Wu Et Al-2013-Crystal Research and Technology

You might also like

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

Cryst. Res. Technol. 48, No. 3, 145152 (2013) / DOI 10.1002/crat.

201200438

Structure and optical properties of Mg-doped ZnO nanoparticles


by polyacrylamide method
Yan Wu*, Jin Yun, Linqin Wang, and Xiang Yang
Faculty of Materials Science and Chemistry, China University of Geosciences, Wuhan 430074, China
Received 21 October 2012, revised 30 January 2013, accepted 31 January 2013
Published online 27 February 2013
Key words Mg-doped ZnO nanoparticles, polyacrylamide method, optical properties.
Mg-doped ZnO (Mgx Zn1-x O) nanoparticles with precise stoichiometry are synthesized through polyacrylamide
polymer method. Calcination of the polymer precursor at 650 C gives particles of the homogeneous solid
solution of the Mgx Zn1-x O system in the composition range (x < 0.15). ZnO doping with Mg causes shrinkage
of lattice parameter c. The synthesized Mgx Zn1-x O nanoparticles are typically with the diameter of 70 85 nm.
Blue shift of band gap with the Mg-content is demonstrated, and photoluminescence (PL) from ZnO has been
found to be tunable in a wide range from green to blue through Mg doping. The blue-related PL therefore
appeared to be caused by energetic shifts of the valence band and/or the conduction band of ZnO. Mgx Zn1-x O
nanoparticles synthesized by polyacrylamide-gel method after modified by polyethylene glycol surfactant have
a remarkable improvement of stability in the ethanol solvent, indicating that these MZO nanoparticles could
be considered as the candidate for the application of solution processed technologies for optoelectronics at
ambient temperature conditions.

C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Introduction

ZnO, an important semiconductor material with direct band gap 3.2 eV, shows attractive potential in UV laser
diodes and light emitting diodes due to its large exciton binding energy (60 meV) at room temperature. ZnO
nanostructures have been also extensively investigated in order to develop more efficient optoelectronic devices
[13]. One of the mostly challenges of application of these optoelectronics is to develop a simple low-cost
manufacturing method based on solution process at ambient temperature conditions [2,46].
Recently Mg-doped ZnO materials have attracted much attention because of their unique UV-luminescent
properties based on radiative recombination of the electronhole pairs [7,8]. Stoichiometry and homogeneity of
composition are key to the potential applications of such mixed metal oxides [9]. Previous attempts at doping
nanocrystals have been fraught with problems because the synthetic schemes used to dope frequently yield
inhomogeniously doped materials. Wet chemistry synthesis route seems to meet these requirements, enabling
the preparation of metal ion doped ZnO with a high crystalline quality and precise control of stoichiometry
by adjusting the different ion proportions in the solution. Fabricating doped ZnO nanoparticles in the networks
of hydro-, micro-, and nano-gel systems is also considered as the most important approach due to its direct
applicability in various catalyst and electronic devices [1013]. Polyacrylamide gel process is a well-known
method for synthesizing various mixed metal ions oxide nanoparticles at a large scale. This polymer gel method
has been used to prepare many different oxide nanoparticles, metallic and oxide compounds, such as NdFe10 Mo2
[14] and Ce1-x Bix O2-x/2 solid electrolytes [15]. With polymeric chains forming a network, metal ions are entrapped
evenly within the polyacrylamide gel, which is helpful for forming uniform oxide nanoparticles [16]. Therefore,
in this study, we develop a solution-processing method, i.e. polymer gel synthesis of smaller size and finer
distribution of magnesium doped zinc oxide nanoparticles in hydrogel networks and test its suspension stability

Corresponding author: e-mail: wuyan@cug.edu.cn


C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

146

Y. Wu et al.: Structure and optical properties of Mg-doped ZnO nanoparticles by polyacrylamide method

under the treatment of polyethylene glycol (PEG) surfactant, which could be as the candidate metal oxide
materials for the future application of solution-processed optoelectronics at ambient temperature conditions.

