Jump to content

Nanoparticle: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Reverted edits by 203.26.38.39 (talk) to last version by 207.6.249.104
Line 20: Line 20:
[[Michael Faraday]] provided the first description, in scientific terms, of the optical properties of nanometer-scale metals in his classic 1857 paper. In a subsequent paper, the author (Turner) points out that: "It is well known that when thin leaves of gold or silver are mounted upon glass and heated to a temperature which is well below a red heat (~500 °C), a remarkable change of properties takes place, whereby the continuity of the metallic film is destroyed. The result is that white light is now freely transmitted, reflection is correspondingly diminished, while the electrical resistivity is enormously increased."
[[Michael Faraday]] provided the first description, in scientific terms, of the optical properties of nanometer-scale metals in his classic 1857 paper. In a subsequent paper, the author (Turner) points out that: "It is well known that when thin leaves of gold or silver are mounted upon glass and heated to a temperature which is well below a red heat (~500 °C), a remarkable change of properties takes place, whereby the continuity of the metallic film is destroyed. The result is that white light is now freely transmitted, reflection is correspondingly diminished, while the electrical resistivity is enormously increased."
<ref>{{cite journal|journal = Phil. Trans. Roy. Soc. London|title = Experimental relations of gold (and other metals) to light|author = Faraday, Michael|volume = 147 |year=1857|pages=145–181|doi = 10.1098/rstl.1857.0011}}</ref><ref>{{cite journal|author=Beilby, G.T.|journal=Proc. Roy. Soc. A|page=226|year=1903|volume=72|title=The Effects of Heat and of Solvents on Thin Films of Metal|url=http://www.jstor.org/pss/116470}}</ref><ref>{{cite journal|author=Turner, T.|type=Transparent silver and other metallic films|journal=Proc. Roy. Soc. Lond. A|volume=81|page=301|year=1908|url=http://www.jstor.org/pss/93060|title=Transparent Silver and Other Metallic Films|issue=548}}</ref>
<ref>{{cite journal|journal = Phil. Trans. Roy. Soc. London|title = Experimental relations of gold (and other metals) to light|author = Faraday, Michael|volume = 147 |year=1857|pages=145–181|doi = 10.1098/rstl.1857.0011}}</ref><ref>{{cite journal|author=Beilby, G.T.|journal=Proc. Roy. Soc. A|page=226|year=1903|volume=72|title=The Effects of Heat and of Solvents on Thin Films of Metal|url=http://www.jstor.org/pss/116470}}</ref><ref>{{cite journal|author=Turner, T.|type=Transparent silver and other metallic films|journal=Proc. Roy. Soc. Lond. A|volume=81|page=301|year=1908|url=http://www.jstor.org/pss/93060|title=Transparent Silver and Other Metallic Films|issue=548}}</ref>

Scientist Jason Stufflebeam is one of the leading researchers of Nanoparticles in the state of Georgia.


==Uniformity==
==Uniformity==

Revision as of 11:49, 28 October 2010

In nanotechnology, a particle is defined as a small object that behaves as a whole unit in terms of its transport and properties. It is further classified according to size: in terms of diameter, fine particles cover a range between 100 and 2500 nanometers, while ultrafine particles, on the other hand, are sized between 1 and 100 nanometers. Similar to ultrafine particles, nanoparticles are sized between 1 and 100 nanometers. Nanoparticles may or may not exhibit size-related properties that differ significantly from those observed in fine particles or bulk materials.[1][2] Although the size of most molecules would fit into the above outline, individual molecules are usually not referred to as nanoparticles.

Nanoclusters have at least one dimension between 1 and 10 nanometers and a narrow size distribution. Nanopowders[3] are agglomerates of ultrafine particles, nanoparticles, or nanoclusters. Nanometer-sized single crystals, or single-domain ultrafine particles, are often referred to as nanocrystals. Nanoparticle research is currently an area of intense scientific interest due to a wide variety of potential applications in biomedical, optical and electronic fields.

