Anda di halaman 1dari 7

13670

J. Phys. Chem. C 2009, 113, 1367013676

Biomineralization of Fine Selenium Crystalline Rods and Amorphous Spheres


Gurinder Kaur, Mohammad Iqbal, and Mandeep Singh Bakshi*,
Nanotechnology Research Laboratory, College of North Atlantic, Labrador City, A2 V 2Y1 Newfoundland, Canada,
College of North Atlantic, Prince Philip DriVe Campus, St. Johns, A1C 5P7 Newfoundland, Canada, and
Department of Chemistry, Acadia UniVersity, Elliot Hall, WolfVille, B4P 2R6, NoVa Scotia Canada
ReceiVed: April 21, 2009; ReVised Manuscript ReceiVed: June 12, 2009

A simple aqueous phase method containing a water-soluble carrier protein, bovine serum albumin (BSA),
has been presented for the synthesis of well-defined morphologies of nanobiomaterials. BSA has been used
as a shape-directing agent to synthesize crystalline Se nanobars (NBs) and amorphous nanospheres in aqueous
phase at a relatively low temperature of 85 C. Na2SeO3 is used as the Se source to achieve nanoselenium
following hydrazine reduction. Well-defined multifacet NBs are produced when the amount of Na2SeO3 is at
least 6 times greater than that of BSA (on the basis of per residue), while amorphous spheres are formed with
nearly a 1:1 ratio. Both morphologies have been fully characterized by field emission scanning electron
microscopy (FESEM), high-resolution transmission electron microscopy (HRTEM), energy dispersive X-ray
spectroscopy (EDX), X-ray diffraction (XRD), and X-ray photoelectron spectroscopic (XPS) analysis. Results
have shown that the shape-directing ability of unfolded BSA helped to achieve the formation of crystalline
NBs, while its soft template effect directed the nanosphere formation.
1. Introduction
Bionanomaterials are highly important constituents of biocompatible devices with many applications in bioengineering,
biomedical imaging, molecular diagnostics, and most importantly a new class of hybrid materials.1 Material properties affect
biological outcomes including the half-life of drugs, biocompatibility of implanted devices, and release rates and toxicity
of drug carriers.1g Similarly, physical and chemical properties
of biomaterials can have a profound impact on cell proliferation
and remodeling of tissues.1f A precise shape-controlled synthesis
of a biomaterial is possible only if capping biomolecules could
selectively control the crystal growth. Anionic phospholipids
(PLs) have been found to be excellent capping/stabilizing agents
for gold nanoparticles (Au NPs).2 Surprisingly, their zwitterionic
homologues (phosphocholines) showed the least shape controlled effects.2b Fine PL-capped Au NPs were then used as
model air pollutants to study their effect on the surface activity
of semisynthetic pulmonary surfactants.3 More recently, bovine
serum albumen (BSA), a water-soluble and highly important
carrier protein, showed remarkable shape-controlled effects on
PbS nanocrystals with respect to a temperature variation within
40-80 C.4 The unfolded form of BSA worked effectively in
controlling the crystal structure and led to well-defined cubic
nanomorphologies in comparison to its native folded state. The
exposed hydrophobic domains of unfolded form provided
desired surface activity to control the crystal growth. Use of a
carrier protein like BSA in a shape-controlled synthesis of
bionanomaterials provides a direct opportunity to produce
desired biomaterials for devices with applications in bioengineering. Although BSA has been used as a capping/stabilizing
agent for different materials,5 precise shape-controlled morphologies are still elusive. We herein report the synthesis of
fine crystalline nanobars (NBs) and amorphous spheres of
* Corresponding author. E-mail: ms_bakshi@yahoo.com.

College of North Atlantic, Labrador City.

College of North Atlantic, St. Johns.