Experimental

2.1 Synthesis of MZO nanoparticles The system Mgx Zn1-x O (MZO) was prepared as polycrystalline
nanoparticles with various compositions (0 x 0.15) by applying the sol-gel technique. Mg-doped ZnO
nanoparticles were synthesized by the following procedure. Zinc acetate dihydrate (Zn(CH3 COO)2 2H2 O),
Magnesium acetate (Mg(CH3 COO)2 ), were used as starting precursors to prepare solution in water. The pH value
of aqueous solution was adjusted to 5 by adding HNO3 . Acrylamide (AM) and N, N -methylenebisacrylamide
(MABM) were used as a monomer and as a network reagent, respectively. After mixed AM and MABM (the
mass ratio is 5:1 if not specially mentioned), the mixture were added to 0.2 mol/L metal salt aqueous solution.
The aqueous solution was heated to solute completely, and then polymerized at 80 C for 2 hours to obtain a white
translucent wet gel by a solicitation action of ammonium persulfate (NH4 )2 S2 O8 . The amount of (NH4 )2 S2 O8 is
0.1% of the mass amount of AM/MABM mixture. The wet gel was dried to obtain a dried gel in a drying cabinet
at 120 C for about 10 hours. The xerogel was calcined at 400 C for 1 hour to burn the organic compounds.
Finally the preheated powder was sintered at high temperature for 1 hour to form the crystallized nanoparticles
and prepared for different characterization.
2.2 Characterization of MZO nanoparticles Thermal decomposed properties were characterized by STA
409 PC simultaneous Thermogravimetric Analyzer and Differential Scanning Calorimeter (TGA-DSC), with the
heating rate of 10 C/min in air. The structural properties were studied by XRD spectra recorded using Bruker
AXS-D8 X-ray diffractometer operated at 40 kV and 40 mA with Cu K radiation ( = 1.5406 ). The data
were collected with a step size of 0.01 and a counting time of 5 s per step in the 2 range from 10 to 70 . FEI
Quanta200 Environmental Scanning Electron Microscopy equipped with energy-dispersive X-ray spectroscopy
(EDX) was operated at 25 kV. Ultraviolet-visible (UV-vis) absorption experiments were performed on a Lambda
35, Perkin-Elmer double beam UV-vis absorption spectrometer equipped with a deuterium lamp and a tungstenhalogen lamp. The samples for this study were used in the form of powder and pure BaSO4 used as reference.
The optical absorption measurements of the prepared samples were done by pressing the MZO nanopowders into
copper sample holders with the diameter of 10 mm. Room temperature photoluminescence (PL) of the samples
were recorded on a Jasco Instruments FP-6500 fluorescence spectrophotometer with a xenon arc lamp as the
excitation source.
2.3 Evaluation of dispersion stability Different concentrations of PEG surfactant in ethanol (25 mg/ml,
5 mg/ml, and 2.5 mg/ml) were prepared. Then MZO (Mg0.05 Zn0.95 O, 0.5 g) nanopowders were added into each
batch of PEG solutions for ultrasonic treatment. The ultrasonic agitation time was 60 min with the bath temperature
50 C. Sequentially the suspension was evaporated into dried powders in the preheated oven with the temperature
of 120 C for 24 h. Finally each batch of dried powders mixed with 100 ml ethanol was under ultrasonic treatment
for 30 min to be ready for the evaluation of dispersion stability.
The apparent sedimentation stability (S) was assessed by measuring the sedimentation speed of the ZnO and
PEG-ZnO nanoparticles suspension in ethanol (0.5 g in 100 ml ethanol) with the aging time. They were located
into a glass tube of 0.5 cm inner diameter and 20 cm length, respectively. Then the sedimentation stabilities were
monitored at a given aging time, as follows [17]:
S(%) =

lt
100
l0

(1)

where, l0 is the length of initial opaque dispersion and lt is the length of sediment part at a given time t in the
tube, respectively.

Results and discussion

3.1 Thermogravimetric/Differential thermal analysis Figure 1 shows the TGA-DSC curves of 5% Mgdoped ZnO xerogel precursor under the synthesized condition of 5:1 mass ratio of the monomer AM and the
network reagent MABM. The xerogel slightly loses weight up to 5% when the temperature increasing from 50
to 288 C, which could correspond to the decomposition of small-molecule end-groups like those of CN, CH

C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Cryst. Res. Technol. 48, No. 3 (2013)

147

Fig. 1 TG-DSC curves of 5% Mg-doped ZnO xerogel


precursor.

