Materials Science in Semiconductor Processing: A. Gaber, A.Y. Abdel-Latief, M.A. Abdel-Rahim, Mahmoud N. Abdel-Salam
Materials Science in Semiconductor Processing: A. Gaber, A.Y. Abdel-Latief, M.A. Abdel-Rahim, Mahmoud N. Abdel-Salam
Materials Science in Semiconductor Processing: A. Gaber, A.Y. Abdel-Latief, M.A. Abdel-Rahim, Mahmoud N. Abdel-Salam
a r t i c l e in f o abstract
Available online 24 July 2013 Tin dioxide (SnO2) nanoparticles were synthesized by a conventional precipitation
Keywords: method using the reaction between tin chloride pentahydrate and ammonia solutions.
SnO2 nanocrystalline The obtained powders were calcined at varied temperatures from 300 to 1050 1C, and
Precipitation then characterized by using thermogravimetric analysis, differential thermal analysis and
XRD Fourier transformation infrared spectroscopy. The average crystallite size, determined by
Optical band gap x-ray diffraction, was found to be in the range of 3.45–23.5 nm. The analysis exhibited a
tetragonal phase. The activation energy of crystal growth was calculated and found to be
12.12 kJ/mol. The microstructure of nanoparticles was examined by high resolution
transmission electron microscopy. Optical properties were investigated by a UV–vis
absorption spectrophotometer. The calculated optical band gap lies between 4.75–4.25 eV
as a result of increasing the calcination temperatures and crystallite size.
& 2013 Published by Elsevier Ltd.
crystallite size and lattice strain has been reported. The direct shown in Fig. 1. The total weight loss is 16.17%, which
optical band gap has been determined. The novelty of this occurred in two stages: I and II. A large amount of weight
work is to relate the calcination temperature with the carrier loss (8.76%) appears in the first stage in the temperature
density, crystallite size, dislocation density and microstrain of range between room temperature and 149 1C which is
SnO2 nanoparticles. attributed to dehydration of water molecules (physically
adsorbed water) on the surface of particles. This process
2. Experimental procedure was confirmed by an endo-thermic peak at around 72 1C in
DTA thermogram. The second stage, taking place from 150 1C
2.1. SnO2 sample preparation to 400 1C, is related to the elimination of hydroxyl group and
decomposition of ammonia. This process was described by
Tin (IV) chloride pentahydrate SnCl4 5H2O 98% [Stream an exo-thermic peak centered at 241 1C in DTA. The results
Chemicals, USA] and ammonia solution NH3 H2O 30% agree with those of Song et al. [12] and are consistent with
[Loba Chemicals, India] were used as starting materials. FTIR analysis of the samples, which indicates the disappear-
Chemicals used in this study were analytically pure and ance of ammonia at temperatures close to 300 1C. The
used without further purification. weight loss in this stage is 5.07%. After 400 1C the weight
Nanocrystalline tin dioxide (SnO2) was prepared by add- loss occurred by the elimination of the little amount of
ing NH3 H2O (30%) dropwise to 0.05 M aqueous solution of residual hydroxyl group during the development of tin
SnCl4 5H2O at 40 1C. The pH of the mixed solution was dioxide nanoparticles [14]. These results are in agreement
adjusted at 8 to obtain a white precipitate of tin hydroxide. with XRD analysis, which determines that the particles have
The product was separated using a centrifuge operating at good crystallization at 600 1C. The results indicate that the
3000 rpm for 20 min. The white gel was washed with water molecules and hydroxyl group inhibit the growth of
deionized water and ethanol several times to remove chlor- particles, until the temperature reaches 600 1C which shows
ine ions. The resulting product was dried for 24 h at 100 1C in rapid increase in the growth of crystallite size [3]. After that
a drying oven, and then ground in a mortar. Finally, the the precipitated phases do not cause any thermal or weight
powder was calcined at different temperatures in a muffle changes.
furnace for 2 h in the range of 300–1050 1C.
72
optical absorption and the direct optical band gap was
determined by analyzing the obtained data. II
110
101
211
T=900°C
200
T=1050°C
count. (a.u)
T=600°C
617
T=900°C
T=750°C
T=600°C
T % (a.u)
T=450°C
609.7
T=300°C
T=100°C
T=300°C
20 30 40 50 60 70 80 90 100
T=100°C
598.5
2 (deg.)
1400.7
1635.4
25
3385.3
556.8
20
5
bending vibration of CH2 [15]. The FTIR spectrum confirms
that a few organic groups, hydroxyl groups and water
molecules still remain on the surface although the samples 150 300 450 600 750 900 1050
were calcined at high temperatures [20]. Temperature (°C)
Fig. 3. (a) The XRD charts of the calcined samples at different tempera-
3.3. XRD structural analyses of tin oxide nanoparticles tures, and (b) dependence of the crystallites growth on the calcination
temperature.