Acadia University.

selenium (Se) under different experimental conditions using


BSA as a shape-directing agent.
Selenium is an important inorganic semiconducting material
with a large Bohr radius. In addition to its interesting physical
properties such as thermoelectric and nonlinear optical responses, high conductivity, and piezoelectric effects,6 Se in
appropriate amounts is an essential element for living organisms.7 Se has been shown to prevent cancer in numerous animal
model systems when fed at levels exceeding the nutritional
requirement.8 Clarke et al. showed cancer chemopreventive
efficacy using a Se supplement in humans.9 Protein-selenium
bioconjugate nanomaterials are reported to be cytotoxic for
tumor cells.10 Se NPs have also been studied for their antioxidant
activity.11a Apart from its many useful applications, excess of
Se intake causes selenosis in animals and humans.11b Use of
BSA in synthesizing the shape-controlled BSA-Se bioconjugate
semiconducting nanomaterials are expected to have several
advantages oversimple Se nanomaterials as far as their biomedical applications are concerned.11c
2. Experimental Section
2.1. Synthesis of Se NPs. Sodium selenite, hydrazine, and
BSA, all 99% pure, were purchased from Aldrich. Double
distilled water was used for all preparations.
Se nanocrystals were synthesized by a chemical reduction
of sodium selenite by hydrazine. In a typical procedure, 10 mL
of aqueous BSA (1-10 10-4 g/mL) along with 2.5%
hydrazine was taken in a round-bottom glass flask. Under
constant stirring, sodium selenite (6-25 mM) was added in it.
After mixing all the components at room temperature, the
reaction mixture was kept in a water thermostat bath (Julabo F
25) at 85 C for 48 h. The color of the solution changed from
colorless to deep orange within 2 h and remained the same for
48 h. Initial pH of the BSA+water solution was 6.7, which
increased to 7.8 upon addition of hydrazine. Addition of
Na2SeO3 further increased the pH to 9 as a result of the
following reaction:

10.1021/jp903685g CCC: $40.75 2009 American Chemical Society


Published on Web 07/01/2009

Biomineralization of Se NBs and Amorphous Spheres

J. Phys. Chem. C, Vol. 113, No. 31, 2009 13671

N2H4(aq) + SeO32-(aq) f Se(s) + N2(g) + 2OH-(aq) + H2O

(1)
Within 2 h, pH further rose to 11 due to the denaturation of
BSA (at 85 C) and remained fairly constant for 48 h. The N
(normal) to B (basic) transitions, which take place around pH
) 7-9, affect Cys-Cys as well as C-S bonds and cause
unfolding at alkaline pH. The samples were purified by spinning
the reaction product at 10 000 rpm for 10 min with repeated
washing with distilled water.
2.2. Methods. Field Emission Scanning Electron Microscopy (FESEM), Transmission Electron Microscopy (TEM),
X-ray Diffraction (XRD), and X-ray Photoelectron Spectroscopy (XPS) Measurements. FESEM analysis was carried out
on a Zeiss NVision 40 Dual Beam FIB/SEM instrument. TEM

analysis was done on a JEOL 2010F at an operating voltage of


200 kV. Photomicrographs were obtained in bright field
scanning/imaging mode, using a spot size of 1 nm and a
camera length of 12 cm. Energy dispersive X-ray (EDX)
microanalysis was carried out using an Oxford-INCA Atmospheric Ultrathin Window (UTW), and the data was processed
using the Oxford INCA Microanalysis Suite, version 4.04. XRD
patterns were recorded by using Bruker-AXS D8-GADDS with
Tsec ) 480. Samples were prepared on glass slides by putting a
concentrated drop of aqueous sample and then dried in vacuum
desiccator. The chemical composition was confirmed with the
help of XPS measurements. A portion of an aqueous NP solution
was placed onto a clean silicon wafer and then it was put into
the introduction chamber of the XPS instrument. The liquid was

Figure 1. (a) Low-resolution FESEM image of several bundles of NBs. (b) Magnified image of several multifaceted NBs. (c) Close-up image
showing surface-adsorbed BSA and the presence of BSA in between the NBs. Inset, dark-field image showing the BSA coating. (d) HAADF image
showing bright patches of adsorbed BSA and a few dark patches were created by the exposure to electron beam. (e) TEM images of a single NB
with a selected area electron diffraction (SAED) image (inset) and (f) its EDX spectrum. (g) Line EDX line spectrum across two fused NBs
showing emission due to Se.

13672

J. Phys. Chem. C, Vol. 113, No. 31, 2009

Kaur et al.

