Sunteți pe pagina 1din 7

Chemical Engineering Journal 254 (2014) 252–258

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Preparation of zinc oxide/silica nanocomposite particles via consecutive


sol–gel and flame-assisted spray-drying methods
W. Widiyastuti a,⇑, Iva Maula a, Tantular Nurtono a, Fadlilatul Taufany a, Siti Machmudah a,
Sugeng Winardi a, Camellia Panatarani b
a
Department of Chemical Engineering, Institut Teknologi Sepuluh Nopember, Kampus ITS Sukolilo, Surabaya 60111, Indonesia
b
Department of Physics, Universitas Padjadjaran, Jl. Raya Bandung-Sumedang km. 21, Sumedang 45363, Indonesia

h i g h l i g h t s

 Water glass was used as a silica source to prepare ZnO/SiO2 nanocomposite.


 The mixed ZnO–SiO2 colloids were dried by flame-assisted spray-drying methods.
 The addition of SiO2 to ZnO resulted in the increase of the particles photoluminescence intensity.
 Silica content of the nanocomposites is significantly affected the particles characteristics.

a r t i c l e i n f o a b s t r a c t

Article history: ZnO/SiO2 nanocomposites were prepared using consecutive sol–gel and flame-assisted spray-drying
Received 7 February 2014 methods. Zinc oxide and silica sols were prepared individually from zinc acetate and water glass, respec-
Received in revised form 19 May 2014 tively, via a sol–gel method. Subsequently, zinc oxide sol, silica sol, and a mixture of both were spray
Accepted 23 May 2014
dried in a flame reactor to generate powder of pure zinc oxide, pure silica, and a ZnO/SiO2 nanocomposite,
Available online 5 June 2014
respectively, in order to compare the characteristics. The generated particles were characterized via lumi-
nescence spectrophotometer, Fourier Transform Infrared (FTIR), X-ray diffraction (XRD), and scanning
Keywords:
electron microscopy (SEM) to obtain the photoluminescence spectra, chemical bonding, crystallinity,
Colloidal precursor
Photoluminescence
and morphology of the particles, respectively. The photoluminescence emission recorded at 250 nm exci-
Flame spray method tation showed a peak at approximately 400 nm that was characterized by a violet band. The highest
Particle intensity was shown by a 75 mol% ZnO sample followed, in order, by 50 mol% ZnO, 25 mol% ZnO, and
Dryer pure ZnO.
Nanocomposites Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction favorably to those of relatively expensive phosphor materials.


Other applications of zinc oxide particles include the following:
New energy-conserving light sources have generated a tremen- high-density data storage systems, transparent electrodes, gas
dous amount of interest due to increasing energy demands and sensing, bio-detection, UV light emitters, window material for dis-
concerns of global warming. Solid-state light emitting diodes plays, solar cells, and lasers [2].
(LEDs) are now receiving increased attention as energy-saving Aerosol or wet processes can be utilized to synthesize ZnO
sources, particularly phosphor-converted LEDs due to their high nanoparticles. Particles on the nano scale are preferred due to
luminescence efficiency. Phosphors are usually limited to mono- the unique and improved properties that are a result of their size,
chromatic emissions. Therefore, to generate white light, a blend composition and structure [3]. ZnO nanoparticles were synthesized
of phosphors that emits in blue, green and red wavelengths is using an integrated pulse combustion-spray pyrolysis process [4].
required. Zinc oxide is a wide-band-gap semiconductor that exhib- The fluctuations in the flow of hot gasses in pulse combustion sys-
its a strong and stable broad-band emission. ZnO has been consid- tems enhance energy transfer and promote an increase in crystal-
ered as an alternative source to replace expensive GaN-based LEDs line nanoparticles compared with the micron-sized particles that
[1], because they have semiconducting properties that compare are produced by conventional spray pyrolysis. Other methods
include gel-template combustion [5] and flame spray pyrolysis
⇑ Corresponding author. Tel.: +62 31 5946240; fax: +62 31 5999282. [6]. An aerosol-based spray-drying system for producing particles
E-mail address: widi@chem-eng.its.ac.id (W. Widiyastuti). with a controlled morphology has been described in detail [7].

