Você está na página 1de 9

Copyright © 2018 American Scientific Publishers Nanoscience and

All rights reserved Nanotechnology Letters


Printed in the United States of America Vol. 10, 1–9, 2018

Doping Dependent Structural, Optical, Thermal and


Catalysis Properties of Synthesized Cadmium
Sulfide Nanoparticles
W. T. Salam1 † , M. Ikram1 ∗ † , I. Shahzadi2 , M. Imran3 4 , M. Junaid1 , M. Aqeel1 , S. Anjum5 ,
A. Shahzadi2 , H. Afzal1 , U. Sattar1 , A. Asghar1 , A. Wahab1 , M. Naz6 , M. Nafees1 , and S. Ali1 7
1
Solar Cell Applications Research Lab, Department of Physics, Government College University Lahore, 54000, Pakistan
2
Punjab University College of Pharmacy, Punjab University, Lahore, 54000, Pakistan
3
Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, 29 Zhongguancun East Road,
Haidian District, Beijing 100190, China
4
University of Chinese Academy of Sciences, Beijing 100049, China
5
Department of Physics, COMSATS University Islamabad, Lahore Campus, Lahore 54000, Pakistan
6
Biochemistry Lab, Department of Chemistry, Government College University Lahore, 54000, Punjab, Pakistan
7
Department of Physics, Riphah Institute of Computing and Applied Sciences (RICAS), Riphah International University,
14 Ali Road, Lahore, Pakistan

A simple and cost-effective route has been employed to synthesize undoped and copper doped
(2, 4 and 6) % cadmium sulfide (CdS) nanoparticles. Resulting materials were characterized by
various techniques as XRD results confirmed the formation of hexagonal crystal structure for doped
and undoped CdS nanoparticles with crystallite size varying from 11–17 nm calculated by Sherrer
equation. The Williamson Hall (W-H) plot is used to examine particle size and lattice strain of pre-
pared pure and Cu doped CdS nanoparticles and compared the particle size with Sherrer equation.
Moreover, significantly reduced the intensity of diffraction is also evident accompanied by slight
peak-shifting towards higher angles. Agglomerated spherical morphology of the CdS nanoparticles
with decreasing particle size upon Cu doping was observed by SEM. FTIR spectra was found
to consist of stretching vibrations from Cd-S and other related groups with negligible changes in
the vibrational modes upon doping. The UV-Vis analysis demonstrates a significant increase in
absorption intensity upon doping along with slight blue shift leading to an increased bandgap for
doped CdS. The Raman spectroscopy measurement was performed on nanoparticles and thermal
properties in different environment were investigated by DSC/TGA. These studies represent a cost-
effective way to prepare CdS and related materials with tunable bandgap. This report highlights the
effect of bared and various concentrations of Cu doped CdS nanoparticles in the degradation of
cationic dye methylene blue by NaBH4 .
Keywords: CdS Nanoparticles, Copper, FESEM, FTIR, Chemical Precipitation Techniques,
Methylene Blue.

1. INTRODUCTION in bulk form. It is widely used in applications such as het-


It has been commonly observed that the CdS forma- erojunction solar cells, laser materials, optoelectronic and
tion as p-type is very challenging mostly due to self- photoconductive devices, light emitting diodes, and optical
compensation effects because of sulfur vacancies.1 Many detectors.3–5 CdS shows n-type conductivity to the intrinsic
research groups introduced dopants to alter the properties defects present in it while substitutional doping tends to
of the inorganic semiconductor to enhance the efficiency generate an acceptor level thus changing its conductivity
of photovoltaic devices.2 Cadmium sulfide is a group II–VI from n-type to p-type.6 7
semiconductor material with a bandgap (B.G) of 2.42 eV In thin-film solar cells, n-type CdS is used as a win-
dow material4 while in bulk-hetero-junction (BHJ) solar

Author to whom correspondence should be addressed. cells CdS nanoparticles are incorporated as electron accep-

These two authors contributed equally to this work. tor materials in the active layer along with electron donor