Figure 2. (a) Dark-field TEM image of a single bar showing the area in the box used for HRTEM analysis (b). (c) The same NB sensitive to
electron beam splits into two pieces and the boxed area used for HRTEM analysis (d). (e) XRD patterns showing only one prominent peak due to
predominant growth at {100} planes.

then pumped away. The XPS analyses were carried out with a
Kratos Axis Ultra spectrometer using a monochromatic Al KR
source (15 mA, 14 kV). Survey and high-resolution analyses
were carried out with an analysis area of 300 700 m using
pass energies of 160 and 20 eV, respectively. Special care was
taken to completely remove the uncapped BSA before XPS
measurements.
2.3. Protein Assay. Bradford method12 was used to determine
the total protein contents in the BSA-NP conjugate suspension.
For this purpose, standard BSA (reference) solutions of concentrations 0, 2, 4, 6, 8, and 10 g/L were prepared in 100 L
distilled water. 10 L of each of these solutions was taken in
triplicate in different wells of the UV-plate. 1 mg of the dried
Se samples (purified and dried at 40 C) was taken in doublet
in the UV-plate. 20 L of pure water was added to the wells
containing the reference (BSA) and 30 L was added to the

wells with dried samples. After this, 170 L of the Bradford


reagent were mixed in all the wells to get total volume of 200
L. The absorbance of each solution was measured and from
the absorbance values the amount of BSA conjugated to Se NCs
in both samples was calculated.
3. Results and Discussion
Figure 1a shows the FESEM image of Se NBs synthesized
with [Na2SeO3] ) 6 mM in the presence of BSA ) 1 10-4
g/mL. A BSA macromolecule contains 607 residues with
average molecular weight of 66432 Da.13 Assuming all residues
of the same nature, each one will contribute about 109.4 mols,
and thus [Na2SeO3/BSA] mole ratio ) 6.7/residue. Bundles of
fine multifaceted bars are evident (Figure 1b) with an average
aspect ratio of 4.7 ( 1.8 (Supporting Information, Figure S1).
Figure 1c indicates (block arrows) the presence of conjugated

Biomineralization of Se NBs and Amorphous Spheres

J. Phys. Chem. C, Vol. 113, No. 31, 2009 13673

Figure 3. Low- (a) and high-resolution XPS spectra of (b) Se 3d, (c) C 1s, and (d) N 1s (see details in text).

BSA (BSAc) on the surface of NBs. A close up bright-field


TEM image (inset) further confirms this. It appears that most
of the surface of each NB is covered with BSAc. A high angle
annular dark field (HAADF) image (Figure 1d) clearly shows
bright surface patches on each NB indicated by block arrows.
A dark dotted circle encloses a few spots of BSA-covered areas
of the sample that were damaged by the electron beam. A TEM
image of a single bar is shown in Figure 1e, which is positioned
along the [110] zone as evident from the diffraction image
(inset). EDX analysis of this particle further confirms the
presence of Se (Figure 1f). A few weak Cu emissions are due to
the Cu grid. Se EDX line spectrum is performed across the short
axes of two NBs (Figure 1g). A simultaneous emission even
from the attached portions of both NBs indicates a slight degree
of fusion facilitated by the BSAc.
Se has a trigonal arrangement due to spiral chains of Se atoms
associated with each other through van der Waals interactions
in a hexagonal lattice.14 Such an arrangement provides a
unidirectional growth tendency. A thermodynamically stable
trigonal (t) Se is expected to favor growth along the 100
direction and eventually lead to the formation of onedimensional (1D) nanostructures such as NBs. Figure 2 demonstrates the high-resolution transmission electron microscopy
(HRTEM) characterization of a single NB. A lattice resolved
image of a magnified part of a single NB (Figure 2a) is shown
in Figure 2b with lattice spacing of 0.38 nm, which refers to
{100} crystal planes of trigonal geometry. Interestingly, the
HAADF image is highly sensitive to the electron beam, and it
breaks the NB into two pieces, as shown in Figure 2c. A lattice
resolved image of a broken part (see the scan area in Figure
2c) gives a lattice spacing of 0.30 nm, which corresponds to
{101} crystal planes. Powder pattern XRD (Figure 2e) gives