Fig. 2 The X-ray diffraction patterns of the


5% Mg-doped ZnO nanopowders samples calcinating at different temperatures.

and C = O. When further increasing the temperature up to about 650 C, the weight drops down to about 50%.
During this temperature range, there are two broad exothermal peaks centered at 540 C and 590 C, which
corresponds to the decomposition of carboxyl groups [18]. After 650 C, the weight tends to keep constant.
Thereby we conclude that the organic compounds have been completely ruled out and the crystallized oxide
forms.
3.2 Structural analysis

3.2.1 The effect of calcinating temperature


Figure 2 presents four different spectra of 5% Mg-doped ZnO nanoparticles calcined at different temperatures.
For as-synthesized powder calcined at 400 C, there are seven broad peaks corresponding to (100), (002),
(101), (102), (110), (103), and (112) peaks which corresponding to the standard ZnO pattern of hexagonal ZnO
(JCPDS card No. 361451). But there are two more extra peaks marked by red arrows for the impurities of zinc
carboxyl-containing compounds. When the annealing temperature increasing to 550 C, all the peaks belong to
the hexagonal lattice of ZnO, and no indication of a secondary phase is found. The peaks become sharp with the
annealing temperature increasing further. The lattice parameters a and c were calculated from (100) and (002)
oriented XRD peaks using equations (2) and (3) [19]. The grain sizes of the crystallite were calculated from the
full width at half maximum (FWHM) of the peak by using the Debye Scherrer formula [20]. The average particle
sizes (d) of MZO samples were estimated from the width of lines in the XRD spectrum using the Sherrers
equation (4).
www.crt-journal.org


C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

148

Y. Wu et al.: Structure and optical properties of Mg-doped ZnO nanoparticles by polyacrylamide method

Fig. 3 The X-ray diffraction patterns of the


ZnO powder samples synthesized with different
molar ratio of the monomer and the network
reagent.

a=

3 sin

(2)

c=

sin

(3)

d=

0.9
cos

(4)

where is the X-ray wavelength (1.5406 ), is the width of the line at the half-maximum intensity, and is the
half of the diffraction angle. The d was calculated for the highest three intense peaks of ZnO spectrum which are
(101), (002), and (100). The average calculated particle size results are 39 3 nm, 54 3 nm, 72 3 nm, and
76 3 nm for the samples calcinating at 400 C, 550 C, 650 C, and 750 C, respectively.
3.2.2 The effect of the mass ratio of monomer and network reagent
We introduced the AM-MABM copolymer in order to investigate the factors affecting the nanoparticles size by
controlling the crystalline growth process. The mass ratio R of monomer and the network reagent varies from
3:1 to 7:1. The calcinating temperature is 650 C and the calcinating time is 1 hour. figure 3 presents their X-ray
diffraction patterns, which demonstrate that all of them have formed the crystallized wurtzite ZnO structure.
Based on the Sherrers equation, the particle sizes are 61 3 nm, 72 3 nm, 72 3 nm, 77 3 nm and 100
3 nm for the ratio of 3:1, 4:1, 5:1, 6:1, and 7:1 respectively. It is obvious that the nanoparticles size increases
dramatically when the ratio of the monomer and the network reagent in the polymer synthesis is large than 6:1.
The mass ratio of 6:1 could be the crucial point to change the network into different structure because that the
relative volumes of the network structures vary depending on the crosslinker concentration [16,21].
3.2.3 The effect of the various Mg dopant concentrations
The full X-ray diffractions of the powder samples with varying Mg doping rate (x = 00.15) at 650 C for
30 min in the air are depicted in figure 4a. Up to x < 0.15, all the peaks belong to the hexagonal lattice of ZnO,
and no indication of a secondary phase is found. For x 0.15 two phases ZnO and MgO were identified in the
diffraction patterns. Therefore, the solubility limit of Mg2+ in ZnO was determined to be lower than x = 0.15
in this work. This value apparently higher than that of about 4.25% by the solvothermal method [22], but lower
than those of 33% with pulse laser deposition [23], 16.5% by metaorganic chemical vapor deposition [24]. While
it is close to those of about 15% reported with rheological phase reaction precursor route [1] and 10% when
using a solution chemistry method [25]. The composition and percentage of Mg in doped ZnO samples were

