Sei sulla pagina 1di 10

Materials Research Bulletin 74 (2016) 387–396

Contents lists available at ScienceDirect

Materials Research Bulletin


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

Microwave-assisted polyol synthesis and characterization of


pvp-capped cds nanoparticles for the photocatalytic degradation of
tartrazine
Maher Darwisha , Ali Mohammadia,b,* , Navid Assia
a
Department of Drug and Food Control, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran
b
Nanotechnology Research Centre, Faculty of Pharmacy, Tehran University of Medical Sciences, Tehran, Iran

A R T I C L E I N F O A B S T R A C T

Article history: Polyvinylpyrrolidone capped cadmium sulfide nanoparticles have been successfully synthesized by a
Received 16 June 2015 facile polyol method with ethylene glycol. Microwave irradiation and calcination were used to control the
Received in revised form 1 November 2015 size and shape of nanoparticles. Characterization with scanning electron microscopy revealed a restricted
Accepted 3 November 2015
nanoparticles growth comparing with the uncapped product, hexagonal phase and 48 nm average
Available online 10 November 2015
particle size were confirmed by X-ray diffraction, and finally mechanism of passivation was suggested
depending on Fourier transform infrared spectra.
Keywords:
The efficiency of nanoparticles was evaluated by the photocatalytic degradation of tartrazine in
A. Chalcogenides
A. Nanostructures
aqueous solution under UVC and visible light irradiation. Complete degradation of the dye was observed
B. Chemical synthesis after 90 min of UVC irradiation under optimized conditions. Kinetic of reaction fitted well to the pseudo-
C. X-ray diffraction first-order kinetic and Langmuir–Hinshelwood models. Furthermore, 85% degradation of the dye in 9 h
D. Catalytic properties under visible light suggests that cadmium sulfide is a promising tool to work under visible light for
environmental remediation.
ã 2015 Published by Elsevier Ltd.

1. Introduction higher surface area-to-volume ratio, and increased band-gab


energy which in turn lead to higher redox potentials [4].
During the last two decades, photocatalytic materials received a Cadmium sulfide (CdS), a typical metal chalcogenide semicon-
large quantity of research interest as they have a great potential to ductor with a direct band gap Eg  2.43 eV at room temperature [5],
be applied in detoxification of environmental organic pollutants has become one of the most considerable materials in research
and as a clean energy source [1]. Semiconductor heterogeneous communities due to its diverse promising applications in the field
photocatalysts such as TiO2, ZnO, SnO2, WO3, Fe2O3, and CdS have of solar cells, photoelectronic devices and photocatalysis [6]. This
been intensively investigated in the area of water and air material has shown better catalytic functions compared to those of
purification and in remediation reactions [2]. TiO2 due to the rapid generation of electron–hole pairs by
Like other advanced oxidation procedures (AOPs), the common photoexcitation [7]. Nevertheless, CdS has the fatal photocorrosion
characteristic of photocatalytic materials is the generation of very problem due to the self-oxidation by the generated hole [8]. To
reactive species such as, principally but not exclusively, hydroxyl overcome such a problem, an effective approach is to cover the
radicals (HO), which initiate a series of reactions leading nanoparticle core surface with a polymeric capping agent which
eventually to the destruction of the target pollutant [3]. Recently, might also help to stabilize the surface, control the growth, and
nanostructures of these photocatalysts have attracted more prevent agglomeration of the nanoparticles [9–11].
consideration as they are expected to have higher photocatalytic Many methods have been utilized for the synthesis of CdS
activity than their bulk counterparts due to their smaller size, nanostructures such as co-precipitation [7,12], polyol [13],
solvothermal [14,15], hydrothermal [16], non-aqueous chemical
method [17], and chemical bath deposition [18]. Among these
different processes, the polyol method appears as an easy to carry
* Corresponding author art: Department of Drug & Food Control and
out with many other advantages: a uniform shape, a narrow size
Nanotechnology Research Centre, Faculty of Pharmacy, Tehran University of
Medical Sciences, P.O. Box 14155-6451, Tehran, Iran. Fax: +98 21 88358801. distribution, and a low degree of agglomeration [19]. This method
E-mail addresses: m-darwish@razi.tums.ac.ir (M. Darwish), generally uses poly-alcohol like ethylene glycol (EG), diethylene
alimohammadi@tums.ac.ir (A. Mohammadi), navid_a30@yahoo.com (N. Assi).