only one sharp peak with prominent growth along the 100
crystal planes of trigonal hexagonal geometry of Se. The
effective capping ability of BSA helps to attain well-defined
rod shape geometries. BSA binds endogenous as well as
exogenous substrates in its hydrophobic pockets.15 The overall
shape of a BSA macromolecule is oblate ellipsoid in its native
state, but denaturation is often followed by a massive unfolding of the protein. The secondary structure consists of
hydrogen-bonded R-helices and -sheets, and is called the largescale structure. At 85 C, BSA is considered to be in its unfolded
form because the overall denaturation temperature is usually
reported to be close to 60 C depending on different methodologies and detection techniques.16 Although the exact mechanism
is still unclear, the unfolded BSA with predominantly hydrophobic domains might be adsorbed favorably on freshly cleaved
Se nucleating centers.10,11,17 Thus, a selective adsorption of the
unfolded form of BSA on low atomic density {100} crystal
planes will direct the crystal growth on {111} planes and will
result in the rod shape formation. The adsorption of BSA on
Se NBs is further confirmed from XPS studies. Figure 3a
presents a low-resolution spectrum of this sample, while highresolution spectra for Se 3d, C 1s, and N 1s are shown in Figure
3b, c, and d, respectively. Elemental selenium is generally
observed between 54.9 and 56.3 eV.18 In our case, Se 3d peak
exists in a weak doublet. Deconvolution of it gives two
prominent peaks of Se 3d5/2 at 55.8 and 55.0 eV. The later
peak refers to the elemental Se, while the former can be due to
oxidation. The XPS peak for C 1s at 284.8 eV refers to C-C
and C-H functional groups19 of BSA macromolecules. Similarly, N 1s produces a peak at 400.1 eV due to associated amine
groups.20 Finally, the amount of BSAc on NBs was determined
by following the Bradford method. In order to determine the

13674

J. Phys. Chem. C, Vol. 113, No. 31, 2009

Kaur et al.

Figure 4. (a) Low-resolution FESEM image of several groups of spheres. (b) Magnified image of a single group containing several spheres. (c)
Close-up image showing fused spheres. (d) HAADF image showing a pair of fused spheres along with (e) a line spectrum across them indicating
the emission due to Se. (f) TEM image of a single sphere with SAED image (inset) and (g) its EDX spectrum.

mode of association of BSA with growing NBs, samples at


regular intervals were drawn from the reaction mixture. The
particles of each sample were thoroughly washed with pure
water to remove unassociated BSA. Then each sample was
carefully dried and used to estimate BSAc (see Experimental
Section). The amount of BSAc increases as the reaction proceeds
(Figure S2) and then becomes constant within 18 h. It means
that NBs continuously grow for 18 h before attaining a limiting
growth. Se 1D nanostructures usually take 24 h to grow into
well-defined geometries21a which is due to the fact that the least
stable allotropes precipitate first, and then slowly develop into
a thermodynamically more stable phase.21
Increase in the amount of BSA from 1 to 10 10-4 g/mL
does not help to further improve the morphology of NBs; instead
the shape of the bars becomes deformed (see Figure S3). But a
[Na2SeO3/BSA] mole ratio of 2.6 per residue, produced large

spheres along with fine long needles (see Figure S4). We were
interested in the nature of large spheres. Interestingly, further
decrease in [Na2SeO3/BSA] mole ratio to 1.3 per residue
suddenly eliminates most of the NBs or needles leaving behind
only large groups of spheres. A low magnification FESEM
image of such several groups of spheres is shown in Figure 4a.
Size distribution histogram computes an average size of a sphere
equal to 346 ( 110 nm (Figure S5). A high-magnification
image (Figure 4b) shows several interconnected spheres in a
single group. The spheres are in fact fused together sidewise
(Figure 4c). In order to further evaluate the mode of fusion, we
carefully selected a pair of spheres and got the HAADF image.
This image (Figure 4d) fully confirms the fact that the spheres
are indeed fused with each other. An EDX line spectrum (Figure
4e) running across the two balls further confirms that the fused
part does not contain BSA because full Se emission is visible

Biomineralization of Se NBs and Amorphous Spheres

J. Phys. Chem. C, Vol. 113, No. 31, 2009 13675

Figure 5. (a,b) HRTEM images of a nanosphere surface to show the absence of crystalline morphology and BSA coating. (c) XRD patterns
showing no peaks indicating the amorphous nature.