Nanosci. Nanotechnol. Lett. 2018, Vol. 10, No. xx 1941-4900/2018/10/001/009 doi:10.1166/nnl.2018.2839 1


Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles Salam et al.

polymers.3 In thin-film solar cells, use of CdS as n-type 2.2. Synthesis


window layer comes with certain disadvantages such as The copper doped CdS nanoparticles were synthesized by
small band gap and lattice mismatch with p-type thin the precipitation method using ethylene glycol as a solvent
films.5 While in BHJ solar cells, an efficient device oper- and capping agent. Solution (I) was obtained by mixing
ation is determined by the generation of photo-induced 0.02 mol of CdCl2 · 2H2 O and an equimolar amount of
charge carriers at the organic-inorganic interface which C2 H5 NS into 120 ml of ethylene glycol. Different weight
in-turn depends on the difference between the bandgap percentages of CuCl2 · 2H2 O were dissolved in ethylene
energies of organic and inorganic components. A large glycol and mixed slowly in the solution (I). The resulting
difference in the bandgap energies would result in short- solution was stirred at 100  C on a hotplate for four hours
diffusion lengths for charge carriers and corresponding until the appearance of brown precipitates. These precip-
degraded power conversion efficiency.8–18 itates were filtered, washed with ethanol and deionized
Therefore, thin films and BHJ devices can be made water, and dried at 100  C. Similar procedure was used to
comparatively more efficient by tunning the CdS bandgap. fabricate pure CdS nanoparticles without the addition of
These requirements can be achieved by the addition of CuCl2 · 2H2 O in the solution (I)
impurities such as Cu, Ag and Fe into the pristine CdS.
In past, various techniques such as vacuum deposition,19 2.3. Catalytic Reduction of Methylene Blue
vacuum co-evaporation,6 spray-pyrolysis,20 mechanochem-
Sufficient deionized water was employed to dilute the
ical alloying and dry process21 have been employed to pro-
300 l of methylene blue solution (1 mM) in quartz cell.
duce Cu doped CdS. However, these techniques employ
Then, added 400 l of NaBH4 solution (0.1 M) and 300 l
expensive equipment and generation of high vacuum can
of doped and undoped CdS nanoparticles solution to make
take several hours.
a final volume of 3.00 ml. The decolourisation of the dye
Organic dye effluents are widely used in paper, plas-
indicate methylene blue degradation because of reducing
tic, textile and food industries, though the hazardous
agent (NaBH4 . In addition, degradation of the MB at
effects of these coloured materials are the major threat
664 nm was recorded spectrophotometrically. The reaction
to the humans and aquatic organisms. Hence it is
in the absence of nano-catalyst is conducted as a control
essential to remove these toxic materials from waste
and absorption spectra were monitored at periodic inter-
water resources.22 23 Traditional treatment methods include
ultrafiltration, adsorption, biodegradation, photochemical vals in the range of 250–750 nm at 25  C.
chemical, and electrochemical process24–28 are inappropri-
ate and insufficient to remove huge amounts of organic 2.4. Characterization
dyes from industrial effluents; such as adsorption is a The structural analysis and phase identification of the pre-
non-destructive technique, it simply convert these coloured pared samples were performed using the instrument PAN-
materials from one phase to another. Nowadays, metal alytical X’Pert PRO XRD Company Ltd. Holland operated
nanoparticles emerged as an effective nano-catalyst that at 40 kV with CuK radiation of wavelength 1.54 nm.
rapidly replacing conventional water treatment method due The surface morphology was analyzed by using FESEM
to their finite size and large surface to volume.29 JSM-5910 with the accelerating voltage of 20 kV. The
In this study, we report the synthesis of doped and FTIR spectra of CdS and Cu doped CdS nanoparticles
undoped CdS nanoparticles using chemical precipitation were recorded on Bruker TENSOR II in the range 600–
technique. Structural, morphological, and optical proper- 4000 cm−1 and the UV visible spectroscopy was carried
ties were evaluated as a function of doping. The catalytic out on Genesys 10S UV-Vis spectrophotometer for optical
activity of undoped and Cu doped CdS nanoparticles were analysis.
investigated by using them in the degradation reactions of
methylene blue by NaBH4 .
3. RESULTS AND DISCUSSION
Figure 1(a) represents powder X-ray diffraction (XRD)
2. MATERIALS AND METHODS patterns of pure and Cu-doped CdS nanoparticles. The
2.1. Materials XRD pattern of as-prepared CdS nanoparticles was fully
Following materials were used for the synthesis of matched with JCPDS card number 01-075-1545. The
doped and undoped CdS nanoparticles: cadmium chlo- obtained diffraction peaks from the planes (102) and (103)
ride (CdCl2 · 2H2 O) as the cadmium source, thioacetamide are the characteristic peaks of hexagonal close packed
(C2 H5 NS) for the sulfur source, copper chloride (CuCl2 · (hcp) structure.30 It can be observed that doping signifi-
2H2 O) as dopant metal salt and ethylene glycol (C2 H6 O2  cantly reduces the intensity of diffraction peaks resulting in
as a solvent. Methylene blue (MB) and sodium borohy- a decreased crystallite size (Fig. 1(b)). Crystallite size for
dride (NaBH4  were procured from Sigma-Aldrich. All doped and undoped samples is shown in Figure 1(c). The
materials were of analytical reagent grade and purchased doping effect is also manifested in the slight shifting of
from Merck. XRD peaks towards higher angles which can be attributed

