nanocomposite, in presence of supercritical nitrogen, improves the dispersion of nanoclay in compatibilized PP matrix. The XRD plot (Fig. 5) shows shift of the peak to higher d-spacing value. This indicates improved dispersion of nanoclay. It can be concluded that microcellular foams having tiny cells in large numbers can sustain the stress in a better way as compared to foams having bigger cell size and less cell density. This is particularly beneficial for the production of microcellular foams. Microcellular foams have been considered as a lightweight and high strength material for structural applications. Incorporation of nanoclay may also be helpful in creation of open cellular morphology along with suitable leaching agents like PVOH and NaCl. Open cell morphology is highly desirable for fabrication of scaffolds for tissue engineering. Nanoparticles may be helpful in generating open cell foams under external fields such as ultrasonic fields, because nanoparticles may behave like a stress concentrator. Further it can be concluded that the toughness, energy at break point, maximum % strain for PP/MMT microcellular nanocomposite is inferior to microcellular polypropylene. Therefore it can be concluded that mechanical properties of microcellular polypropylene are inferior to neat polypropylene but improvement in most of the mechanical properties of microcellular polypropylene can be obtain by addition of small amount of MMT nanoclay. References
Chan, C.-M., et al., Polypropylene/Calcium Carbonate Nanocomposites, Polymer, 43, 2981 2992 (2002), DOI:10.1016/S0032-3861(02)00120-9 Hwang, S.-S., et al., Effect of Clay and Compatibilizer on Mechanical/Thermal Properties of Microcellular Injection Molded Low Density Polyethylene Nanocomposites, Int. Commun.Heat Mass Trans., 36, 471 479 (2009), DOI:10.1016/j.icheatmasstransfer.2009.02.005 Hwang, S.-S., et al., Effect of Organoclay on the Mechanical/Thermal Properties of Microcellular Injection Molded PBT-clay Nanocomposites, Int. Commun. Heat Mass Transfer, 37, 1036 1043 (2010), DOI:10.1016/j.icheatmasstransfer.2010.06.010

3.4 Impact Properties Impact properties are highly dependent upon the microstructure of foamed polypropylene. Since polypropylene is a highly crystalline polymer and its impact strength is contributed by crystalline regions. Size and number of spherulities defining the crystal structure are responsible for impact properties. A preliminary study of impact experiments on microcellular polystyrene was conducted by Waldman (1982), the author measured the notched Charpy impact toughness of microcellular polystyrene foams and found it to be higher than that of the neat polymer, while in a falling weight experiment the impact toughness was reported to be less than that of the neat polymer matrix. Impact strength is calculated as energy per unit thickness required breaking a test specimen under flexural impact. Alternatively, the results may be reported as energy lost per unit crosssectional area at the notch (J/m2). In Europe, ISO 180 methods are used and results are based only on the cross-sectional area at the notch (J/m2). Test specimen is held as a vertical cantilevered beam and is impacted by a swinging pendulum. The energy lost by the pendulum is equated with the energy absorbed by the test specimen. A notch of 458 was cut on every sample with the help of notch cutting machine. Fig. 18 shows the result for Izod impact strength for PP and PP/nanoclay under foamed and solid conditions. The presence of voids in polymer matrix results in crack initiation and propagation. Therefore it has been found that impact strength of neat PP and PP/MMT nanocomposite under foamed condition is lower than their solid counterparts. Since, nanoparticles cause

Fig. 18. Impact Strength for PP and PP/nanoclay in foamed and solid conditions

Intern. Polymer Processing XXVI (2011) 4


IPP_ipp-2375 27.5.11/stm media kthen

S. J. A. Rizvi, N. Bhatnagar: Microcellular PP vs. Microcellular PP/MMT Nanocomposites


Lee, L. J., et al., Polymer Nanocomposite Foams, Compos. Sci. Tech., 65, 2344 2363, (2005), DOI:10.1016/j.compscitech.2005.06.016 Leszczynska, A., et al., Polymer/Montmorillonite Nanocomposites with Improved Thermal Properties Part I Factors Influencing Thermal Stability and Mechanisms of Thermal Stability Improvement, Thermochim Acta, 453, 75 96 (2007), DOI:10.1016/j.tca.2006.11.002 Okamoto, M.: Macromolecular Engineering in Precise Synthesis, Materials Properties, Applications, Wiley-VCH, Weinheim (2007) Pinnavaia, T. J., Beall, G. W., Polymerclay Nanocomposites, John Wiley and Sons, New York (2000) Rohlmann, C. O., et al., Comparative Analysis of Nanocomposites Based on Polypropylene and Different Montmorillonites, Eur. Polym.J., 44, 2749 2760, (2008), DOI:10.1016/j.eurpolymj.2008.07.006 Shyh-shin Hwang et al., Effect of organoclay on the mechanical/ thermal properties of microcellular injection molded polystyrene-clay nanocomposites, Int. Communications in Heat and Mass Transfer, 36, 799 805 (2009), DOI:10.1016/j.icheatmasstransfer.2009.06.011 Utracki, L. A., Clay-containing Polymeric Nanocomposites, Shawburi: Rapra Technology, ?New Dehli? (2004) Waldman, F.A., The Processing of Microcellular Foam, Phd Thesis, Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA, (1982) Yuan, M. J., et al., Microcellular Nanocomposite Injection Molding Process, SPE ANTEC Tech. Papers, 691 695 (2003)

Date received: May 12, 2010 Date accepted: February 20, 2011

Bibliography DOI 10.3139/217.2375 Intern. Polymer Processing XXVI (2011) 4; page 1 8 Carl Hanser Verlag GmbH & Co. KG ISSN 0930-777X You will find the article and additional material by entering the document number IPP2375 on our website at www.polymer-process.com

8
IPP_ipp-2375 27.5.11/stm media kthen

Intern. Polymer Processing XXVI (2011) 4

You might also like