TEM (a, b, and c) images of prepared mesoporous silica nanoparticles with mean outer diameter: (a) 20nm, (b) 45nm, and (c) 80nm. SEM (d) image corresponding to (b). The insets are a high magnification of mesoporous silica particle.[4]

The National Nanotechnology Initiative has led to generous public funding for nanoparticle research in the United States.

Background

Although nanoparticles are generally considered an invention of modern science, they actually have a very long history. Nanoparticles were used by artisans as far back as the 9th century in Mesopotamia for generating a glittering effect on the surface of pots[citation needed].

Even these days, pottery from the Middle Ages and Renaissance often retain a distinct gold or copper colored metallic glitter. This so called luster is caused by a metallic film that was applied to the transparent surface of a glazing. The luster can still be visible if the film has resisted atmospheric oxidation and other weathering.

The luster originated within the film itself, which contained silver and copper nanoparticles dispersed homogeneously in the glassy matrix of the ceramic glaze. These nanoparticles were created by the artisans by adding copper and silver salts and oxides together with vinegar, ochre and clay, on the surface of previously-glazed pottery. The object was then placed into a kiln and heated to about 600 °C in a reducing atmosphere.

In the heat the glaze would soften, causing the copper and silver ions to migrate into the outer layers of the glaze. There the reducing atmosphere reduced the ions back to metals, which then came together forming the nanoparticles that give the colour and optical effects.

Luster technique showed that ancient craftsmen had a rather sophisticated empirical knowledge of materials. The technique originated in the islamic world. As Muslims were not allowed to use gold in artistic representations, they had to find a way to create a similar effect without using real gold. The solution they found was using luster.[5]

Michael Faraday provided the first description, in scientific terms, of the optical properties of nanometer-scale metals in his classic 1857 paper. In a subsequent paper, the author (Turner) points out that: "It is well known that when thin leaves of gold or silver are mounted upon glass and heated to a temperature which is well below a red heat (~500 °C), a remarkable change of properties takes place, whereby the continuity of the metallic film is destroyed. The result is that white light is now freely transmitted, reflection is correspondingly diminished, while the electrical resistivity is enormously increased." [6][7][8]

Scientist Jason Stufflebeam is one of the leading researchers of Nanoparticles in the state of Georgia.

Uniformity

File:Coll 3.jpg
Colloidal crystal composed of amorphous hydrated colloidal silica (particle diameter 600 nm)

The chemical processing and synthesis of high performance technological components for the private, industrial and military sectors requires the use of high purity ceramics, polymers, glass-ceramics and material composites. In condensed bodies formed from fine powders, the irregular particle sizes and shapes in a typical powder often lead to non-uniform packing morphologies that result in packing density variations in the powder compact.

Uncontrolled agglomeration of powders due to attractive van der Waals forces can also give rise to in microstructural inhomogeneities. Differential stresses that develop as a result of non-uniform drying shrinkage are directly related to the rate at which the solvent can be removed, and thus highly dependent upon the distribution of porosity. Such stresses have been associated with a plastic-to-brittle transition in consolidated bodies, and can yield to crack propagation in the unfired body if not relieved. [9] [10] [11]

In addition, any fluctuations in packing density in the compact as it is prepared for the kiln are often amplified during the sintering process, yielding inhomogeneous densification. Some pores and other structural defects associated with density variations have been shown to play a detrimental role in the sintering process by growing and thus limiting end-point densities. Differential stresses arising from inhomogeneous densification have also been shown to result in the propagation of internal cracks, thus becoming the strength-controlling flaws. [12][13] [14]

It would therefore appear desirable to process a material in such a way that it is physically uniform with regard to the distribution of components and porosity, rather than using particle size distributions which will maximize the green density. The containment of a uniformly dispersed assembly of strongly interacting particles in suspension requires total control over interparticle forces. Monodisperse nanoparticles and colloids provide this potential. [15]