2 Nanosci. Nanotechnol. Lett. 10, 1–9, 2018


Salam et al. Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles

Fig. 1. XRD patterns of (a) W-H plot (b) and comparison between particle size and lattice strain (c) and crystallite size dependence on the concentration
of Cu in CdS (d).

to the difference in the atomic radii of Cu (1.35 Å) and measured crystallite size with W-H plot and Scherer equa-
Cd (1.51 Å) atoms. Copper may have been doped substi- tion is shown Figure 1(d). With higher amount of dop-
tutionally into the CdS causing lattice contraction.7 ing (6%), contribution of peak broadening due to lattice
In order to check the broadening of the peak with lattice strain is much prominent relative to particle effect broad-
strains, different peaks were utilized from XRD pattern. ening which enhanced the induced lattice strain as well
The strain induced broadening of the Bragg’s diffraction as crystallite size (W-H) Figure 1(c). From Figure 1(d),
peak can be calculated using Wilson and Stokes31 formula the crystallite size decreased gradually according to Sher-
Eq. (1). rer equation because it does not consider the broadening
 effect due to lattice strain.
 = hkl (1)
4 tan  Field emission scanning electron microscopy (FESEM)
The broadening is contribution of lattice strain induced and was used to analyze the surface morphology of undoped
crystallite size and can be written as and Cu doped CdS nanoparticles (Fig. 2). Figure 2(a)
reveals the formation of highly agglomerated spherical
hkl = D +  (2) nanoparticles of pure CdS in the range of 0.2–1.5 m.

Using Scherer equation and (1) results in following


Eq. (3)
K
hkl = + 4 tan  (3)
DCos
Rearranging the Eq. (3) gives
K
hkl cos  = + 4 sin  (4)
D
The W-H plot of hkl cos  versus 4 sin  for pure and
doped CdS samples with concentration (0, 2, 4 and 6) %
are shown in Figure 1(b). The extracted values of crystal-
lite size (inverse of Y-intercept) and strain (slope of the
line) induced in the broadening of the peak are shown
in Figure 1(c). It can be seen from Figure 1(b); the lat-
tice strain values are fluctuating due to the increasing Fig. 2. SEM images of (a) pure CdS, (b) 2% Cu:CdS, (c) 4% Cu:CdS
amount of Cu substitution in Cd site. The comparison of and (d) 6% Cu:CdS.

Nanosci. Nanotechnol. Lett. 10, 1–9, 2018 3


Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles Salam et al.