http://dx.doi.org/10.1016/j.cej.2014.05.104
1385-8947/Ó 2014 Elsevier B.V. All rights reserved.
W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258 253

Besides aerosol processing, the synthesis of ZnO nanoparticles 1240 294:0 1:09
¼ 3:301 þ 2
þ ð1Þ
through wet-chemical processing is also an area of particular inter- k1=2 dp dp
est, because it provides an economical alternative using moderate
temperatures and pressures. Particle size is controlled by the rate Colloidal silica particles were prepared by diluting water glass
of nucleation and coagulation that is highly influenced by synthesis into distilled water to generate a 0.1 M colloidal solution at a tem-
conditions such as temperature, precursor concentration, and sol- perature of 60 °C. After the temperature was decreased to room
vent type. A synthesis of ZnO colloidal nanoparticles with sizes that temperature, the water glass solution was passed through a cation
range from 3 to 6 nm has been demonstrated via a sol–gel exchange resin to form a silicic acid solution. To the silicic acid
approach [8]. However, an increase in particle size after aging solution, 0.1 M KOH was added drop-by-drop while the solution
resulted in a significant color shift in the photoluminescence spec- was stirred rigorously to adjust the pH to 8. Under these condi-
tra [9]. Nanoparticles tended to coagulate and form larger particles, tions, the solution was polymerized to form colloidal silica parti-
which led to a decrease in their functional characteristics. In an cles [16]. The particle size distribution of the prepared colloidal
attempt to maintain the characteristics of nanoparticles, they were silica, as measured by Malvern Particle Size Distribution, is
mixed with inert material to generate a nanocomposite. Further- depicted in Fig. 1. The average geometric particle size (dg) and its
more, trapping the particles in a solid matrix hindered the growth geometric standard (rg) deviation were 11.25 and 1.29, respec-
of ZnO nanoparticles, and the properties of the particles would not tively. The surface charges of colloidal silica and colloidal ZnO par-
change. Nanocomposite structures are formed by incorporating ticles were 13.10 and 5.00 mV, respectively, by measuring their
functional nanoparticles into a transparent matrix material, which zeta potentials (Zetasizer, Malvern).
stabilizes the particle size and growth and maintains the enhanced The colloidal zinc oxide and silica were mixed to obtain 25, 50,
nanoparticles [10]. and 75 mol% ZnO. For comparison, pure colloidal zinc oxide and sil-
Transparent silica nanoparticles are well suited for use in a ica were also used as precursors. The mixed colloidal solution was
matrix because they have almost no effect on the embedded nano- used as a precursor for the flame-assisted spray-drying system, as
particles. Variations in synthesis methods cause ZnO nanoparticles shown in Fig. 2. In the flame reactor, propane gas was used as a fuel
to exhibit different optical properties when embedded into amor- and air was used as an oxidizer. The flow rate of the fuel was
phous silica nanoparticles by varying the particle size and compo- 0.50 L/min and the flow rate of the oxidizer was 10-fold greater.
sition [11]. A matrix of silica particles can play an important role in In order to carry uncharged droplets from the nebulizer (ultrasonic
determining the controllable luminescence position of ZnO nano- nebulizer, NE-U17, Omron) to the flame reactor, air with a flow
particles. Zinc oxide–silica nanocomposites have been prepared rate of 2 L/min was used as the carrier gas.
by sol–gel and spray-combustion methods using tetraethoxysilane In order to examine the water content of pure ZnO and pure sil-
(TEOS) and hexamethyldisiloxane (HMDSO), respectively [12,13]. ica samples, thermogravimetry analysis (TGA, Mettler Toledo) was
However, TEOS and HMDSO are relatively expensive and toxic used to measure the change in weight of the generated particles by
organic precursor materials. Zinc oxide nanoparticles embedded increasing the temperature at a heating rate of 10 °C/min. Scanning
in a silica nanoparticle matrix were investigated by consecutive electron microscopy (SEM, Zeiss) at 20 kV was used to examine the
sol–gel and spray-drying methods by using commercial colloidal morphology of the generated particles. An X-ray diffractometer
silica [14,15]. Water glass is a promising candidate for the synthe- (XRD, Philips) using filter CuKa radiation (k = 1.54 Å) operated at
sis of silica colloidal nanoparticles, and it is both relatively 40 kV and 30 mA was used to record the diffraction spectra with
inexpensive and plentiful. The preparation of colloidal silica nano- 2h from 10° to 60°. The Fourier Transform Infrared (FTIR, Perkin
particles from water glass has been accomplished under both Elmer) spectrum was analyzed for wave numbers ranging from
acidic and basic conditions [16]. Silicic acid was generated before 4000–500 cm1. FTIR was used to characterize the type of chemical
the formation of colloidal silica nanoparticles. In addition, silicic bonding by evaluating the chemical bond’s vibrating energy that
acid can be synthesized from biomass that is abundant in tropical was altered by the absorbance of infrared radiation. A lumines-
regions. cence spectrometer (Perkin Elmer) equipped with a xenon source
The objective of this work was to synthesize a ZnO/SiO2 nano- was used to examine the excitation and emission luminescence
composite using a flame-assisted spray drying method with mixed spectra.
colloidal ZnO and silica precursors synthesized by sol–gel methods.
Colloidal ZnO and silica were synthesized from zinc acetate and
water glass, respectively. The use of water glass as a silica source
was a novel attempt in a ZnO/SiO2 nanocomposite. The silica con-
tent in the colloidal precursor was studied for its effect on the char-
acteristics of the generated particles.