from the total surface area. Figure 4f shows a TEM image of a


single ball. Surprisingly, the diffraction image (inset) of this
ball gives no diffraction rings, suggesting the amorphous nature
of spheres. EDX spectrum (Figure 4g) further proves that this
ball is entirely made up of Se. Hence, the interparticle fusion
of such spheres is predominantly caused by their amorphous
nature and not because of BSA capping. Figure 5a,b demonstrates two HRTEM images from different angles. No lattice
planes were observed in both cases, and there was no sign of
associated/capped BSA. Likewise, no diffraction peak were
observed in the XRD patterns (Figure 5c), further confirming
the presence of amorphous nature. Amorphous Se (a-Se) spheres
have already been reported by other groups22 and are believed
to be formed as a result of an insufficient amount of capping/
stabilizing agent.
The above results clearly indicate that the NB and sphere
formation is related to the [Na2SeO3/BSA] mole ratio. The fine
NBs shown in Figure 1 are obtained at a mole ratio of 6.7 per
residue, or in other words, when the amount of Na2SeO3 is much
higher than that of BSA. Large spheres (Figure 5) on the other
hand are synthesized with a ratio of 1.3 per residue. At an
intermediate ratio of 2.6, both rods and spheres are observed
(Figure S4). However, a large increase in the ratio even up to
28 does not help to achieve any more precise shape control;
instead long needles of several micrometers become the major
reaction product, as observed by other researchers.23 Thus, an
appropriate balance between the amounts of both Na2SeO3 and
BSA is essential to achieve shape-controlled morphologies. The
mechanism of the reaction is considered to follow the following
steps: In the first step, SeO3- ions adsorb electrostatically on
the surface of unfolded BSA. Hydrazine reduces SeO3- ions
into Se nucleating centers, which are simultaneously adsorbed

at the BSA-water interface.11,24 We do not see any significant


effect of hydrazine on BSA. UV-visible spectrum shows that
the intensity of the tryptophan absorbance around 278 nm
slightly increases in the presence of hydrazine (Figure S6),
which might be due to some conformational changes in view
of an increased pH. At high temperature (85 C), more SeO3ions will help to grow such nucleating centers by following an
autocatalytic process. When the [Na2SeO3/BSA] mole ratio is
6.7, there are more than 6 times SeO3- ions in comparison to
electropositive sites per residue (due to ammonium groups)
available on BSA. The actual number of electropositive sites
must be even far less than this because the unfolded state is
considered to be predominantly hydrophobic in nature. Thus,
an unfolded state not only helps in achieving the capping/
stabilizing effects but also acts as a soft template where an
excess of SeO3- ions will direct the crystal growth in the 100
direction. Such a mechanism would obviously lead to the NB
formation. On the contrary, when the [Na2SeO3/BSA] mole ratio
is 1.3, all SeO3- ions may find their way into small Se nucleating
centers adsorbed on the unfolded BSA with little probability of
further autocatalytic process. Intramolecular dynamics influenced
by the conformation changes and changing topographic effects24
can induce Ostwald repining among the nucleating centers and
may lead to the formation of large spheres of amorphous nature.
Protein adsorption characteristics are generally governed by the
surface topography, i.e., curvature,25 and albumen is usually less
ordered on large substrates.24 This effect should be more
pronounced on a curved surface of a sphere rather than on a
NB, hence it will reduce the capping ability of BSA to control
the crystal growth in an ordered manner.