Whereas Figures 2(b)–(d) displaying particles size distri- peak. The broadening and intensity of the peaks increased
bution in the range of 0.05–1.65 m, 0.1–1.25 m and with increasing amount of Cu suggesting compositional
0.1–1 m respectively. Increasing concentration of Cu disorder.
enhances the agglomeration while decreasing the particles Figure 5(a) indicates reducing capacity of NaBH4 with
size. This is in agreement with XRD results which reveal MB after 40 mins, the addition of undoped CdS with
a reduced intensity of Bragg reflections and smaller crys- NaBH4 representing continuous decrease in peak intensity
tallite size for doped CdS nanoparticles. of MB, Almost 95% reduction of MB was observed in
Optical properties of pure and Cu doped CdS nanopar- less than 15 min, indicating the effectiveness of CdS as
ticles were probed by UV-Vis spectroscopy (UV-Vis) as nano-catalystn Figure 5(b). 2% Cu doped CdS with NaBH4
shown in Figure 3(a). Pristine CdS nanoparticles demon- reduce the dye within 25 min measured by spectropho-
strate strong absorption in UV region. Doping enhances tometer at the regular interval of approximately 6 min
the absorption intensity accompanied by a blue shift32 (Fig. 5(c)). 4% and 6% Cu doped CdS showing succes-
(Fig. 3(a)). This slight shifting in absorption edge is the sive decrease in the concentration of MB within 40 min at
consequence of the quantum confinement effect produced the peak intensity of ≈664 nm Figures 5(d)–(e). Catalytic
due to increased nucleation rate with doping.33 To mea- activity proposing that NaBH4 is not actively participate
sure the energy band gap from the absorption spectra for the reduction of methylene blue whereas undoped CdS
for undoped and doped CdS nanoparticles was calculated representing significant catalytic potential as compare to
using Tauc’s equation as shown in Figure 3(b). Cu doped CdS nanoparticles. The decreased catalytic effi-
Figure 3(c) shows the effect of lattice strain and crys- ciency of Cu doped CdS nanoparticles might be attributed
tallite size on band gap of undoped and Cu doped CdS. due to increased agglomeration between particles. In addi-
The lattice strain (Fig. 3(c)) and crystallite size (Fig. 3(d)) tion, the reduction process was found to be accelerated in
decreased gradually with the addition of Cu and this trend the presence of CdS NPs which exhibits rapid decrease
reverse with higher amount of doping.34 35 Figures 3(c)–(d) in absorption intensity of MB solution. CdS NPs sup-
show that the increase in band gap is proportional to dop- ports in electron relay from BH4 (donor) to MB (acceptor
ing concentrations and inversely proportional to crystal- material). BH4 ions are nucleophilic in nature while MB
lite size and strain. As the ionic radius of Cu+2 (0.73 Å) is electrophilic in nature. Where the CdS accept electron
is smaller than the radius of Cd+2 (0.95 Å). Cu+2 may from BH4 ions and convey them to MB. By the addition of
occupy Cd+2 sites producing compressive strain that con- Cu, the acceptance rate of doped NPs is decreased as Cu
sequently reduces the particle size of Cu doped Cds NPs. (1.90) is more electronegative relative to Cd (1.69) restrict
When the size of the CdS nanocrystals becomes smaller the charge mobility of electrons leads to decrease the cat-
than the exciton radius, a remarkable quantum size effect alytic activity. The degradation phenomena of MB in the
leads to a size-dependent increase in the band-gap and a presence of reducing agent with undoped and doped CdS
blue shift in the absorption onset.36 NPs are represented in Eq. (5).
The FTIR spectra of CdS and Cu doped CdS nanoparti-
cles were recorded in the range 400–4000 cm−1 as shown
in Figure 4(a). It can be observed that peaks appear at
same positions for doped and undoped samples reflecting
minimal changes in crystal structures upon doping. Peak
appearing at 670–690 cm−1 is assigned to Cd–S stretch-
ing vibrations37 while 1638 cm−1 is the bending of N–H
originating from the thioacetamide compound.38 The peak
observed at 2361 cm−1 corresponds to S–H stretching
from the hydrogen sulfide gas (H2 S) produced during the
synthesis.39 The peak at 3198 cm−1 is the O–H stretching
that may interact with ethanol through the hydrogen bond-
ing which becomes more prominent in 0.02 % Cu doped
as it has more moisture (8% as described below in DSC- (5)
TGA graph) than other while the other O–H bond attaches Methylene blue reduced to leucomethylene blue in the
itself on the surface of the CdS shown at 3466 cm−1 .40 presence of NaBH4 and nano catalyst (undoped and Cu
The Raman spectra of prepared nanoparticles are shown doped CdS) at room temperature as shown in above
in Figure 4(b). The observed peaks for CdS around 295 equation.42 43 The presence of NaBH4 in solution increases
and 589 cm−1 are attributing to 1st and 2nd order scat- the pH of system; which ultimately retard the degradation
tering of LO-longitudinal optical phonon mode respec- of BH− +
4 ions, liberating H ions that reduce MB.
42 43
Thus
tively and well matched with reported.41 The 1st order the metal nanoparticles are efficient catalysts, suitable for
peak (295 cm−1  is asymmetric appears like the superpo- removing MB from industrial effluents and can be produc-
sition of modes (cubic 1LO or hexagonal A1(LO)/E1(LO) tively used in waste water treatment.