2. Material and methods

Zinc acetate (Zn(CH3COO)22H2O, p.a. 99.5%, Merck), lithium


hydroxide (LiOH, 98%, Merck), ethanol (absolute, Merck), potas-
sium hydroxide (KOH, Merck), and water glass (Na2OnSiO2mH2O,
8% Na2O, 27% SiO2, Merck) were used as received. Colloidal zinc
oxide particles were prepared based on a method established by
Spanhel and Anderson using 0.1 M zinc acetate and 0.14 M lith-
ium hydroxide with ethanol as a solvent [8]. By using the correla-
tion between average particle diameter (dp) and the absorption
shoulder (k1/2), based on the UV–Vis absorbance measurement
(spectrophotometer, Thermo Scientific Genesys), at the wave-
length where the absorption was half of the excitonic peak, the
size of the colloidal ZnO was 3.75 nm. The correlation was as fol-
lows [17]: Fig. 1. Particle size distribution of as-prepared colloidal silica.
254 W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258

to complete evaporation and resulted in nanoparticle formation


[18]. Fig. 3b shows that the content of ZnO was 75 mol%. The
solvent contained 75 volume% of ethanol and the rest was water.
The particles remained smaller compared with the particles gener-
ated by an ethanol solvent with volume contents of 50% and 25%,
as depicted by Fig. 3c and d, respectively. The 50 and 25 mol%
ZnO particles were generated one-droplet-to-one-particle, but
the surfaces of the particles were not smooth compared to pure sil-
ica using only water as a solvent, as shown in Fig. 3e. This indicated
that the selection of the solvent plays a role in forming distinctions
in morphology size, shape, and surface.
In order to examine the distribution of zinc oxide and silica
nanoparticles in the composite, in the case of the 25 mol% ZnO, ele-
mental mapping of Zn, Si, and O by EDX analysis was done, as
shown in Fig. 4, which shows that both the ZnO and silica nanopar-
ticles were well-dispersed among the others. ZnO nanoparticles
were smaller than the silica nanoparticles, and had a higher degree
of kinetic energy. However, the surface charge of the silica nano-
particles was larger than that of the ZnO nanoparticles even though
both were negatively charged. The same charges led to the repul-
sive interactions among colloidal nanoparticles. Size and surface
charge contributed to the homogeneity of the generated nanocom-
posite particles by maintaining the well-dispersed nature of the
Fig. 2. Flame-assisted spray-drying system. nanoparticles inside the droplet. Further explanation of the effects
of colloidal surface charge, size and concentration on particle mor-
phology has been discussed in detail for silica–polystyrene colloi-
3. Results and discussion dal systems in a previous study [19].
The particle size distribution calculated from SEM images using
SEM images of the samples are shown in Fig. 3. When using the Image-J free software shown in Fig. 5 indicates that the mean
pure ZnO, as shown in Fig. 3a, smaller particles were observed particle size was increased with the silica content. The morphology
which did not follow the one-droplet-to-one-particle mechanism of the particles was more spherical with increases in the silica con-
that is common to particle formation using a spray-pyrolysis pro- tent. However, particles with a relatively broad size distribution
cess [3]. This smaller-particle morphology was induced by the were generated because droplets were uncharged or neutral during