http://dx.doi.org/10.1016/j.materresbull.2015.11.002
0025-5408/ ã 2015 Published by Elsevier Ltd.
388 M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396

glycol (DEG) or 1,2-propanediol as both solvent and reducing agent study the catalytic properties of these nanoparticles for oxidation
[20]. Furthermore, due to the relatively high dipole moment and of tartrazine, as an organic pollutant, in aqueous solution under
loss factor of polyol solvents, they are also suitable for microwave UVC and visible light irradiations, taking in consider the effect of
assisted route. In this regard, using microwave in the synthesis of different parameters such as initial pH of solution, amount of
nanoparticles is considered a modern and rapidly developing nanoparticles, and concentration of pollutant in the degradation
method. It has been evidenced that microwave irradiation gives a process.
narrow particle size distribution of nanocrystals with a high purity
in a short reaction time in addition of being cheap and 2. Expermental
environmentally friendly [21–23].
Tartrazine (Acid Yellow 23; FD&C Yellow No. 5), as an azo dye, 2.1. Materials and reagents
has been chosen for this study due to its extensively use as a
colorant in food, cosmetics, pharmaceuticals and textile industry All reagents used in our experiments were of analytical grade
[24], as well as its high stability against biodegradation and and used as received without any further purification. Tartrazine
conventional wastewater treatment procedures after disposal powder ( 85%) was obtained from Sigma–Aldrich (Germany).
from industrial effluent [25]. The toxic concentration of tartrazine Cadmium chloride 2.5-hydrate (CdCl2 2.5H2O) and thiourea (SC
on human has been reported to be 7.5 mg k1 [26]. This compound (NH2) 2) were obtained from Scharlau (Spain). Polyvinylpyrroli-
is known to cause allergic reactions such as asthma and urticaria done (MW  25,000 g mol1), ethylene glycol, glacial acetic acid,
and appears to cause more allergic and intolerance reactions than and ammonia solution (25%) were obtained from Merck
other azo dyes [27]. (Germany). Double distilled water was used for the preparation
Previously reported works on the removal of tartrazine from of all samples.
aqueous solutions include mainly conventional methods like
adsorption [28,29], ecocoagulation [30], and filtration [31] which 2.2. Synthesis of nanoparticles
are known not to effectively degrade pollutant but merely transfer
it to another phase where it is more concentrated [32]. Whereas In a typical process, 9 g of cadmium chloride 2.5-hydrate and 3 g
other reports have shown that tartrazine could be degraded of PVP were separately dissolved in 30 mL and 20 mL of EG,
employing AOPs like: ozonation [33], electrochemical oxidation respectively, and then heated with stirring for 60 min at 75  C. The
[34], photo Fenton oxidation [35], UV/H2O2 [36,37], photolytic two solutions were then mixed slowly and heated with stirring for
[38], and photocatalytic oxidation [27,39–46]. 40 min at 100  C. Concurrently, 3 g of thiourea was dissolved in
Most of photocatalytic methods have utilized TiO2, which was 20 mL of EG and heated with stirring for 100 min at 75  C and then
extensively studied and known to have some drawbacks such as added slowly to the hot solution of (PVP-Cd2+). The temperature of
expensive precursors and inability to absorb visible light [47,48]. the new mixture was gradually raised to 170–180  C within 45 min
Hence, the aim of the present work was first, to fabricate CdS and kept at this temperature until most of the solvent was
nanoparticles by a chemical synthesis utilizing the polyol method evaporated. In this stage, the color of the solution changed from
with ethylene glycol as a solvent and polyvinylpyrrolidone (PVP) as light yellow into a dark orange suspension indicating the formation
a capping agent. Microwave irradiation followed by calcination of CdS. Next, the suspension was microwaved (LG MC-2820S
was used for size and shape controlling. Further, to investigate the microwave) for 5 min with a power of 720 W and 50% duty cycle
structure, size, and morphology of the synthesized products using (30 s of irradiation and 30 s stop). The obtained powder was then
various characterization techniques. In second, the work aimed to calcined in a furnace (Sybron Thermolyne Type 1500) for 120 min

Fig. 1. A schematic of the preparation of PVP-CdS nanoparticles.


M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396 389

at 450  C and finally, black crystallites of PVP-capped CdS out using (OSRAM HWL 160 W) lamp in the same experimental
nanoparticles were obtained. The scheme of preparation is setup used under UVC irradiation. In addition, a light protected
illustrated in Fig. 1. A control synthesis in the absence of PVP experiment under the same conditions of visible light experiment
and the exact other steps was carried out to evaluate the impact of was carried out and considered as a blank to assess the
PVP on the yielded product. mineralization of tartrazine after same time of irradiation. Samples
were analyzed by a total organic carbon analyzer type (TOC-V,
2.3. Characterization of the CdS nanoparticles Shimadzu).