Fig. 3a shows the XRD charts of dried sample and
samples calcined in the range of 300–1050 1C. The The dependence of crystallite size on the calcination
obtained peaks indicate a rutile tetragonal phase of SnO2 temperatures is shown in Fig. 3b. The decline in the values
nanocrystalline with lattice parameters a ¼4.742 Å and of crystallite size in the temperatures range below 600 1C
c ¼3.187 Å [21]. The average crystallite size is related to is related to the presence of adsorbed water and hydroxyl
the peak broadening. Therefore, crystallite size was depen- groups which was confirmed by TGA and DTA results,
dent on the calcination temperatures and increases with which in turn leads to prevention of the growth of crystal-
increase in the calcination temperatures. The sharpness lites rapidly. Then the crystallite growth increases with
and numbers of reflection increase with increase in the increase in the calcination temperatures which can be
calcination temperatures, due to the development of ascribed by more agglomeration between crystals.
crystallinity. Miller indices are provided in the figure, and The activation energy of crystal growth of nanoparticles
all peaks refer to SnO2 nanoparticles with no detection of (E) can be calculated by Scott’s equation (Eq. (2)), which
other phases or materials; the planes (110), (101) and (211) describes the process of nanoparticles growth under the
have the highest intensity. The average crystallite size of assumption of homogeneous growth [22]:
tin dioxide nanocrystals was estimated by the modified E
Scherrer equation [17]: D ¼ A expð Þ ð2Þ
RT
kλ
D¼ ð1Þ where D is the crystallite size, A is constant, R is the ideal
β cos θ
gas constant and T is the absolute temperature. A plot of ln
where β is defined as β¼(β2observed β2instrumental)1/2, k is the (D) versus 1/T for the samples calcined from 600 1C is
shape factor equal to 0.89, λ is the X-ray wave length for Cu shown in Fig. 4; from the slope of the straight line the
kα radiation (λ¼1.5418 Å), θ is the Bragg diffraction angle value of the activation energy of crystal growth is equal to
(in degrees) and β (in radians) is the full width at half 12.12 kJ/mol. The value of activation energy implies that
maximum (FWHM) of the observed peak which is cor- the chemical driving force of nanoparticles growth is low
rected by considering the instrumental broadening of the and the kinetics of growth is controlled by interfacial
standard Si. reaction [23]. This result is fairly consistent with the
A. Gaber et al. / Materials Science in Semiconductor Processing 16 (2013) 1784–1790 1787
β cos θ / λ
on the line broadening according to the Williamson–Hall
equation [25].
β cos θ 1 sin θ 0.05 T=900°C
¼ þ m ð3Þ
λ D λ
Fig. 5 represents the slope of (β cos θ)/λ versus (sin θ)/λ. T=1050°C
0.04
Lattice strain is determined from the slope of the straight
line, while the crystallite size is calculated from the
intercept of the Y-axis. The values of m and D are estimated 0.03
and listed in Table 1. The lattice strain decreases with 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0
increase in the calcination temperature from 600 1C to sinθ / λ
1050 1C, due to the relaxation of particles, whereas the
Fig. 5. Williamson–Hall plot of SnO2 nanocrystals at different calcining
crystallite size increases. From these results one can temperatures.
conclude that the lattice-strain of the crystallite is calcina-
tion temperature dependent. The crystallite sizes calcu- Table 1
lated from the Williamson–Hall equation are slightly less The mean crystallite size that is calculated by the modified Scherrer
than those calculated by the modified Scherrer equation. equation and Williamson–Hall equation and lattice strain for samples
that are calcined at different temperatures.
3.4. HRTEM analysis for SnO2 nanoparticles Temperature Crystallite size by Crystallite size by Strain
(1C) Scherrer (nm) Williamson–Hall (nm) (ε, %)
The micro-structure of the SnO2 samples calcined at
600 13.3 11.24 0.98
450 1C, 600 1C, and 750 1C investigated by the HRTEM in
750 14.4 12.05 0.96
Fig. 6 shows that the nanoparticles appear as spherical 900 18.4 16.13 0.62
particles with uniform size as the calcination temperature 1050 23.54 19.6 0.5
increases. The average particle size increases as a function
of the calcination temperature and the average diameter is
equal to 6.07 nm, 13 nm and 21 nm for T ¼450 1C, 600 1C agglomerated particles have regular distribution and the
and 750 1C, respectively. The HRTEM results are in fair shapes of particles are, to some extent, similar and
agreement with the crystallite sizes determined by XRD homogeneous.