Monodisperse powders of colloidal silica, for example, may therefore be stabilized sufficiently to ensure a high degree of order in the colloidal crystal or polycrystalline colloidal solid which results from aggregation. The degree of order appears to be limited by the time and space allowed for longer-range correlations to be established. Such defective polycrystalline colloidal structures would appear to be the basic elements of submicrometer colloidal materials science, and, therefore, provide the first step in developing a more rigorous understanding of the mechanisms involved in microstructural evolution in high performance materials and components. [16] [17]

Properties

Silicon nanopowder

Nanoparticles are of great scientific interest as they are effectively a bridge between bulk materials and atomic or molecular structures. A bulk material should have constant physical properties regardless of its size, but at the nano-scale size-dependent properties are often observed. Thus, the properties of materials change as their size approaches the nanoscale and as the percentage of atoms at the surface of a material becomes significant. For bulk materials larger than one micrometer (or micron), the percentage of atoms at the surface is insignificant in relation to the number of atoms in the bulk of the material. The interesting and sometimes unexpected properties of nanoparticles are therefore largely due to the large surface area of the material, which dominates the contributions made by the small bulk of the material.

For example, nanoparticles of usually yellow gold and gray silicon are red in color; gold nanoparticles melt at much lower temperatures (~300 °C for 2.5 nm size) than the gold slabs (1064 °C);[18] and absorption of solar radiation in photovoltaic cells is much higher in materials composed of nanoparticles than it is in thin films of continuous sheets of material – the smaller the particles, the greater the solar absorption.

Other size-dependent property changes include quantum confinement in semiconductor particles, surface plasmon resonance in some metal particles and superparamagnetism in magnetic materials. Ironically, the changes in physical properties are not always desirable. Ferromagnetic materials smaller than 10 nm can switch their magnetisation direction using room temperature thermal energy, thus making them unsuitable for memory storage.[19]

Suspensions of nanoparticles are possible since the interaction of the particle surface with the solvent is strong enough to overcome density differences, which otherwise usually result in a material either sinking or floating in a liquid. Nanoparticles also often possess unexpected optical properties as they are small enough to confine their electrons and produce quantum effects. For example gold nanoparticles appear deep red to black in solution.

The high surface area to volume ratio of nanoparticles provides a tremendous driving force for diffusion, especially at elevated temperatures. Sintering can take place at lower temperatures, over shorter time scales than for larger particles. This theoretically does not affect the density of the final product, though flow difficulties and the tendency of nanoparticles to agglomerate complicates matters. Moreover, nanoparticles have been found to impart some extra properties to various day to day products. For example the presence of titanium dioxide nanoparticles imparts what we call the self-cleaning effect, and the size being nanorange, the particles can not be observed. Zinc oxide particles have been found to have superior UV blocking properties compared to its bulk substitute. This is one of the reasons why it is often used in the preparation of sunscreen lotions.[20], and is completely photostable.[21]

Clay nanoparticles when incorporated into polymer matrices increase reinforcement, leading to stronger plastics, verifiable by a higher glass transition temperature and other mechanical property tests. These nanoparticles are hard, and impart their properties to the polymer (plastic). Nanoparticles have also been attached to textile fibers in order to create smart and functional clothing.[22]

Metal, dielectric, and semiconductor nanoparticles have been formed, as well as hybrid structures (e.g., core-shell nanoparticles). Nanoparticles made of semiconducting material may also be labeled quantum dots if they are small enough (typically sub 10 nm) that quantization of electronic energy levels occurs. Such nanoscale particles are used in biomedical applications as drug carriers or imaging agents.

Semi-solid and soft nanoparticles have been manufactured. A prototype nanoparticle of semi-solid nature is the liposome. Various types of liposome nanoparticles are currently used clinically as delivery systems for anticancer drugs and vaccines.

Nanoparticles with one half hydrophilic and the other half hydrophobic are termed Janus particles and are particularly effective for stabilizing emulsions. They can self-assemble at water/oil interfaces and act as solid surfactants.