The synthesized CdS nanoparticles were characterized using 3. Results and discussion
various analytical techniques. Scanning electron microscopy
(KYKY-EM3200 SEM) was used to investigate the morphologies 3.1. Morphology and structure
of nanoparticles. X-ray diffraction patterns of the nanoparticles
were recorded by a Bruker AXS D8-ADVANCE diffractometer fitted The SEM images presented in Fig. 2 show that PVP-capped CdS
with a (Cu Ka l = 1.5418 Å) radiation tube. The average crystallite (PVP-CdS) particles are in the range of nanometer with narrower
size of the synthesized CdS was calculated with well-known particle size distribution (40–59 nm) and less agglomeration
Scherrer’s equation (Eq. (1)) with full width at half maxima comparing to those of the uncapped CdS which has an average
(FWHM) of X-ray diffraction pattern: particle size of 115 nm (75–180) indicating clearly the advantages of
PVP in restricting the growth and preventing agglomeration. This is
0:89l
D¼ ð1Þ simply due to the stereo-hindrance effect resulted from repulsive
bcosu force among the polyvinyl groups of PVP and due to the restrained
where l is the wavelength of X-ray radiation source, b is FWHM of Ostwald ripening kinetics in such a way that the growth rate was
the peak and u is Bragg’s diffraction angle. decreased with the size of CdS nanoparticles. Consequently, smaller
Furthermore, in order to determine the chemical composition particles with enhanced monodispersity were obtained.
of nanoparticles and study the behavior of PVP, Fourier transform X-ray powder diffraction patterns of the capped nanoparticles
infrared spectra in the range of 400–4000 cm1 of many samples presented in Fig. 3 show a perfect match with the hexagonal phase
during the synthesis process was recorded in transmission mode of CdS (JCPDS No. 01-089-2944). Diffraction peaks of monoclinic
(Thermo Nicolet 8700 FTIR), the powder samples were grounded sulfur are also found (JCPDS No. 01-074-2108). The average
with KBr and compressed into a pellet for analysis.

2.4. Photocatalytic activity under UVC light

Photocatalytic activity of the nanoparticles was evaluated by


measuring the degradation of tartrazine which was used as a test
pollutant. An amount of (5–50) mg of the catalyst was dispersed in
200 mL of tartrazine aqueous solution (10–100) mg L1. The pH of
the suspension was adjusted to the desired value by ammonia or
glacial acetic acid solutions and determined by a digital pH meter
(Metrohm pH lab 827). The suspension in the photoreaction
container was exposed to UV light source (CH Lighting T8 30 W
UVC) positioned 10 cm above for 90–180 min. At given intervals of
illumination, 5 mL samples were taken out and centrifuged
(Hettich EPA 12) at 2500 rpm for 10 min in order to completely
remove all nanoparticles. The degradation ratio was monitored
using UV–vis spectrophotometer (T80 + UV–vis Spectrometer PG
Instruments Ltd.). According to the Beer–Lambert law, the
concentration of tartrazine is proportional to its absorbance.
Hence, the degradation efficiency was calculated by the following
formula:
   
ðC 0  CÞ ðA0  AÞ
D% ¼  100 ¼  100 ð2Þ
C0 A0
C0 and A0 are the initial concentration and absorbance, respective-
ly, while C and A are the concentration and absorbance after
intervals of illumination (t). The reaction rate constant (k) was
calculated using plots of ln (C/C0) versus illumination time
according to the following formula:
 
C
ln ¼ kt ð3Þ
C0

2.5. Photocatalytic degradation and mineralization under visible light

After obtaining the optimal conditions for tartrazine degrada-


tion with the synthesized nanoparticles under UVC, an assessment
of the activity of these nanoparticles under visible light was carried
Fig. 2. SEM images of (a) uncapped and (b) PVP-capped CdS nanoparticles.
390 M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396

Fig. 3. XRD patterns of the PVP-CdS nanoparticles.