breakup of droplets contained in the precursor solution, because spraying. The morphology of particles generated from uncharged
the use of ethanol as a solvent led to an increase in the evaporation droplet had broad size distribution and agglomerated particles
rate compared to when water was used as the solvent. Ethanol has [20].
a lower boiling point and increases the energy for combustion, TG-DTA analysis for the generated particles of pure silica and
because it can also act as a fuel. The same result has also been pure ZnO is shown in Fig. 6. The analysis was performed under
reported when methanol has been used as a solvent—increasing air atmospheric conditions within a temperature range of from
the methanol content in a water–methanol solvent mixture led 25 to 600 °C. The water content of the generated particles was

Fig. 3. SEM image of the generated particles from a precursor of (a) pure colloidal zinc oxide, (b) 75 mol% zinc oxide, (c) 50 mol% zinc oxide, (d) 25 mol% zinc oxide, and (e)
pure colloidal silica.
W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258 255

Fig. 4. Elemental mapping of Si, Zn, and O of ZnO/SiO2 nanocomposite (25 mol% zinc oxide).

Fig. 5. Particle size distribution of the generated particles from a precursor of (a) pure colloidal zinc oxide, (b) 75 mol% zinc oxide, (c) 50 mol% zinc oxide, (d) 25 mol% zinc
oxide, and (e) pure colloidal silica.

examined based on decreases in sample weight at temperatures peak of silica at around 2h = 22° can be assigned to the silica in
from 25 to 100 °C. The weight of pure silica and pure ZnO samples an amorphous phase. The XRD pattern shows the characteristic
at 100 °C were 99.93% and 99.71% of the original amount, respec- peaks of silica as an amorphous material between 2h = 20° and 40°.
tively. By assuming water evaporates completely at 100 °C, the The intensity was decreased by the addition of silica concentra-
water content of the pure silica and pure ZnO samples were only tion, as shown in Table 1. The crystalline size that was calculated
0.07% and 0.29%. At 600 °C, the weight loss for pure silica and pure based on the Scherer number used a half-width of the highest peak
ZnO samples was only 2.43% and 5.45%, respectively. In addition, in the XRD pattern—the [1 0 1] peak at 2h = 36.253°. The crystalline
no exothermic or endothermic peaks in the heat flow curve indi- size was decreased by adding colloidal silica in the precursor,
cated that no reaction or phase transformation had occurred. which was caused by a decrease in the final zinc precursor concen-
Nanocomposite particles were not analyzed by assuming the water tration after adding an amount of colloidal silica, as shown in
content was in the range of pure silica and pure ZnO samples. Table 1.
Fig. 7 shows the XRD patterns of the generated particles for var- Fig. 8 shows the FTIR pattern of the samples, which illustrates
ious zinc oxide–silica compositions. For particles of 25, 50, 75, and the type of chemical bonding by pure zinc oxide, pure silica, and
100 mol% ZnO, the ZnO structure could be characterized as hexag- the composites of zinc oxide–silica. For pure zinc oxide and
onal wurtzite (JCPDS 36-1451). On the other hand, the diffraction 25–75% ZnO samples, infrared absorption reduction peaks were
256 W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258