Synthesis

There are several methods for creating nanoparticles, including both attrition and pyrolysis. In attrition, macro or micro scale particles are ground in a ball mill, a planetary ball mill, or other size reducing mechanism. The resulting particles are air classified to recover nanoparticles. In pyrolysis, a vaporous precursor (liquid or gas) is forced through an orifice at high pressure and burned. The resulting solid (a version of soot) is air classified to recover oxide particles from by-product gases. Pyrolysis often results in aggregates and agglomerates rather than single primary particles.

A thermal plasma can also deliver the energy necessary to cause evaporation of small micrometer size particles. The thermal plasma temperatures are in the order of 10,000 K, so that solid powder easily evaporates. Nanoparticles are formed upon cooling while exiting the plasma region. The main types of the thermal plasma torches used to produce nanoparticles are dc plasma jet, dc arc plasma and radio frequency (RF) induction plasmas. In the arc plasma reactors, the energy necessary for evaporation and reaction is provided by an electric arc which is formed between the anode and the cathode. For example, silica sand can be vaporized with an arc plasma at atmospheric pressure. The resulting mixture of plasma gas and silica vapour can be rapidly cooled by quenching with oxygen, thus ensuring the quality of the fumed silica produced. In RF induction plasma torches, energy coupling to the plasma is accomplished through the electromagnetic field generated by the induction coil. The plasma gas does not come in contact with electrodes, thus eliminating possible sources of contamination and allowing the operation of such plasma torches with a wide range of gases including inert, reducing, oxidizing and other corrosive atmospheres.

The working frequency is typically between 200 kHz and 40 MHz. Laboratory units run at power levels in the order of 30–50 kW while the large scale industrial units have been tested at power levels up to 1 MW. As the residence time of the injected feed droplets in the plasma is very short it is important that the droplet sizes are small enough in order to obtain complete evaporation. The RF plasma method has been used to synthesize different nanoparticle materials, for example synthesis of various ceramic nanoparticles such as oxides, carbours/carbides and nitrides of Ti and Si (see Induction plasma technology).

Inert-gas condensation is frequently used to make nanoparticles from metals with low melting points. The metal is vaporized in a vacuum chamber and then supercooled with an inert gas stream. The supercooled metal vapor condenses into nanometer-sized particles, which can be entrained in the inert gas stream and deposited on a substrate or studied in situ.

Sol-gel

File:Coll 1.jpg
SEM micrograph of amorphous colloidal silica particles (average particle diameter 600 nm) formed in basic solution from TEOS.

The sol-gel process is a wet-chemical technique (also known as chemical solution deposition) widely used recently in the fields of materials science and ceramic engineering. Such methods are used primarily for the fabrication of materials (typically a metal oxide) starting from a chemical solution (sol, short for solution) which acts as the precursor for an integrated network (or gel) of either discrete particles or network polymers. [23]

Typical precursors are metal alkoxides and metal chlorides, which undergo hydrolysis and polycondensation reactions to form either a network "elastic solid" or a colloidal suspension (or dispersion) – a system composed of discrete (often amorphous) submicrometer particles dispersed to various degrees in a host fluid. Formation of a metal oxide involves connecting the metal centers with oxo (M-O-M) or hydroxo (M-OH-M) bridges, therefore generating metal-oxo or metal-hydroxo polymers in solution. Thus, the sol evolves towards the formation of a gel-like diphasic system containing both a liquid phase and solid phase whose morphologies range from discrete particles to continuous polymer networks.[24]

In the case of the colloid, the volume fraction of particles (or particle density) may be so low that a significant amount of fluid may need to be removed initially for the gel-like properties to be recognized. This can be accomplished in any number of ways. The most simple method is to allow time for sedimentation to occur, and then pour off the remaining liquid. Centrifugation can also be used to accelerate the process of phase separation.

Removal of the remaining liquid (solvent) phase requires a drying process, which is typically accompanied by a significant amount of shrinkage and densification. The rate at which the solvent can be removed is ultimately determined by the distribution of porosity in the gel. The ultimate microstructure of the final component will clearly be strongly influenced by changes implemented during this phase of processing. Afterwards, a thermal treatment, or firing process, is often necessary in order to favor further polycondensation and enhance mechanical properties and structural stability via final sintering, densification and grain growth. One of the distinct advantages of using this methodology as opposed to the more traditional processing techniques is that densification is often achieved at a much lower temperature.