line broadening. The particles have a high degree of The selected area electron diffraction (SAED) pattern
agglomeration caused by mutual influence between parti- images of tin dioxide nanocrystals calcined at different
cles which arises from some forces such as van der Walls temperatures are characterized by diffraction rings with
forces, capillary forces and electrostatic forces [26–27]. discrete spots. SAED indicates full crystallinity of tin dioxide
Furthermore, the presence of water molecules and hydro- nanoparticles with polycrystalline nature and is in good
xyl groups in powder calcined at high temperature par- agreement with the XRD patterns. The lattice spacing is equal
tially affects the agglomeration of particles. The to 0.28, 0.31 and 0.26 nm corresponding to the (101), (110)
and (101) planes of tetragonal phase of tin oxide nanoparticles
3.2 and is in agreement with XRD data [21]. The interplanar
spacing does not depend on the calcination temperature but it
depends on the orientation of imaged crystals.
3.0
3.5. Optical properties of SnO2 nanoparticles
lnD (nm)
0.28 nm
0.31 nm
0.26 nm 0.31
0.26 nm
Fig. 6. HRTEM images and SAED of SnO2 nanocrystals calcined at different temperatures: (a, b) T ¼450 1C, (c, d) T ¼ 600 1C, AND (e, f) T ¼ 750 1C.
where A is the absorbance, ρ is the theoretical density of The behavior of α versus hυ can be divided into two
SnO2 (6.9 g cm 1) [29], l is the optical path length of a regions: the exponential region and high absorption
quartz cell (1 cm), c is the molar concentration and M is region. The density of localized tail state Ee in the for-
the molecular weight of SnO2 powders. Fig. 7 describes bidden energy gap can be determined in the exponential
typical optical absorption spectra of samples A2, A3, A4 region using Urbach's equation (Eq. (5)) by plotting lnα vs.
and A6 calcined at 300 1C, 450 1C, 600 1C and 900 1C, hυ as shown in Fig. 8 [31].
respectively. The sharp spectrum with high absorbance is
α ¼ αo ehυ=Ee ð5Þ
related to a stable nanosuspension solution containing
highly crystallized SnO2 nanoparticles. Furthermore, the where αo is a constant and hυ is the photon energy.
absorption edges for all samples are shifted to shorter The optical band gap Eg in a semiconductor is deter-
wave length (blue-shift) which indicates higher optical mined by plotting (αhυ)1/n vs. (hυ) using Eq. (6) [29].
band gap [30]. Assuming the nature of the transition n has different
A. Gaber et al. / Materials Science in Semiconductor Processing 16 (2013) 1784–1790 1789
1.4
A2, T=300 °C E
1.3 4.8
A3, T=450 °C E 0.6
1.2 A4, T=600 °C
4.7
A6, T=900 °C
1.1
4.6
1.0 0.5
Ee (eV)
Eg (eV)
0.9 4.5
A (a. u)
dir
0.8 4.4
0.4
0.7
4.3
0.6
0.5 4.2 0.3
0.4 4.1
300 400 500 600 700 800 900
0.3
100 200 300 400 500 600 700 800 900 1000 Temperature (°°C)
λ (nm)
Fig. 7. UV–vis absorption spectra for all SnO2 samples calcined at indicated 4.8 0.6
temperatures.
4.7
16.4 0.5
4.6
Ee (eV)
Eg (eV)
16.2
A2, T=300°C 4.5 0.4
dir
A3, T=450°C
A4, T=600°C
16.0 4.4
A6, T=900°C
0.3
ln (α )
4.3 E
15.8 E
4.2 0.2
0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0
15.6
Crystallite size (nm)
Fig. 10. Dependence of (a) direct energy gap and (b) localized tail state on
15.4 the calcination temperature and crystallite size of SnO2 nanoparticles.
1.40 1.45 1.50 1.55 1.60 1.65 1.70
hν (eV)
where B is the absorption constant that depends on
Fig. 8. ln(α) as a function of photon energy (hυ) for all SnO2 samples transition probability. The direct optical band gap energy
calcined at indicated temperatures.
of SnO2 nanoparticles is calculated in the high energy
region of the absorption edge by plotting (αhυ)2 vs. (hυ)
8.0x1015 and the values of Eg are obtained from the extrapolation of
A2,T=300°C
the linear portion of the curve (αhυ)2 ¼ 0 (Fig. 9).
A3,T=450°C
Fig. 10a, b gives the values of optical band gaps and the
15
A4,T=600°C
6.0x10 density of localized tail states of SnO2 nanoparticles.
A6,T=900°C
(α h ν)2 (cm-1 eV)2
Table 2
Variations of carrier concentration, dislocation density and optical band gap of SnO2 nanoparticles samples at different calcination temperatures.
Samples no. Temperature (1C) Eg (eV) Ee (eV) ΔEg (eV) Carrier concentration (N2/3) 1016 Dislocation density (δ 1016 ) lines/m2