Nanocomposite

From Wikipedia, the free encyclopedia

A nanocomposite is a multiphase solid material where one of the phases has one, two or three dimensions of less than 100 nanometers (nm), or structures having nano-scale repeat distances between the different phases that make up the material.[1] In the broadest sense this definition can include porous media, colloids, gels and copolymers, but is more usually taken to mean the solid combination of a bulk matrix and nano-dimensional phase(s) differing in properties due to dissimilarities in structure and chemistry. The mechanical, electrical, thermal, optical, electrochemical, catalytic properties of the nanocomposite will differ markedly from that of the component materials. Size limits for these effects have been proposed,[2] <5 nm for catalytic activity, <20 nm for making a hard magnetic material soft, <50 nm for refractive index changes, and <100 nm for achieving superparamagnetism, mechanical strengthening or restricting matrix dislocation movement.

Nanocomposites are found in nature, for example in the structure of the abalone shell and bone. The use of nanoparticle-rich materials long predates the understanding of the physical and chemical nature of these materials. Jose-Yacaman et al. [3] investigated the origin of the depth of colour and the resistance to acids and bio-corrosion of Maya blue paint, attributing it to a nanoparticle mechanism. From the mid-1950s nanoscale organo-clays have been used to control flow of polymer solutions (e.g. as paint viscosifiers) or the constitution of gels (e.g. as a thickening substance in cosmetics, keeping the preparations in homogeneous form). By the 1970s polymer/clay composites were the topic of textbooks,[4] although the term "nanocomposites" was not in common use.

In mechanical terms, nanocomposites differ from conventional composite materials due to the exceptionally high surface to volume ratio of the reinforcing phase and/or its exceptionally high aspect ratio. The reinforcing material can be made up of particles (e.g. minerals), sheets (e.g. exfoliated clay stacks) or fibres (e.g. carbon nanotubes or electrospun fibres). The area of the interface between the matrix and reinforcement phase(s) is typically an order of magnitude greater than for conventional composite materials. The matrix material properties are significantly affected in the vicinity of the reinforcement. Ajayan et al. [1] note that with polymer nanocomposites, properties related to local chemistry, degree of thermoset cure, polymer chain mobility, polymer chain conformation, degree of polymer chain ordering or crystallinity can all vary significantly and continuously from the interface with the reinforcement into the bulk of the matrix.

This large amount of reinforcement surface area means that a relatively small amount of nanoscale reinforcement can have an observable effect on the macroscale properties of the composite. For example, adding carbon nanotubes improves the electrical and thermal conductivity. Other kinds of nanoparticulates may result in enhanced optical properties, dielectric properties, heat resistance or mechanical properties such as stiffness, strength and resistance to wear and damage. In general, the nano reinforcement is dispersed into the matrix during processing. The percentage by weight (called mass fraction) of the nanoparticulates introduced can remain very low (on the order of 0.5% to 5%) due to the low filler percolation threshold, especially for the most commonly used non-spherical, high aspect ratio fillers (e.g. nanometer-thin platelets, such as clays, or nanometer-diameter cylinders, such as carbon nanotubes).

Ceramic-matrix nanocomposites

In this group of composites the main part of the volume is occupied by a ceramic, i.e. a chemical compound from the group of oxides, nitrides, borides, silicides etc.. In most cases, ceramic-matrix nanocomposites encompass a metal as the second component. Ideally both components, the metallic one and the ceramic one, are finely dispersed in each other in order to elicit the particular nanoscopic properties. Nanocomposite from these combinations were demonstrated in improving their optical, electrical and magnetic properties [5] as well as tribological, corrosion-resistance and other protective properties.[6]

The binary phase diagram of the mixture should be considered in designing ceramic-metal nanocomposites and measures have to be taken to avoid a chemical reaction between both components. The last point mainly is of importance for the metallic component that may easily react with the ceramic and thereby lose its metallic character. This is not an easily obeyed constraint, because the preparation of the ceramic component generally requires high process temperatures. The most safe measure thus is to carefully choose immiscible metal and ceramic phases. A good example for such a combination is represented by the ceramic-metal composite of TiO2 and Cu, the mixtures of which were found immiscible over large areas in the Gibbs’ triangle of Cu-O-Ti.[7]