The precursor sol can be either deposited on a substrate to form a film (e.g. by dip-coating or spin-coating), cast into a suitable container with the desired shape (e.g. to obtain a monolithic ceramics, glasses, fibers, membranes, aerogels), or used to synthesize powders (e.g. microspheres, nanospheres). The sol-gel approach is a cheap and low-temperature technique that allows for the fine control of the product’s chemical composition. Even small quantities of dopants, such as organic dyes and rare earth metals, can be introduced in the sol and end up uniformly dispersed in the final product. It can be used in ceramics processing and manufacturing as an investment casting material, or as a means of producing very thin films of metal oxides for various purposes. Sol-gel derived materials have diverse applications in optics, electronics, energy, space, (bio)sensors, medicine (e.g. controlled drug release) and separation (e.g. chromatography) technology.[25][26]

Colloids

Nanostars of vanadium(IV) oxide

The term colloid is used primarily to describe a broad range of solid-liquid (and/or liquid-liquid) mixtures, all of which contain distinct solid (and/or liquid) particles which are dispersed to various degrees in a liquid medium. The term is specific to the size of the individual particles, which are larger than atomic dimensions but small enough to exhibit Brownian motion. If the particles are large enough, then their dynamic behavior in any given period of time in suspension would be governed by forces of gravity and sedimentation. But if they are small enough to be colloids, then their irregular motion in suspension can be attributed to the collective bombardment of a myriad of thermally agitated molecules in the liquid suspending medium, as described originally by Albert Einstein in his dissertation. Einstein proved the existence of water molecules by concluding that this erratic particle behavior could adequately be described using the theory of Brownian motion, with sedimentation being a possible long-term result. This critical size range (or particle diameter) typically ranges from nanometers (10−9 m) to micrometers (10−6 m).[27]

Morphology

Scientists have taken to naming their particles after the real world shapes that they might represent. Nanospheres[28], nanoreefs [29], nanoboxes [30] and more have appeared in the literature. These morphologies sometimes arise spontaneously as an effect of a templating or directing agent present in the synthesis such as miscellar emulsions or anodized alumina pores, or from the innate crystallographic growth patterns of the materials themselves.[31] Some of these morphologies may serve a purpose, such as long carbon nanotubes being used to bridge an electrical junction, or just a scientific curiosity like the stars shown at right.

Amorphous particles usually adopt a spherical shape (due to their microstructural isotropy) – whereas the shape of anisotropic microcrystalline whiskers corresponds to their particular crystal habit. At the small end of the size range, nanoparticles are often referred to as clusters. Spheres, rods, fibers, and cups are just a few of the shapes that have been grown. The study of fine particles is called micromeritics.

Characterization

File:TEM of 15nm Fe3O4 magnetic nano particles2.JPG
TEM image of magnetic Fe3O4 nanoparticle.[32]

Nanoparticle characterization is necessary to establish understanding and control of nanoparticle synthesis and applications. Characterization is done by using a variety of different techniques, mainly drawn from materials science. Common techniques are electron microscopy (TEM, SEM), atomic force microscopy (AFM), dynamic light scattering (DLS), x-ray photoelectron spectroscopy (XPS), powder X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF), ultraviolet-visible spectroscopy, dual polarisation interferometry and nuclear magnetic resonance (NMR).

Whilst the theory has been known for over a century (see Robert Brown), the technology for Nanoparticle tracking analysis (NTA) allows direct tracking of the Brownian motion and this method therefore allows the sizing of individual nanoparticles in solution.