13676

J. Phys. Chem. C, Vol. 113, No. 31, 2009

4. Conclusions
This study addresses a very significant and important problem
of shape-controlled synthesis of ordered bionanomaterials in the
presence of proteins. Ordered morphologies are highly important
to build biochip devices. The results conclude that BSA works
well in controlling the overall geometry of NPs when the
[Na2SeO3/BSA] mole ratio is 6. At this mole ratio, denatured
BSA becomes even more hydrophobic as a result of the
neutralization of oppositely charged sites by SeO3- ions. A
predominantly hydrophobic BSA is a better shape-directing
agent. However, at a too high mole ratio of 28, BSA proves to
be a poor capping agent because of the presence of too many
nucleating centers whose growth cannot be simultaneously
controlled by BSA macromolecules. On the other hand, when
the mole ratio is 1, then unfolded BSA macromolecule works
as a soft template by accommodating maximum nucleating
centers on it and thereby facilitating the Ostwald ripening.
Therefore, in order to observe the best shape directing effect of
BSA, the following features have to be taken into consideration:
(a) BSA should be in the unfolded and predominantly hydrophobic state. (b) The precursor concentration should be greater
than that of BSA so that the growing nucleating centers can be
properly stabilized. (c) Too many nucleating centers cannot be
simultaneously stabilized by BSA because of its time-dependent
surface adsorption.
Acknowledgment. The authors would like to extend sincere
thanks to Dr. Richard Sawyer at CNA, Lab West, for arranging
financial assistance for the work. We thankfully acknowledge
the help rendered by Julia Huang and Carmen Andrei at the
Canadian Centre for Electron Microscopy, McMaster University.
Supporting Information Available: Size distribution histograms, SEM images, other information is available free of
charge via the Internet at http://pubs.acs.org.
References and Notes
(1) (a) Taft, B. J.; Lazareck, A. D.; Withey, G. D.; Yin, A.; Xu, J. M.;
Kelley, S. O. J. Am. Chem. Soc. 2004, 126, 12750. (b) Li, W.; Gao, C.
Langmuir 2007, 23, 4575. (c) Gao, C.; Muthukrishnan, S.; Li, W.; Yuan,
J.; Xu, Y.; Muller, A. H. E. Macromolecules 2007, 40, 1803. (d)
Eggenberger, K.; Merkulov, A.; Darbandi, M.; Nann, T.; Nick, P. Bioconjugate Chem. 2007, 18, 1879. (e) Wang, Y.; Tang, Z.; Tan, S.; Kotov, N. A.
Nano Lett. 2005, 5, 243. (f) Langer, R.; Tirrel, D. A. Nature 2004, 428,
487. (g) Mooney, D. J.; Mikos, A. G. Sci. Am. 1995, 280, 60.
(2) (a) Bakshi, M. S.; Possmayer, F.; Petersen, N. O. Chem. Mater.
2007, 19, 1257. (b) Bakshi, M. S.; Possmayer, F.; Petersen, N. O. J. Phys.
Chem. C 2007, 111, 14113. (c) Bakshi, M. S.; Kaur, G.; Thakur, P.; Banipal,
T. S.; Possmayer, F.; Petersen, N. O. J. Phys. Chem. C 2007, 111, 5932.
(d) Meister, A.; Drescher, S.; Mey, I.; Wahab, M.; Graf, G.; Garamus, V. M.;
Hause, G.; Mogel, H. J.; Janshoff, A.; Dobner, B.; Blume, A. J. Phys. Chem.
B 2008, 112, 4506.
(3) (a) Bakshi, M. S.; Zhao, L.; Smith, R.; Possmayer, F.; Petersen,
N. O. Biophys. J. 2008, 94, 855. (b) Nakahara, H.; Lee, S.; Sugihara, G.;
Chang, C.-H.; Shibata, O. Langmuir 2008, 24, 3370.
(4) Bakshi, M. S.; Thakur, P.; Kaur, G.; Kaur, H.; Banipal, T. S.;
Possmayer, F.; Petersen, N. O. AdV. Funct. Mater. 2009, 19, 1451.