The concept of ceramic-matrix nanocomposites was also applied to thin films that are solid layers of a few nm to some tens of µm thickness deposited upon an underlying substrate and that play an important role in the functionalization of technical surfaces. Gas flow sputtering by the hollow cathode technique turned out as a rather effective technique for the preparation of nanocomposite layers. The process operates as a vacuum-based deposition technique and is associated with high deposition rates up to some µm/s and the growth of nanoparticles in the gas phase. Nanocomposite layers in the ceramics range of composition were prepared from TiO2 and Cu by the hollow cathode technique [8] that showed a high mechanical hardness, small coefficients of friction and a high resistance to corrosion.

Metal-matrix nanocomposites

Metal matrix nanocomposites can also be defined as reinforced metal matrix composites. This type of composites can be classified as continuous and non-continuous reinforced materials. One of the more important nanocomposites is Carbon nanotube metal matrix composites, which is an emerging new material that is being developed to take advantage of the high tensile strength and electrical conductivity of carbon nanotube materials. Critical to the realization of CNT-MMC possessing optimal properties in these areas are the development of synthetic techniques that are (a) economically producible, (b) provide for a homogeneous dispersion of nanotubes in the metallic matrix, and (c) lead to strong interfacial adhesion between the metallic matrix and the carbon nanotubes. In addition to carbon nanotube metal matrix composites, boron nitride reinforced metal matrix composites and carbon nitride metal matrix composites are the new research areas on metal matrix nanocomposites.[9]

A recent study, comparing the mechanical properties (Young's modulus, compressive yield strength, flexural modulus and flexural yield strength) of single- and multi-walled reinforced polymeric (polypropylene fumarate—PPF) nanocomposites to tungsten disulfide nanotubes reinforced PPF nanocomposites suggest that tungsten disulfide nanotubes reinforced PPF nanocomposites possess significantly higher mechanical properties and tungsten disulfide nanotubes are better reinforcing agents than carbon nanotubes.[10] Increases in the mechanical properties can be attributed to a uniform dispersion of inorganic nanotubes in the polymer matrix (compared to carbon nanotubes that exist as micron sized agggregates) and increased crosslinking density of the polymer in the presence of tungsten disulfide nanotubes (increase in crosslinking density leads to an increase in the mechanical properties). These results suggest that inorganic nanomaterials, in general, may be better reinforcing agents compared to carbon nanotubes.

Another kind of nanocomposite is the energetic nanocomposite, generally as a hybrid sol–gel with a silica base, which, when combined with metal oxides and nano-scale aluminum powder, can form superthermite materials.[11][12][13]

Polymer-matrix nanocomposites

In the simplest case, appropriately adding nanoparticulates to a polymer matrix can enhance its performance, often dramatically, by simply capitalizing on the nature and properties of the nanoscale filler [14] (these materials are better described by the term nanofilled polymer composites [14]). This strategy is particularly effective in yielding high performance composites, when good dispersion of the filler is achieved and the properties of the nanoscale filler are substantially different or better than those of the matrix. An example of this would be reinforcing a polymer matrix by much stiffer nanoparticles [15][16] of ceramics, clays, or carbon nanotubes. It should be noted that the improvement in mechanical properties may not be limited to stiffness or strength. Time-dependent properties could be improved by addition of the nanofillers.[17] Alternatively, the enhanced crystallization behavior under flow conditions [18] and other physical properties of high performance nanocomposites may be mainly due to the high aspect ratio and/or the high surface area of the fillers,[19][20] since nanoparticulates have extremely high surface area to volume ratios when good dispersion is achieved. Nanoparticle dispersibility in the polymer matrix is a key issue, which limits the appliciable particle volume fraction and there for also the multi-functionality of the composite material.[21] Recent research on thin films (thickness <50 micrometer) made of polymer nanocomposites has resulted in a new and scalable synthesis technique, which allows the facile incorporation of greater nano-material quantities .[21] Such advances will enable the future development of multi-functional small scale devices (i.e. sensor, actuator, medical equipment), which rely on polymer nanocomposites.