4 Nanosci. Nanotechnol. Lett. 10, 1–9, 2018


Salam et al. Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles

Fig. 3. UV visible spectra (a) and Tauc plots (b), graph between bandgap and lattice strain with Cu concentration (c) and crystallite size versus band
gap for pure and doped CdS (d).

The DSC-TGA analysis of prepared undoped and doped loss (8%–30%, Table I) recognized by TGA/DTGA curves
2, 4, 6 % doped CdS were performed in different envi- for all graphs, clearly visible in the temperature range of
ronments (in air, N2 gas) keeping ramp rate of 5  C/min about 550  C to 650  C. This endothermic transition as
as shown in Figures 6(a)–(d) and (e)–(h) respectively. The
vet by DSC/DDSC curves are owing to expected subli-
exothermal peaks are deliberated in up direction. Anal-
ysis of TGA curves for doped and pure CdS in both mation of ultra-fine nanoparticles of doped and pure CdS
environmental conditions show a mass loss (upto 2%–8%, presence in the samples. These samples treated in N2 envi-
Table I) starting from about 150  C to 260  C, most ronment suffer a final and complete mass loss after 900  C
probably due to remove of water of crystallization and as demonstrated by TGA and DTGA curves; this endother-
organic material contents used for washing purpose dur-
mic sublimation of remaining particle with comparatively
ing the synthesis process.44 45 The corresponding DTGA
peaks appear after 225  C moreover DSC/DDSC curves big particle size is correspondingly supported by respective
10(a)–(d) also confirm the endothermic nature of transi- DSC and DDSC plots. Doped and pure CdS nanoparticles
tion in same temperature range. Another common mass sublime at relatively low temperature in control and inert

Fig. 4. FTIR spectra (a) and Raman spectroscopy of various ratios of Cu in CdS nanoparticles (b).

Nanosci. Nanotechnol. Lett. 10, 1–9, 2018 5


Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles Salam et al.

Fig. 5. (a–e) time dependent UV–vis spectra for the reduction of MB, MB with NaBH4 (a), MB with NaBH4 + CdS (b), MB with NaBH4 + 2% Cu
doped CdS (c), MB with NaBH4 + 4% Cu doped CdS (d) and MB with NaBH4 + 6% Cu doped CdS (e).