C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Cryst. Res. Technol. 48, No. 3 (2013)

149

Table 1 Calculated lattice parameters of MZO samples calcined at 650 C. Estimated standard deviations for the least
significant digit are given in parentheses.
Sample
ZnO
Mg0.02 Zn0.98 O
Mg0.05 Zn0.95 O
Mg0.10 Zn0.90 O
Mg0.125 Zn0.975 O
Mg0.15 Zn0.85 O

a ()

c ()

c/a

Particle size (nm)

3.250(6)
3.260(6)
3.250(7)
3.250(8)
3.251(3)
3.251(9)

5.206(3)
5.205(6)
5.205(2)
5.203(7)
5.202(2)
5.205(4)

1.601(7)
1.601(4)
1.601(2)
1.600(7)
1.600(2)
1.600(8)

77 3
82 3
84 3
72 3
77 3
81 3

Fig. 4 The full (a) and enlarged (b) X-ray diffraction patterns of the powder samples Mgx Zn1-x O for x = 0.01,
0.05, 0.1, 0.125 and 0.15, running upwards.

confirmed by energy dispersive X-ray (EDX) spectroscopy. The substitution of Mg2+ in ZnO was confirmed by
lattice parameters as a function of the dopant content of Mg. The lattice parameters are important to determine
whether Mg (II) have been doped into the lattice of ZnO nanoparticles or not. Variation of the lattice parameters
a, c, c/a ratio and particle sizes for samples calcined at 650 C are presented in table 1. The crystal sizes of these
samples are around 7085 nm. The lattice parameter c of the MZO clearly shrinks with the dopant level up to
12.5% as shown in figure 4b, which is in agreement with the data obtained from poly(acrylic acid) based method
[10,23] and epitaxially grown films on sapphire by a laser deposition method [23]. However, when the dopant
concentration further increases to 15%, both the parameter c and the c/a ratio start increasing. These results reveal
that the limitation concentration of Mg dopant is less than 15%, which is consistent with the results of XRD
pattern of Mg0.15 Zn0.85 O sample presented in figure 4.
3.3 Optical properties Figure 5 depicts the absorption spectra of all the prepared ZnO and Mg doped ZnO
samples with the crystal sizes around 7085 nm. Mg doping shifts the absorption onset to blue (360338 nm)
of Mg doping levels from 0 to 15%, indicating an increase of the band gap. Moreover, with the increase in Mg
concentration, the band edge shifts toward the lower wavelength (higher energy) side. Higher doping levels result
in more pronounced shifts. The indicated band gaps are similar to that which has been observed on other forms
of ZnO doped with Mg [24,27]. The systematic shift with doping levels suggests that the introduction of dopant
ions, rather than size effects, bring about these changes [20,28]. Further, a number of previous experiments and
theoretical investigations have suggested that size effects have little or no influence on the band structure of ZnO
nanocrystals with diameters greater than 7.0 nm [28,29] .The indicated band gap shifts are similar to those results
which have been observed on other forms of ZnO doped with Mg and Cd [24,27,30]. We therefore attribute
the origin of the increase of band gap to the influence of dopant ions. Further proof follows from the emission
characteristics of the nanoparticles.
The room temperature PL spectra of Mg-doped ZnO nanoparticles excited at 325 nm are shown in figure 6.
All samples show a broad green band with a blue-shift by increasing Mg doping concentration compared to
the emission peak of undoped sample. The green emission involves transition from the band edge (or shallow
www.crt-journal.org


C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

150

Y. Wu et al.: Structure and optical properties of Mg-doped ZnO nanoparticles by polyacrylamide method

Fig. 5 UV-vis spectra of Mg-doped ZnO nanoparticles.