Nanoscale dispersion of filler or controlled nanostructures in the composite can introduce new physical properties and novel behaviors that are absent in the unfilled matrices. This effectively changes the nature of the original matrix [14] (such composite materials can be better described by the term genuine nanocomposites or hybrids [14]). Some examples of such new properties are fire resistance or flame retardancy,[22] and accelerated biodegradability.

In a recent study, polymeric nanocomposites were fabricated using various one-dimensional carbon nanostructures such as single- and multi- walled carbon nanotubes, as well as two-dimensional carbon and inorganic nanomaterials such as graphene platelets,[23] graphene nanoribbon,[24] single- and multi- walled graphene oxide nanoribbons, graphene oxide nanoplatelets and molybdenum disulfide nanoplatelets as reinforcing agents, to improve the mechanical properties of poly(propylene fumarate) nanocomposites, for bone tissue engineering applications.[25] Significant mechanical reinforcement (i.e. increases in the Young's modulus, compressive yield strength, flexural modulus and flexural yield strength) were observed at low loading concentrations (0.01-0.2 wt%) of nanomaterials. The results suggest that mechanical reinforcement is dependent on the nanostructure morphology, defects, dispersion of nanomaterials in the polymer matrix, and the cross-linking density of the polymer. In general, two-dimensional nanostructures can reinforce the polymer better than one-dimensional nanostructures, and inorganic nanomaterials are better reinforcing agents than carbon based nanomaterials. In addition to mechanical properties, multi-walled carbon nanotubes based polymer nanocomposites have also been used for the enhancement of the electrical conductivity.[26]