environment (N2 Gas) as bulk CdS has higher sublimation confirm the endothermic nature of transition in same tem-
point 980–1000  C. perature range. Another common mass loss (8%–30%,
The DSC-TGA analysis of prepared undoped and doped Table I) recognized by TGA/DTGA curves for all graphs,
2, 4, 6 % doped CdS were performed in different envi- clearly visible in the temperature range of about 550  C to
ronments (in air, N2 gas) keeping ramp rate of 5  C/min 650  C. This endothermic transition as vet by DSC/DDSC
as shown in Figures 6(a)–(d) and (e)–(h) respectively. The curves are owing to expected sublimation of ultra-fine
exothermal peaks are deliberated in up direction. Analysis nanoparticles of doped and pure CdS presence in the sam-
of TGA curves for doped and pure CdS in both environ- ples. These samples treated in N2 environment suffer a
mental conditions show a mass loss (upto 2%–8%, Table I) final and complete mass loss after 900  C as demonstrated
starting from about 150  C to 260  C, most probably due by TGA and DTGA curves; this endothermic sublima-
to remove of water of crystallization and organic mate- tion of remaining particle with comparatively big particle
rial contents used for washing purpose during the synthe- size is correspondingly supported by respective DSC and
sis process.44 45 The corresponding DTGA peaks appear DDSC plots. Doped and pure CdS nanoparticles sublime at
after 225  C moreover DSC/DDSC curves 10(a)–(d) also relatively low temperature in control and inert environment

6 Nanosci. Nanotechnol. Lett. 10, 1–9, 2018


Salam et al. Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles

Fig. 6. DSC/TGA graphs of CdS (a–b) and Cu doped CdS 0.02, 0.04 and 0.06% (in N2 and air) (c–h) respectively.

(N2 Gas) as bulk CdS has higher sublimation point Figures 6(e)–(h). This behavior reveals the exothermic
980–1000  C. nature of this transition but corresponding TGA and
Doped and undoped CdS demonstrate completely dif- DTGA plot show very fascinating and curious behavior.
ferent thermal behavior in air after 650  C around 750  C It is noteworthy that the control CdS and 2% Cu doped
to 820  C can be seen in DSC and DDSC curves CdS suffered a weight loss while 4% and 6% doped

Nanosci. Nanotechnol. Lett. 10, 1–9, 2018 7


Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles Salam et al.