Functionalization

The surface coating of nanoparticles is crucial to determining their properties. In particular, the surface coating can regulate stability, solubility and targeting. A coating that is multivalent or polymeric confers high stability. For biological applications, the surface coating should be polar to give high aqueous solubility and prevent nanoparticle aggregation. In serum or on the cell surface, highly charged coatings promote non-specific binding, while polyethylene glycol linked to terminal hydroxyl or methoxy groups repel non-specific interactions.[33] [34] Nanoparticles can be linked to biological molecules which can act as address tags, to direct the nanoparticles to specific sites within the body,[35] specific organelles within the cell,[36] or to follow specifically the movement of individual protein or RNA molecules in living cells.[37] Common address tags are monoclonal antibodies, aptamers, streptavidin or peptides. These targeting agents should ideally be covalently linked to the nanoparticle and should be present in a controlled number per nanoparticle. Multivalent nanoparticles, bearing multiple targeting groups, can cluster receptors, which can activate cellular signaling pathways, and give stronger anchoring. Monovalent nanoparticles, bearing a single binding site,[38] [39] [40] avoid clustering and so are preferable for tracking the behavior of individual proteins.

Safety

Nanoparticles present possible dangers, both medically and environmentally.[41] Most of these are due to the high surface to volume ratio, which can make the particles very reactive or catalytic.[42] They are also able to pass through cell membranes in organisms, and their interactions with biological systems are relatively unknown.[43] However, free nanoparticles in the environment quickly tend to agglomerate and thus leave the nano-regime, and nature itself presents many nanoparticles to which organisms on earth may have evolved immunity (such as salt particulates from ocean aerosols, terpenes from plants, or dust from volcanic eruptions)[citation needed]. A more complete analysis is provided in the article on nanotechnology.

According to the San Francisco Chronicle, "Animal studies have shown that some nanoparticles can penetrate cells and tissues, move through the body and brain and cause biochemical damage they also have shown to cause a risk factor in men for testicular cancer. But whether cosmetics and sunscreens containing nanomaterials pose health risks remains largely unknown, pending completion of long-range studies recently begun by the FDA and other agencies."[44] Diesel nanoparticles have been found to damage the cardiovascular system in a mouse model.[45]