See also

References

  1. 1.0 1.1 P.M. Ajayan, L.S. Schadler, P.V. Braun (2003). Nanocomposite science and technology. Wiley. ISBN 3-527-30359-6. 
  2. Kamigaito, O, What can be improved by nanometer composites? J. Jpn. Soc. Powder Powder Metall. 38:315-21, 1991 in Kelly, A, Concise encyclopedia of composites materials, Elsevier Science Ltd, 1994
  3. Jose-Yacaman, M.; Rendon, L.; Arenas, J.; Serra Puche, M. C. (1996). "Maya Blue Paint: An Ancient Nanostructured Material". Science 273 (5272): 223–5. doi:10.1126/science.273.5272.223. PMID 8662502. 
  4. B.K.G. Theng "Formation and Properties of Clay Polymer Complexes", Elsevier, NY 1979; ISBN 978-0-444-41706-0
  5. F. E. Kruis, H. Fissan and A. Peled (1998). "Synthesis of nanoparticles in the gas phase for electronic, optical and magnetic applications – a review". J. Aerosol Sci. 29 (5–6): 511–535. doi:10.1016/S0021-8502(97)10032-5. 
  6. S. Zhang, D. Sun, Y. Fu and H. Du (2003). "Recent advances of superhard nanocomposite coatings: a review". Surf. Coat. Technol. 167 (2–3): 113–119. doi:10.1016/S0257-8972(02)00903-9. 
  7. G. Effenberg, F. Aldinger and P. Rogl (2001). Ternary Alloys. A Comprehensive Compendium of Evaluated Constitutional Data and Phase Diagrams. Materials Science-International Services. ISBN status = May be invalid - please double check Check |isbn= value (help). 
  8. M. Birkholz, U. Albers, and T. Jung (2004). "Nanocomposite layers of ceramic oxides and metals prepared by reactive gas-flow sputtering". Surf. Coat. Technol. 179 (2–3): 279–285. doi:10.1016/S0257-8972(03)00865-X. 
  9. S. R. Bakshi, D. Lahiri, and A. Argawal, Carbon nanotube reinforced metal matrix composites - A Review, International Materials Reviews, vol. 55, (2010), http://web.eng.fiu.edu/agarwala/PDF/2010/12.pdf
  10. Lalwani, G; Henslee, AM; Farshid, B; Parmar, P; Lin, L; Qin, YX; Kasper, FK; Mikos, AG; Sitharaman, B (September 2013). "Tungsten disulfide nanotubes reinforced biodegradable polymers for bone tissue engineering.". Acta biomaterialia 9 (9): 8365–73. PMID 23727293. 
  11. Gash, AE. "Making nanostructured pyrotechnics in a Beaker" (pdf). Retrieved 2008-09-28. 
  12. Gash, AE. "Energetic nanocomposites with sol-gel chemistry: synthesis, safety, and characterization, LLNL UCRL-JC-146739" (pdf). Retrieved 2008-09-28. 
  13. Ryan, Kevin R.; Gourley, James R.; Jones, Steven E. (2008). "Environmental anomalies at the World Trade Center: evidence for energetic materials". The Environmentalist 29: 56. doi:10.1007/s10669-008-9182-4. 
  14. 14.0 14.1 14.2 14.3 Manias, Evangelos (2007). "Nanocomposites: Stiffer by design". Nature Materials 6 (1): 9–11. doi:10.1038/nmat1812. PMID 17199118. 
  15. Y. Mai, Z. Yu, (2006). Y. Mai, Z. Yu, ed. Polymer Nanocomposites. Woodhead Publ. ISBN 978-1-85573-969-7. 
  16. "Polymer-Clay Nanocomposites", T. J. Pinnavaia, G. W. Beall (eds.), Wiley, 2001; ISBN 978-0-471-63700-4
  17. Zandiatashbar, Ardavan; Picu, Catalin R.; Koratkar, Nikhil (2012). "Control of Epoxy Creep Using Graphene". Small 8 (11): 1676–1682. doi:10.1002/smll.201102686. 
  18. Nilesh Patil, Luigi Balzano, Giuseppe Portale and Sanjay Rastogi (2010). "Influence of shear in the crystallization of polyethylene in the presence of SWCNTs". Carbon 48 (14): 4116. doi:10.1016/j.carbon.2010.07.022. 
  19. Usuki, Arimitsu; Kawasumi, Masaya; Kojima, Yoshitsugu; Okada, Akane; Kurauchi, Toshio; Kamigaito, Osami (1993). "Swelling behavior of montmorillonite cation exchanged for ω-amino acids by ∊-caprolactam". Journal of Materials Research 8 (5): 1174. doi:10.1557/JMR.1993.1174. 
  20. Usuki, Arimitsu; Kojima, Yoshitsugu; Kawasumi, Masaya; Okada, Akane; Fukushima, Yoshiaki; Kurauchi, Toshio; Kamigaito, Osami (1993). "Synthesis of nylon 6-clay hybrid". Journal of Materials Research 8 (5): 1179. doi:10.1557/JMR.1993.1179. 
  21. 21.0 21.1 Sotirou, Georgios A.; Blattmann, Christoph O.; Pratsinis, Sotiris E. (2013). "Flexible, multifunctional, magnetically actuated nanocomposite films". Advanced Functional Materials 23 (1): 1616–3028. doi:10.1002/adfm.201201371. 
  22. "Flame Retardant Polymer Nanocomposites" A. B. Morgan, C. A. Wilkie (eds.), Wiley, 2007; ISBN 978-0-471-73426-0
  23. M.A. Rafiee et. al, Enhanced Mechanical Properties of Nanocomposites at Low Graphene Content, ACS Nano, 3 (12), (2009) 3884-90. [pdf]
  24. M.A. Rafiee et al. graphene nanoribbon composite, ACS Nano 4 (12) (2010) 7415–20.
  25. Lalwani, Gaurav; Allan M. Henslee, Behzad Farshid, Liangjun Lin, F. Kurtis Kasper, Yi-Xian Qin, Antonios G. Mikos, and Balaji Sitharaman (February 13, 2013). "Two-Dimensional Nanostructure-Reinforced Biodegradable Polymeric Nanocomposites for Bone Tissue Engineering". Biomacromolecules. doi:10.1021/bm301995s. PMID 23405887. 
  26. Singh BP et al. <http://link.springer.com/article/10.1007/s11671-008-9179-4#page-1f>

Further reading

This article is issued from Wikipedia. The text is available under the Creative Commons Attribution/Share Alike; additional terms may apply for the media files.