Kaur et al.
(5) (a) Xie, J.; Lee, J. Y.; Wang, D. I. C. J. Phys. Chem. C 2007, 111,
10226. (b) Shang, Li.; Wang, Y.; Jiang, J.; Dong, S. Langmuir 2007, 23,
2714. (c) Meziani, M. J.; Pathak, P.; Harruff, B. A.; Hurezeanu, R.; Sun,
Y.-P. Langmuir 2005, 21, 2008. (d) Brewer, S. H.; Glomm, W. R.; Johnson,
M. C.; Knag, M. K.; Franzen, S. Langmuir 2005, 21, 9303. (e) Mikhaylova,
M.; Kim, D. K.; Berry, C. C; Zagorodni, A.; Toprak, M.; Curtis, A. S. G.;
Muhammed, M. Chem. Mater. 2004, 16, 2344.
(6) Gao, X.; Zhang, J.; Zhang, L. AdV. Mater. 2002, 14, 290.
(7) (a) Rosenfield, I.; Beath, O. A. Selenium: Geotoxicity, Biochemistry,
Toxicity and Nutrition; Academic Press: New York, 1964. (b) Bock, A. In
Encyclopedia of Inorganic Chemistry; Bruce-King, R., Ed.; Wiley: New
York, 1994; Vol. 7, p 3700. (c) Mugesh, G.; du Mont, W.; Sies, H. Chem.
ReV. 2001, 101, 2125.
(8) Ip, C. Selenium inhibition of chemical carcenogenesis. Fed. Proc.
1984, 44, 2573.
(9) Clark, L. C.; Combs, G. F.; Turnbull, B. W. Effects of selenium
supplementation for cancer prevention in patients with carcinoma of skin;
A randomized controlled trial. J. Am. Med. Assoc. 1996, 276, 1957.
(10) Sieber, F.; Daziano, J.; Gunther, W. H.; Krieg, M.; Miyagi, K.;
Sampson, R. W.; Ostrowski, M. D.; Anderson, G. S.; Tsujino, I.; Bula,
R. J. Phosphorus, Sulfur Silicon Relat. Elem. 2005, 180, 647.
(11) (a) Zingaro, R. A., ; Cooper, W. C., Eds. Selenium; Litton
Educational Publishing: New York, 1974; pp 12-28. (b) Seiler, H. G.; Sigel,
A.; Sigel, H. Handbook on Metals in Clinical and Analytical Chemistry;
M. Dekker: New York, 1994. (c) Aguanno, J. J.; Ladenson, J. H. J. Biol.
Chem. 1982, 257, 8745.
(12) Bradford, M. M. Anal. Biochem. 1976, 72, 248.
(13) (a) Carter, D. C.; Ho, X. J. AdV. Protein Chem. 1994, 45, 153. (b)
Papedopoulou, A.; Green, R. J.; Frazier, R. A. J. Agric. Food. Chem. 2005,
53, 158. (c) Honda, C.; Kamizono, H.; Samejima, T.; Endo, K. Chem.
Pharm. Bull. 2000, 48, 464.
(14) Hippel, A. V. J. Chem. Phys. 1948, 16, 372.
(15) Peters, T., Jr. All about Albumin: Biochemistry, Genetics, and
Medical Applications; Academic Press: San Diego, 1996.
(16) (a) Murayama, K.; Tomida, M. Biochemistry 2004, 43, 11526. (b)
Baier, S.; McClements, D. J. J. Agric. Food Chem. 2001, 49, 2600. (c)
Moriyama, Y.; Takeda, K. Langmuir 2005, 21, 5524. (d) Militello, V.; Vetri,
V.; Leone, M. Biophys. Chem. 2003, 105, 133.
(17) Mishra, B; Hassan, P. A.; Priyadarsini, K. I.; Mohan, H. J. Phys.
Chem. B 2005, 109, 12718.
(18) Naveau, A; Monteil-Rivera, F.; Guillon, E.; Dumonceau, J. EnViron.
Sci. Technol. 2007, 41, 5376.
(19) Bakshi, M. S.; Thakur, P.; Sachar, S.; Kaur, G.; Banipal, T. S.;
Possmayer, F.; Petersen, N. O. J. Phys. Chem. C 2007, 111, 18087.
(20) (a) Sharma, J.; Mahima, S.; Kakade, B. A.; Pasricha, R.; Mandale,
A. B.; Vijayamohanan, K. J. Phys. Chem. B 2004, 108, 13280. (b) Sharma,
J.; Chaki, N. K.; Mandale, A. B.; Pasricha, R.; Vijayamohanan, K. J. Colloid
Interface Sci. 2004, 272, 145.
(21) (a) Li, X.; Li, Y.; Li, S.; Zhou, W.; Chu, H.; Chen, W.; Li, I. L.;
Tang, Z. Cryst. Growth Des. 2005, 5, 911. (b) Jolivet, J.-P.; Henry, M.;
Livage, J. Metal Oxide Chemistry and Synthesis - From Solution to Solid
State; Bescher, E., Translator; John Wiley and Sons, LTD: Chichester, U.K.,
2000; p 47.
(22) (a) Gates, B.; Mayers, B.; Cattle, B.; Xia, Y. AdV. Funct. Mater.
2002, 12, 219. (b) Mees, D. R.; Pysto, W.; Tarcha, P. J. J. Colloid Interface
Sci. 1995, 170, 254.
(23) (a) Zhang, B.; Dai, W.; Ye, X.; Zuo, F.; Xie, Y. Angew. Chem.,
Int. Ed. 2006, 45, 2571. (b) Lu, J.; Xie, Y.; Xu, F.; Zhu, L. J. Mater. Chem.
2002, 12, 2755. (c) Tang, K.; Yu, D.; Wang, F.; Wang, Z. Cryst. Growth
Des. 2006, 6, 2159. (d) Mondal, K.; Roy, P.; Srivastava, S. K. Cryst. Growth
Des. 2008, 8, 1580.
(24) Roach, P.; Farrar, D.; Perry, C. C. J. Am. Chem. Soc. 2006, 128,
3939.
(25) (a) Vertegal, A. A.; Siegel, R. W.; Dordick, J. S. Langmuir 2004,
20, 6800. (b) Lundqvist, M.; Sethson, I.; Johnson, B.-H. Langmuir 2004,
20, 10639.

JP903685G

Anda mungkin juga menyukai