Table 1
Precursor composition of zinc oxide/silica particles.

%ZnO Volume of Volume of Final ZnO Final silica dc


sol ZnO sol silica concentration concentration [nm]
[mL] [mL] [M] [M]
100 50 0 0.100 0 24.20
75 50 16.7 0.075 0.025 21.82
50 50 50 0.050 0.050 18.19
25 50 150 0.025 0.075 16.79
0 0 50 0 0.100 –

of the siloxane link [21]. Reduction peaks at around 1600 cm1


were only observed for samples that contained zinc oxide. These
reduction peaks were attributed to the bending vibration from
absorbed water [22]. The combination of the bend vibration of
Si–O–Si and the stretching modes of Zn–O are indicated by the
reduction peaks at around 400 cm1 that were only apparent in
ZnO/SiO2 nanocomposite samples: ZnO 25%, ZnO 50%, and ZnO
75%.
When the samples were excited at a wavelength of 250 nm at
room temperature, PL emission peaks of 25%, 50%, 75%, and 100%
ZnO were located at 401.5, 394.5, 396, and 399.5 nm, respectively,
as shown in Fig. 9. The peak intensities did not significantly change
the wavelength position with silica concentration in the compos-
ites. The PL emission peaks of nanocomposite samples were not
dependent on the particles size of the bulk powder. This result
indicated that the samples showing a noticeable near ultra violet
– violet region were matched to near band-edge ZnO emissions
as a result of the recombination of free excitons [23]. An emission
spectra that consists of near-UV, violet, blue, and green reportedly
will reveal bands of 3.4, 3.07, 2.8, and 2.4 eV, respectively, and
wavelengths of 365, 404, 443, and 517 nm, respectively [24]. As
the figure shows, the half band gap covered by wavelengths rang-
ing from 345–480 nm corresponds to the near ultra violet – green
bands. In this case, the silica concentration only affected the PL
intensity peak of the composites, but was incapable of shifting
Fig. 6. Thermo-gravimetric and differential analysis (TG-DTA) for the generated the band. The maximum intensities of PL emissions of 25%, 50%,
particles of (a) pure silica, (b) pure zinc oxide. and 75%, were 2.8-, 3.6-, and 3.9-fold higher, respectively, than
pure ZnO. The emission band covering the visible region was
observed near 3500 cm1, and were attributed to the hydroxyl broad, which can be attributed to the deep-level defects of ZnO
group of water. This reduction peak was not observed for pure at different levels of energy.
silica. At around 1100 cm1, a significant reduction peak was Photoluminescence excitation detected at an emission wave-
clearly detected for the pure silica sample and insignificant peaks length of 500 nm is shown in Fig. 10. The highest intensities were
were unmistakable for the nanocomposites samples, which indi- observed at 247, 248.5, and 250 nm wavelengths for samples of
cated asymmetric stretching vibration modes of the Si–O–Si bridge 25%, 50%, and 75% volumes of ZnO, respectively. On the other hand,

Fig. 7. X-ray diffraction pattern of particles at various mole percentages (a) at the same scale of intensity and (b) at a higher scale of intensity only for silica XRD pattern.
W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258 257

for the same emission wavelength, pure ZnO at a wavelength of


approximately 250 nm showed the lowest intensity. That result
indicated that composing ZnO with silica resulted in a higher per-
formance for photoluminescence by comparison with pure ZnO.