Fig. 6 Room temperature photoluminescence


spectra of Mg-doped ZnO nanoparticles.

level close to band) to a defect level. This emission could be due to oxygen interstitials. Also, with increasing
doping concentration, red shift of emission band decreased, which could arise from variation of surface levels
and defects levels [28]. The presence of the blue-shifting is remarkable because it suggests that the nanocrystals
do not produce extra defects due to the introduction of the Mg ions [27]. The band edge emission shifts to the
blue with increasing levels of Mg doping, reflecting the change in the exciton energy seen in the absorption
spectra.
3.4 Stability of the MZO nanoparticles modified by PEG surfactant All the MZO particles have the
similar sedimentation stability. Hereby we take Mg0.05 Zn0.95 O as an example. The sedimentation stabilities of
the bare MZO (Mg0.05 Zn0.95 O) nanoparticles and various concentrations of PEG modified MZO nanoparticles
dispersed in ethanol were assessed and shown in figure 7. All these nanoparticles are in the diameter range of
7085 nm, as shown in table 1, and also they are confirmed by SEM images (not shown here). The results show
a remarkable difference. The first nanosuspension (with bare MZO) precipitated at a higher rate. The bare MZO
nanoparticles were aggregated together because of the high surface polarity. Conventional inorganic particles
have a tendency to separate easily in solvents, because of their high density and low compatibility [31,32]. After
modified with PEG surfactant, the precipitation rate of the suspension decreases dramatically. The aggregation
of the inorganic nanoparticles is prevented. The stability increases with the concentration of the surfactant PEG.
MZO nanoparticles modified with 25 mg/ml PEG maintain their initial stability of above 85% for more than 22 h,
which is qualified for the application of solution-processed technologies [17,33].

C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Cryst. Res. Technol. 48, No. 3 (2013)

151

Fig. 7 Apparent sedimentation stability of


MZO modified with PEG in ethanol solvent.

Conclusions

We have successfully synthesized nanoparticles of Mg-doped ZnO (Mgx Zn1-x O, x < 0.15) with the diameter range
of 7085 nm by using polyacrylamide polymer method. The variation of the lattice parameters and decrease
of c/a ratio of MZO samples suggest that Mg2+ has doped into ZnO lattice and the dopant concentration is
less than 15%. SEM observation shows that all Mgx Zn1-x O samples are an assembly of ellipsoidal nanoparticles
with tiny grain size of 6080 nm. Introduction of Mg2+ does not change the particle size and particle shape.
UV-vis absorption results indicate that the Mg doping shifts the absorption onset towards lower wavelengths
(370350 nm) with increasing of Mg doping levels from 0 to 12.5%, indicating an increase of the bandgap. The
increase of band gap is attributed to the influence of dopant ions. Photoluminescence measurements confirm
these results. MZO nanoparticles modified by PEG suspension in ethanol solvent has a remarkable improvement
of stability in ethanol solvent maintaining the initial stability of above 85% for more than 22 hours. It indicates
that these nanocrystals could be considered as the candidate for the application of MZO nanoparticles in the
solution-processing technologies for optoelectronics.
Acknowledgments We thank the Fundamental Research Founds for National Universities, China University of Geosciences
(Wuhan) for financial support.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]