Table I. DSC/TGA result extracted from Figures 11 and 12. CdS samples possess a mass gain due to formation of sul-
Temperature Mass Thermal fates by thermal oxidation. The undoped CdS nanoparti-
range (sample) Environment Transition variation behavior cles are proved to be remarkably competent catalyst by
150 to 260  C Air and N2 Removal of Mass loss Endothermic
reducing methylene blue in less than 15 min, inspite of this
(all samples) gas water of Cu doped (2–6)% CdS nanoparticles degrade MB within
crystalliza- 25–40 min. This novel approach of CdS nanoparticles gain
tion valuable attractions and offer an efficient and economic
550 to 650  C Air and N2 Sublimation of Mass loss Endothermic
(all samples) gas ultrafine
route to environmental protection.
particles
900 to 1000  C Air Sublimation of Mass loss Endothermic Acknowledgment: Authors would like thank to
(all samples) heavy Higher Education Commission-HEC, Pakistan for finan-
particles
750 to 820  C N2 Oxidation Mass loss Exothermic
cial support through Start Research Grant Project
(pure and along with 21/1669/SRGP/R&D/HEC/2017) and CAS-TWAS Presi-
2% decomposi- dent’s Fellowship for international PhD students China.
Cu–CdS) tion
750 to 820  C N2 Oxidation Mass Exothermic
(4% and 6% increment References and Notes
Cu–CdS) 1. Y. Kashiwaba, I. Kanno, and T. Ikeda, J. Appl. Phys. 31, 1170
After 850  C N2 Thermal Mass loss Endothermic (1992).
(4% and 6% decomposi-
2. M. Grätzel, Nature 414, 338 (2001).
Cu–CdS) tion
3. S. A. Dowland, L. X. Reynolds, A. MacLachlan, U. B. Cappel, and
S. A. Haque, J. Mater. Chem. A 1, 13896 (2013).
4. M. Imran, M. Ikram, A. Shahzadi, S. Dilpazir, H. Khan, I. Shahzadi,
samples possess a mass gain in same temperature range. and S. A. Yousaf, RSC Advances 8, 18051 (2018).
This increment in mass is caused by thermal oxidation of 5. A. B. Wong, S. Brittman, Y. Yu, N. P. Dasgupta, and P. Yang, Nano
Lett. 15, 4096 (2015).
these samples (4% and 6% Cu doped CdS) in air atmo- 6. H. Xie, C. Tian, W. Li, L. Feng, J. Zhang, L. Wu, Y. Cai, Z. Lei,
sphere may yielded the sulfate/oxysulfate of cadmium. and Y. Yang, Appl. Surf. Sci. 257, 1623 (2010).
Further mass loss after 850C in 4% and 6% doped CdS 7. P. Reyes and S. Velumani, Mater. Sci. Eng. B 177, 1452 (2012).
are due to decomposition of cadmium sulfate/ oxysulfate 8. G. H. Lee and Y. S. Kim, J. Nanosci. Nanotechnol. 17, 8192 (2017).
9. Z. K. Yu, W. F. Fu, W. Q. Liu, Z. Q. Zhang, and H. Z. Chen, Chinese
with the release of SO2 gas to form CdO.46 47 Chemical Letters 28, 13 (2017).
10. D. H. Jeon, D. K. Hwang, J. K. Kang, D. Nam, and H. Cheong,
J. Nanosci. Nanotechnol. 17, 8236 (2017).
4. CONCLUSIONS 11. L. Liu, X. M. Yu, B. Zhang, S. X. Meng, and Y. Q. Feng, Chinese
Structural, optical and morphological properties of Chemical Letters 28, 765 (2017).
undoped and Cu doped CdS nanoparticles synthesized by 12. C. Cho, D. Kong, D. Lee, and B. Kim, J. Nanosci. Nanotechnol.
simple chemical precipitation method have been investi- 17, 8418 (2017).
13. H. Kim, J. Jo, G. Lee, M. Shin, and J. C. Lee, J. Nanosci. Nano-
gated. Pure CdS nanoparticles exhibit the hexagonal close- technol. 17, 8425 (2017).
packed structure and maintain this structure after addition 14. Y. Xu, L. S. Qiang, Y. L. Yang, L. G. Wei, and R. Q. Fan, Chin.
of Cu impurities. Doping effects appear in the form of Chem. Lett. 27, 127 (2016).
reduced intensity of XRD peaks and slight shifting towards 15. Y. W and X.Li, Chin. Chem. Lett. 27, 927 (2016).
16. K. P. Kim, D. K. Hwang, D. H. Kim, and S. H. Woo, J. Nanosci.
higher angles. The W-H plot was used to find the lattice
Nanotechnol. 17, 8201 (2017).
strain and particle size of the samples. SEM reveals spher- 17. F. Ullah, H. Chen, and C. Z. Li, Chin. Chem. Lett. 28, 503
ical morphology and high agglomeration for doped and (2017).
undoped CdS nanoparticles with a decrease in particle size 18. G. H. Lee and Y. S. Kim, J. Nanosci. Nanotechnol. 17, 8383 (2017).
upon doping. FTIR spectra consist of stretching vibrations 19. G. Liu, T. Schulmeyer, J. Brötz, A. Klein, and W. Jaegermann, Thin
Solid Films 431, 477 (2003).
from Cd–S and other related functional groups. Undoped 20. H. Afify, I. El Zawawi, and I. Battisha, J. Mater. Sci. Mater. Electron.
CdS demonstrates strong absorption in UV region. Dop- 10, 497 (1999).
ing enhances absorption intensity accompanied by a blue 21. C. Feldman, G. Deutscher, and E. Grünbaum, S. Appl. Surf. Sci.
shift resulting in a significant increase in band gap from 48, 535 (1991).
22. I. A. Alaton and I. A. Balcioglu, J Photochem Photobiol. A Chem.
3.03 eV for undoped to 3.15 eV for 6% Cu doped CdS
141, 247 (2001).
is attributed to compressive strain and decrease in particle 23. M. Faisal, M. A. Tariq, and M. Muneer, Dyes Pigm. 72, 233
size. These findings suggest a cost-effective way to synthe- (2007).
size CdS nanoparticles with tunable bandgap. All doped 24. T. Robinson, G. McMullan, R. Marchant, and P. Nigam, J. Bioresour.
and pure CdS samples illustrate same thermal behavior in Technol. 77, 247 (2001).
25. M. S. Khehra, H. S. Saini, D. K. Sharma, B. S. Chadha, and S. S.
N2 atmosphere but in air, around 750  C to 820  C showed Chimni, Dyes Pigm. 70, 1 (2006).
very interesting behavior, pure and 2% Cu doped CdS 26. C. Wang, A. Yediler, D. Lienert, Z. Wang, and A. Kettrup, Chemo-
samples suffered a weight while 4% and 6% Cu doped sphere 52, 1225 (2003).