4. Conclusion

In summary, ZnO/SiO2 nanocomposites were synthesized by


using consecutive sol–gel and flame-assisted spray-drying meth-
ods. A flame reactor was used to spray the mixed colloidal zinc
oxide and silica precursor from zinc acetate and water glass,
respectively. Water glass used as a silica source was a novel
attempt in a ZnO/SiO2 nanocomposite. The use of water and etha-
nol as solvents in colloidal silica and zinc oxide, respectively,
increased the silica content in the precursor, and led to an increase
in the particle size. The pure silica sample had an amorphous struc-
ture. The ZnO/SiO2 nanocomposite was characterized as a hexago-
nal wurtzite of ZnO, which was the same structure as pure zinc
oxide. The crystalline size was increased with an increase in the
ZnO content. The highest photoluminescence intensity was
observed for the 75 mol% ZnO sample followed by the 50,
25 mol%, and pure ZnO particles, in order, which emitted at a
Fig. 8. Fourier Transform Infrared pattern of particles at various mole percentages.
wavelength approximately of 400 nm when they were excited at
a wavelength of 250 nm, which corresponded to a violet band.

Acknowledgements

This research was supported by the Indonesia Toray Science


Foundation (ITSF) under the scheme of Science and Technology
Research Grant 2012 (STRG 2012). We also extend our gratitude
to Ms. Paulina Ruliawati for assistance with the experiment.

References

[1] P. Thiyagarajan, M. Kottaisamy, N. Rama, M.S.R. Rao, White light emitting


diode synthesis using near ultraviolet light excitation on zinc oxide–silicon
dioxide nanocomposite, Scr. Mater. 59 (7) (2008) 722–725.
[2] S.J. Pearton, D.P. Norton, K. Ip, Y.W. Heo, T. Steiner, Recent progress in
processing and properties of ZnO, Superlattices Microstruct. 34 (1–2) (2003)
3–32.
[3] K. Okuyama, I.W. Lenggoro, Preparation of nanoparticles via spray route, Chem.
Eng. Sci. 58 (3–6) (2003) 537–547.
[4] W. Widiyastuti, W.-N. Wang, A. Purwanto, I.W. Lenggoro, K. Okuyama, A pulse
combustion-spray pyrolysis process for the preparation of nano- and
submicrometer-sized oxide particles, J. Am. Ceram. Soc. 90 (12) (2007)
3779–3785.
Fig. 9. Photoluminescence emission spectra of particles at an excitation of 250 nm.
[5] J. Zhou, Y. Wang, F. Zhao, Y. Wang, Y. Zhang, L. Yang, Photoluminescence of ZnO
nanoparticles prepared by a novel gel-template combustion process, J. Lumin.
119–120 (2006) 248–252.
[6] D.J. Seo, S.B. Park, Y.C. Kang, K.L. Choy, Formation of ZnO, MgO and NiO
nanoparticles from aqueous droplets in flame reactor, J. Nanopart. Res. 5 (3–4)
(2003) 199–210.
[7] A.B.D. Nandiyanto, K. Okuyama, Progress in developing spray drying methods
for the production of controlled morphology particles: from the nanometer to
submicrometer size ranges, Adv. Powder Technol. 22 (1) (2011) 1–19.
[8] L. Spanhel, M.A. Anderson, Semiconductor clusters in the sol–gel process:
quantized aggregation, gelation, and crystal growth in concentrated ZnO
colloids, J. Am. Chem. Soc. 113 (8) (1991) 2826–2833.
[9] S. Monticone, R. Tufeu, A.V. Kanaev, Complex nature of the UV and visible
fluorescence of colloidal ZnO nanoparticles, J. Phys. Chem. B 102 (16) (1998)
2854–2862.
[10] L.L. Beecroft, C.K. Ober, Nanocomposite materials for optical applications,
Chem. Mater. 9 (6) (1997) 1302–1317.
[11] M. Abdullah, S. Shibamoto, K. Okuyama, Synthesis of ZnO/SiO2
nanocomposites emitting specific luminescence colors, Opt. Mater. 26 (1)
(2004) 95–100.
[12] Y. Yang, Y.-Q. Li, H.-Q. Shi, W.-N. Li, H.-M. Xiao, L.-P. Zhu, Y.-S. Luo, S.-Y. Fu, T.
Liu, Fabrication and characterization of transparent ZnO–SiO2/silicone
nanocomposites with tunable emission colors, Compos. B 42 (8) (2011)
2105–2110.
[13] L. Madler, W.J. Stark, S.E. Pratsinis, Rapid synthesis of stable ZnO quantum
dots, J. Appl. Phys. 92 (11) (2002) 6537–6540.
[14] R. Mikrajuddin, F. Iskandar, K. Okuyama, F.G. Shi, Stable photoluminescence of
zinc oxide quantum dots in silica nanoparticles matrix prepared by the
Fig. 10. Photoluminescence excitation spectra of particles at an emission of combined sol–gel and spray drying method, J. Appl. Phys. 89 (11) (2001)
500 nm. 6431–6434.
258 W. Widiyastuti et al. / Chemical Engineering Journal 254 (2014) 252–258