See also

References

  1. ^ Cristina Buzea, Ivan Pacheco, and Kevin Robbie (2007). "Nanomaterials and Nanoparticles: Sources and Toxicity". Biointerphases. 2 (4): MR17–MR71. doi:10.1116/1.2815690. PMID 20419892.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  2. ^ ASTM E 2456 – 06 Standard Terminology Relating to Nanotechnology
  3. ^ Fahlman, B. D. (2007). Materials Chemistry. Springer. pp. 282–283. ISBN 1402061196.
  4. ^ A.B.D. Nandiyanto; S.-G Kim; F. Iskandar; and K. Okuyama 2009 447–453
  5. ^ Philip S. Rawson (1984). Ceramics. University of Pennsylvania Press. ISBN 0812211561.
  6. ^ Faraday, Michael (1857). "Experimental relations of gold (and other metals) to light". Phil. Trans. Roy. Soc. London. 147: 145–181. doi:10.1098/rstl.1857.0011.
  7. ^ Beilby, G.T. (1903). "The Effects of Heat and of Solvents on Thin Films of Metal". Proc. Roy. Soc. A. 72: 226.
  8. ^ Turner, T. (1908). "Transparent Silver and Other Metallic Films". Proc. Roy. Soc. Lond. A (Transparent silver and other metallic films). 81 (548): 301.
  9. ^ Edited by George Y. Onoda, Jr., and Larry L. Hench (1979). Onoda, G.Y., Jr. and Hench, L.L. Eds (ed.). Ceramic Processing Before Firing. New York: Wiley & Sons. ISBN 0471654108. {{cite book}}: |author= has generic name (help)CS1 maint: multiple names: authors list (link)
  10. ^ Aksay, I.A., Lange, F.F., Davis, B.I. (1983). "Uniformity of Al2O3-ZrO2 Composites by Colloidal Filtration". J. Am. Ceram. Soc. 66: C-190. doi:10.1111/j.1151-2916.1983.tb10550.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  11. ^ Franks, G.V. and Lange, F.F. (1996). "Plastic-to-Brittle Transition of Saturated, Alumina Powder Compacts". J. Am. Ceram. Soc. 79: 3161. doi:10.1111/j.1151-2916.1996.tb08091.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  12. ^ Evans, A.G. and Davidge, R.W. (1969). "The strength and fracture of fully dense polycrystalline magnesium oxide". Phil. Mag. 20 (164): 373. doi:10.1080/14786436908228708.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  13. ^ J Mat. Sci. 5: 314. 1970. {{cite journal}}: Missing or empty |title= (help)
  14. ^ Lange, F.F. and Metcalf, M. (1983). "Processing-Related Fracture Origins: II, Agglomerate Motion and Cracklike Internal Surfaces Caused by Differential Sintering". J. Am. Ceram. Soc. 66: 398. doi:10.1111/j.1151-2916.1983.tb10069.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ Evans, A.G. (1987). "Considerations of Inhomogeneity Effects in Sintering". J. Am. Ceram. Soc. 65: 497. doi:10.1111/j.1151-2916.1982.tb10340.x.
  16. ^ Whitesides, G.M.; et al. (1991). "Molecular Self-Assembly and Nanochemistry: A Chemical Strategy for the Synthesis of Nanostructures". Science. 254: 1312. doi:10.1126/science.1962191. {{cite journal}}: Explicit use of et al. in: |author= (help)
  17. ^ Dabbs D. M, Aksay I.A. (2000). "Self-Assembled Ceramics". Ann. Rev. Phys. Chem. 51: 601. doi:10.1146/annurev.physchem.51.1.601. PMID 11031294.
  18. ^ Buffat, Ph.; Burrel, J.-P. (1976). "Size effect on the melting temperature of gold particles". Physical Review A. 13 (6): 2287. doi:10.1103/PhysRevA.13.2287.
  19. ^ Sergey P. Gubin (2009). Magnetic nanoparticles. Wiley-VCH. ISBN 3527407901.
  20. ^ "Sunscreen". U.S. Food and Drug Administration.
  21. ^ Mitchnick, MA; Fairhurst, D; Pinnell, SR (1999). "Microfine zinc oxide (Z-cote) as a photostable UVA/UVB sunblock agent". Journal of the American Academy of Dermatology. 40 (1): 85–90. doi:10.1016/S0190-9622(99)70532-3. PMID 9922017.
  22. ^ "The Textiles Nanotechnology Laboratory".
  23. ^ Brinker, C.J. (1990). Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing. Academic Press. ISBN 0121349705. {{cite book}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  24. ^ L.L.Hench, J.K.West; West, Jon K. (1990). "The Sol-Gel Process". Chem. Rev. 90: 33–72. doi:10.1021/cr00099a003.
  25. ^ Klein, L. (1994). Sol-Gel Optics: Processing and Applications. Springer Verlag. ISBN 0792394240.