D. C. Look, Mater. Sci. Eng. B 80, 383 387 (2001).


S. J. Pearton, D. P. Norton, K. Ip, Y. W. Heo, and T. Steiner, Prog. Mater Sci. 50, 293 (2005).
P. Sharma, K. Sreenivas, and K. V. Rao, J. Appl. Phys. 93, 3963 3970 (2003).
Y. Wu, T. Tamaki, W. Voit, L. Belova, and K. V. Rao, Mater. Res. Soc. Symp. Pro. 1161, I03 (2009).
Y. Wu, T. Tamaki, T. Volotinen, W. Voit, L. Belova, and K. V. Rao, J. Phys. Chem. Lett. 1, 89 (2010).
Y. Wu, K. V. Rao, W. Voit, T. Tamaki, O. D. Jayakumar, L. Belova, Y. S. Liu, P. A. Glans, C. L. Chang, and J. H. Guo,
IEEE Trans. Magn. 46, 2152 (2010).
V. Etacheri, R. Roshan, and V. Kumar, ACS Appl. Mater. Interfaces 4, 2717 (2012).
U. Ozgur, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Dogan, V. Avrutin, S. J. Cho, and H. Morkoc, J. Appl.
Phys. 98, 041301 (2005).
E. R. Leite, A. P. Maciel, I. T. Weber, P. N. Lisboa-Filho, E. Longo, C. O. Paiva-Santos, A. V. C. Andrade, C. A.
Pakoscimas, Y. Maniette, and W. H. Schreiner, Adv. Mater. 14, 905 (2002).
G. Lu, I. Lieberwirth, and G. Wegner, J. Am. Chem. Soc. 128, 15445 (2006).
S. E. Ahn, J. S. Lee, H. Kim, S. Kim, B. H. Kang, K. H. Kim, and G. T. Kim, Appl. Phys. Lett. 84, 5022 (2004).
K. Huang, Z. Tang, L. Zhang, J. Y. Yu, J. G. Lv, X. S. Liu, and F. Liu, Appl. Surf. Sci. 258, 3710 (2012).
X. W. Zhu, Y. Q. Li, Y. Lu, L. C. Liu, and Y. B. Xia, Mater. Chem. Phys. 102, 75 (2007).
S. F. Liu, J. H. Lin, X. L. Qian, J. M. Bayi, and M. Z. Su, Chem. Mater. 8, 2545 (1996).
H. Zhao, S. Feng, and W. Xu, Mater. Res. Bull. 35, 2379 (2000).
R. F. T. Stepto, J. I. Cail, and D. J. R. Taylor, Mater. Res. Innovations 7, 4 (2003).

www.crt-journal.org


C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

152

Y. Wu et al.: Structure and optical properties of Mg-doped ZnO nanoparticles by polyacrylamide method

[17] P. Liu, Colloids Surf. A 291, 155 (2006).


[18] G. J. Lin, H. Yang, T. Xian, Z. Q. Wei, J. L. Jiang, and W. J. Feng, Adv. Powder Technol. 23, 35 (2012).
[19] F. K. Shan, B. I. Kim, G. X. Liu, Z. F. Liu, J. Y. Sohn, W. J. Lee, B. C. Shin, and Y. S. Yu, J. Appl. Phys. 95, 4772
(2004).
[20] Y. F. Yang, Y. Z. Jin, H. P. He, Q. L. Wang, Y. Tu, H. M. Lu, and Z. Z. Ye, J. Am. Chem. Soc. 132, 13381 (2010).
[21] S. Durmaz and O. Okay, Polym. Bull. 46, 409 (2001).
[22] X. Qiu, Y. Xue, G. Li, and L. Li, Chem. Lett. 36, 384 (2007).
[23] A. Ohtomo, M. Kawasaki, T. Koida, K. Masubuchi, H. Koinuma, Y. Sakurai, Y. Yoshida, T. Yasuda, and Y. Segawa,
Appl. Phys. Lett. 72, 2466 (1998).
[24] C. H. Ku, H. H. Chiang, and J. J. Wu, Chem. Phys. Lett. 404, 132 (2005).
[25] X. Qiu, L. Li, J. Zheng, J. Liu, X. Sun, and G. Li, J. Phys. Chem. C 112, 12242 (2008).
[26] M. S. Tomara, R. Melgarejoa, P. S. Dobala, and R. S. Katiyara, J. Mater. Res. 16, 903 (2001).
[27] Y. S. Wang, P. J. Thomas, and P. OBrien, J. Phys. Chem. B 110, 21412 (2006).
[28] B. Karthikeyan and T. Pandiyarajan, J. Lumin. 130, 2317 (2010).
[29] A. Wood, M. Giersig, M. Hilgendorff, A. Vilas-Campos, L. M. Liz-Marzan, and P. Mulvaney, Aust. J. Chem. 56, 1051
(2003).
[30] G. H. Ning, X. P. Zhao, and J. Li, Opt. Mater. 27, 1 (2004).
[31] R. Y. Hong, J. Z. Qian, and J. X. Cao, Powder Technol. 163, 160 (2006).
[32] S. C. Liufu, H. N. Xiao, and Y. P. Li, Powder Technol. 145, 20 (2004).
[33] S. J. Chung, J. P. Leonard, I. Nettleship, J. K. Lee, Y. Soong, D. V. Martello, and M. K. Chyu, Powder Technol. 194, 75
(2009).


C 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

You might also like