[15] N. Hagura, T. Takeuchi, S. Takayama, F. Iskandar, K. Okuyama, Enhanced [20] A. Suhendi, A.B.D. Nandiyanto, M.M. Munir, T. Ogi, L. Gradon, K. Okuyama, Self-
photoluminescence of ZnO–SiO2 nanocomposite particles and the analyses of assembly of colloidal nanoparticles inside charged droplets during spray-
structure and composition, J. Lumin. 131 (1) (2011) 138–146. drying in the fabrication of nanostructured particles, Langmuir 29 (43) (2013)
[16] H.C. Liu, J.X. Wang, Y. Mao, R.S. Chen, The preparation and growth of colloidal 13152–13161.
particles of concentrated silica sols, Colloids Surf. A 74 (1) (1993) 7–13. [21] G. Liu, G. Hong, D. Sun, Synthesis and characterization of SiO2/Gd2O3:Eu core-
[17] Y. Yang, W.-N. Li, Y.-S. Luo, H.-M. Xiao, S.-Y. Fu, Y.-W. Mai, Novel ultraviolet- shell luminescent materials, J. Colloid Interface Sci. 278 (1) (2004) 133–138.
opaque, visible-transparent and light-emitting ZnO-QD/silicone composites [22] S. Panigrahi, A. Bera, D. Basak, Ordered dispersion of ZnO quantum dots in SiO2
with tunable luminescence colors, Polymer 51 (12) (2010) 2755–2762. matrix and its strong emission properties, J. Colloid Interface Sci. 353 (1)
[18] T. Tani, A. Kato, H. Morisaka, Effects of solvent on powder characteristics of (2011) 30–38.
zinc oxide and magnesia prepared by flame spray pyrolysis, J. Ceram. Soc. Jpn. [23] S. Chawla, K. Jayanthi, R.K. Kotnala, Room-temperature ferromagnetism in Li-
113 (3) (2005) 255–258. doped p-type luminescent ZnO nanorods, Phys. Rev. B 79 (12) (2009).
[19] A.B.D. Nandiyanto, A. Suhendi, O. Arutanti, T. Ogi, K. Okuyama, Influences of 125204(1-7).
surface charge, size, and concentration of colloidal nanoparticles on [24] J.-H. Hong, Y.-F. Wang, G. He, J.-X. Wang, The effect of calcination temperature
fabrication of self-organized porous silica in film and particle forms, on the photoluminescence from sol–gel derived amorphous ZnO/silica
Langmuir 29 (21) (2013) 6262–6270. composites, J. Non-Cryst. Solids 356 (50–51) (2010) 2778–2780.

S-ar putea să vă placă și