8 Nanosci. Nanotechnol. Lett. 10, 1–9, 2018


Salam et al. Doping Dependent Structural, Optical, Thermal and Catalysis Properties of Synthesized Cadmium Sulfide Nanoparticles

27. V. Gupta, R. Jain, A. Mittal, M. Mathur, and S. Sikarwar, J. Colloid 38. Y. Yu, Y. Ding, S. Zuo, and J. Liu, Int. J. Photoenergy 2011
Interface Sci. 309, 464 (2007). (2011).
28. S. H. Lin and C. F. Peng, Water Res. 28, 277 (1994). 39. N. Qutub and S. Sabir, J. Nanosci. Nanotechnol. 8, 111
29. S. K. Ghosh, S. Kundu, M. Mandal, and T. Pal, Langmuir 18, 8756 (2012).
(2002). 40. R. Ningthoujam, N. Gajbhiye, A. Ahmed, S. Umre, and S. Sharma,
30. T. Thongtem, A. Phuruangrat, and S. Thongtem, J. Phys. Chem. J. Nanosci. Nanotechnol. 8, 3059 (2008).
Solids 69, 1346 (2008). 41. M. K. Indana, B. R. Gangapuram, R. Dadigala, R. Bandi, and
31. A. Stokes and A. Wilson, Proceedings of the Physical Society V. Guttena, JAST 7, 19 (2016).
56, 174 (1944). 42. V. Vidhu and D. Philip, Micron 56, 54 (2014).
32. M. Mahdi, Z. Hassan, S. Ng, J. Hassan, and S. M. Bakhori, Thin 43. M. Naz, M. Z. Qureshi, A. Shahbaz, A. Haider, M. Ikram,
Solid Films 520, 3477 (2012). M. Nafees, A. Shahzadi, T. Bashir, S. Ali, A. C. Blackburn,
33. K. S. Rathore, D. Patidar, Y. Janu, N. S. Saxena, and K. Sharma, H. Chen, and A. Tricoli, Nanoscci. Nanotechnol. Lett. 10, 889
Chalcogenide Lett. 6, 105 (2008). (2018).
34. C. L. Li, J. Yuan, B. Y. Han, and W. F. Shuangguan, Int. J. Hydrogen 44. M. Nafees, M. Ikram, and S. Ali, Appl. Nanosci. 7, 399
Energy 36, 4271 (2011). (2017).
35. K. Jayanthi, S. Chawla, H. Chander, and D. Haranath, Cryst. Res. 45. M. Nafees, M. Ikram, and S. Ali, Dig. J. Nanomater. Biostruct.
Technol. 42, 976 (2007). 10, 635 (2015).
36. W. Sang, Y. Qian, J. Min, D. Li, L. Wang, W. Shi, and L. Yinfeng, 46. R. I. Dimitrov and B. S. Boyanov, J. Therm. Anal. Calorim. 61, 181
Solid State Commun. 121, 475 (2002). (2000).
37. A. U. Ubale, K. S. Chipade, M. V. Bhute, P. P. Raut, G. P. Malpe, 47. R. I. Dimitrov, N. Moldovanska, and I. K. Bonev, Thermochim Acta
Y. S. Sakhare, and M. R. Belkhedkar, J. Mater. Chem. 2, 165 (2012). 385, 41 (2002).

Received: 28 June 2018. Accepted: 26 October 2018.

Nanosci. Nanotechnol. Lett. 10, 1–9, 2018 9

Você também pode gostar