  26. ^ Robert Corriu, Nguyên Trong Anh (2009). Molecular Chemistry of Sol-Gel Derived Nanomaterials. John Wiley and Sons. ISBN 0470721170.
  27. ^ Pais, A. (2005). Subtle is the Lord: The Science and the Life of Albert Einstein. Oxford University Press. ISBN 0192806726.
  28. ^ Agam, M. A.; Guo, Q (2007). "Electron Beam Modification of Polymer Nanospheres". Journal of Nanoscience and Nanotechnology. 7 (10): 3615. doi:10.1166/jnn.2007.814. PMID 18330181.
  29. ^ Choy J.H., Jang E.S., Won J.H., Chung J.H., Jang D.J., and Kim Y.W. (2004). "Hydrothermal route to ZnO nanocoral reefs and nanofibers". Appl. Phys. Lett. 84: 287. doi:10.1063/1.1639514.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  30. ^ Sun, Y; Xia, Y (2002). "Shape-controlled synthesis of gold and silver nanoparticles". Science. 298 (5601): 2176–9. doi:10.1126/science.1077229. PMID 12481134.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  31. ^ Murphy, Cj (2002). "Materials science. Nanocubes and nanoboxes". Science. 298 (5601): 2139–41. doi:10.1126/science.1080007. PMID 12481122.
  32. ^ N. Pinna; et al. (2005). "Magnetite Nanocrystals: Nonaqueous Synthesis, Characterization, and Solubility". Chem. Mater. 17: 3044. doi:10.1021/cm050060. {{cite journal}}: Explicit use of et al. in: |author= (help)
  33. ^ Prime, KL; Whitesides, GM (1991). "Self-assembled organic monolayers: model systems for studying adsorption of proteins at surfaces". Science (New York, N.Y.). 252 (5010): 1164–7. PMID 2031186.
  34. ^ Liu, W; Greytak, AB; Lee, J; Wong, CR; Park, J; Marshall, LF; Jiang, W; Curtin, PN; Ting, AY (2010). "Compact biocompatible quantum dots via RAFT-mediated synthesis of imidazole-based random copolymer ligand". Journal of the American Chemical Society. 132 (2): 472–83. doi:10.1021/ja908137d. PMC 2871316. PMID 20025223.
  35. ^ Akerman ME, Chan WC, Laakkonen P, Bhatia SN, Ruoslahti E. (2002). "Nanocrystal targeting in vivo". Proceedings of the National Academy of Sciences of the United States of America. 99 (20): 12617–21. doi:10.1073/pnas.152463399. PMC 130509. PMID 12235356.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  36. ^ Hoshino, A; Fujioka, K; Oku, T; Nakamura, S; Suga, M; Yamaguchi, Y; Suzuki, K; Yasuhara, M; Yamamoto, K (2004). "Quantum dots targeted to the assigned organelle in living cells". Microbiology and immunology. 48 (12): 985–94. PMID 15611617.
  37. ^ Suzuki, KG; Fujiwara, TK; Edidin, M; Kusumi, A (2007). "Dynamic recruitment of phospholipase C gamma at transiently immobilized GPI-anchored receptor clusters induces IP3-Ca2+ signaling: single-molecule tracking study". The Journal of cell biology. 177 (4): 731–42. doi:10.1083/jcb.200609175. PMC 2064217. PMID 17517965.
  38. ^ Sung, KM; Mosley, DW; Peelle, BR; Zhang, S; Jacobson, JM (2004). "Synthesis of monofunctionalized gold nanoparticles by fmoc solid-phase reactions". Journal of the American Chemical Society. 126 (16): 5064–5. doi:10.1021/ja049578p. PMID 15099078.
  39. ^ Fu, A; Micheel, CM; Cha, J; Chang, H; Yang, H; Alivisatos, AP (2004). "Discrete nanostructures of quantum dots/Au with DNA". Journal of the American Chemical Society. 126 (35): 10832–3. doi:10.1021/ja046747x. PMID 15339154.
  40. ^ Howarth, M; Liu, W; Puthenveetil, S; Zheng, Y; Marshall, LF; Schmidt, MM; Wittrup, KD; Bawendi, MG; Ting, AY (2008). "Monovalent, reduced-size quantum dots for imaging receptors on living cells". Nature methods. 5 (5): 397–9. doi:10.1038/nmeth.1206. PMC 2637151. PMID 18425138.
  41. ^ Mnyusiwalla, Anisa; Daar, Abdallah S; Singer, Peter A (2003). "Mind the gap : science and ethics in nanotechnology". Nanotechnology. 14: R9. doi:10.1088/0957-4484/14/3/201.
  42. ^ Ying, Jackie (2001). Nanostructured Materials. New York: Academic Press. ISBN 0127444513.
  43. ^ Nanotechnologies: 6. What are potential harmful effects of nanoparticles?
  44. ^ Keay Davidson. "FDA urged to limit nanoparticle use in cosmetics and sunscreens". San Francisco Chronicle. Retrieved 2007-04-20.
  45. ^ Study Pollution Particles Lead to Higher Heart Attack Risk (Update1)