crystallite size of the capped nanoparticles calculated from XRD amount of PVP-CdS (25 mg) were set at pH 3.5, 5.5, 7 or 11 and kept
patterns using the Scherrer’s equation was about 48 nm which is in in dark place with constant stirring for 90 min in order to reach the
agreement with the result obtained from SEM image. equilibrium adsorption. In addition, photolysis experiments, in the
FTIR spectra of pure PVP and PVP-CdS prior to and after absence of catalyst and presence of UVC light with the same other
calcination are shown in Fig. 4. In pure PVP spectrum, the O H experimental setup used in the adsorption study, were carried out
stretching vibration observed as a broad and strong peak at also over 90 min. The role of pH on photocatalytic degradation of
3517 cm1 is due to H2O absorption on the surface of the sample. tartrazine was investigated in the same conditions of photolysis in
The two bands at 2952 and 1423 cm1 correspond to C H the presence of 25 mg of PVP-CdS or uncapped CdS over 90 min.
asymmetric stretching vibration and bending vibration, respec- The results are shown in Fig. 6.
tively. The broad peak with high intensity at 1665 cm1 is ascribed As can be seen, a decrease in the dye concentration between 3
to the C¼O bond stretching and the band at 1286 cm1 corresponds and 8% was obtained among the different pH values in the presence
to C N bond stretching of the pyrrolidone structure. of PVP-CdS only without UVC irradiation. Whereas about 4.5–12%
Two observations can be found when comparing pure PVP of the initial tartrazine concentration was removed at different pH
spectrum with the other two spectra. First, a decrease in all peaks values in the presence of UVC irradiation alone without nano-
intensities after exposing the sample to microwave irradiation and particles. On the other hand, maximum photocatalytic degradation
more decrease after calcination. This can be explained in the light percentage of PVP-CdS was observed in alkaline region. The
of thermal degradation of PVP. Loría-Bastarrachea et al. [49], in degradation increased with increasing pH and reached its
agreement with earlier reports, stated that at 450  C the main maximum value at pH 11. For uncapped CdS, neutral medium
product obtained from thermal degradation of PVP is its was favorable.
corresponding monomer, i.e., vinyl pyrrolidone or less unsaturated Our suggested interpretation is the following: tartrazine has a
compounds such as PVP oligomers. The second observation is the pKa = 9.4 and so is negatively charged in alkaline medium due to its
shift to lower wavenumber of some peaks, specifically C¼O and ionization to the related phenyl salt. Likewise, CdS has a zero point
CN stretching vibrations. These shifting indicate the interaction charge at about pH 7 and is also negatively charged in pH (8–12). As
between PVP and Cd2+. The mechanism of this interaction as a result, the adsorption of tartrazine on CdS surface is minimum at
suggested by Abdelghany et al. [50] implies the formation of pH 11. Here, the enhancement of capped catalyst activity in
coordinate bonds between nitrogen and/or oxygen atoms of the alkaline medium could be imputed to the attachment of CdS with
PVP with cadmium ions in four different probabilities as shown in pyrrolidone ring that might have acted as a mediator to trap the
Fig. 5. Hence, the preceding findings indicate the formation of short carriers generated in the catalyst and adsorb the water or dye
polymeric chain-capped CdS nanoparticles. molecules or even the abundant hydroxyl ions, which would
subsequently facilitate the generation of hydroxyl radicals that
3.2. Effect of tartrazine solution initial pH initiate the oxidization of the pollutant.
Unsurprisingly, PVP-CdS has shown a superior activity over
The pH of the reaction medium is one of the most crucial uncapped CdS in extreme acidic and basic conditions whereas in
parameters in photocatalytic degradation of organic pollutants. It neutral conditions, the activities were nearly equivalent. This is
affects the surface charges of both the catalyst and pollutant and ascribed to two factors: first, the smaller particles size that offered
consequently the adsorption of the pollutant on catalyst surface more surface area for photocatalytic reaction in the case of PVP-
[51]. CdS and second, the photocorrosin or dissolution that uncapped
In order to study the role of pH on the adsorption of tartrazine CdS might underwent in extreme conditions which has reduced
on the catalyst surface, 20 mg L1 solutions of the dye and a fixed drastically its activity. Meanwhile, PVP-CdS maintained an
M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396 391

Fig. 4. FTIR spectra of (A) pure PVP, (B) PVP-CdS after microwave irradiation and, (C) PVP-CdS after calcination.

approximate percentages of degradation in acidic and neutral 50 mg) with 50 mg L1 initial concentration of tartrazine at pH 11
conditions as a consequence of PVP passivation. for 90 min of irradiation. The results are shown in Fig. 7.
As is noted, the photodegradation of the tartrazine was found to
increase with the increased catalyst loading within the range of
3.3. Effect of the catalyst loading
study. Increasing the catalyst loading gave a rise to a higher
number and density of nanoparticles, which in turn caused more
Catalyst loading is a critical parameter in photodegradation of
photons to be absorbed and dye molecules to be adsorbed leading
organic pollutants. It affects the reaction rate and consequently the
to better degradation [53]. The opacity of the suspension was not
cost of treatment [52].
greatly affected by the increased amount of PVP-CdS and no
The effect of PVP-CdS dosage on the degradation of tartrazine
hindering to light penetration was observed. On the other hand,
was investigated in the presence of different amount of catalyst (5–
after 25 mg of catalyst amount there was no considerable
392 M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396

Fig. 4. (Continued)

Fig. 5. Suggested mechanism of interaction between PVP and Cd2+.


Ref. [50].

enhancement in the degradation profile. Hence, 25 mg of PVP-CdS


nanoparticles was chosen to be the optimal amount for further 3.4. Effect of the initial concentration of tartrazine
studies.
The influence of pollutant concentration on photocatalytic
degradation has also been studied. Experiments have been carried
out at different initial concentrations in the range (10–100) mg L1
with constant amount of nanoparticles (25 mg) and initial
solution’s pH 11. The results shown in Fig. 8 indicate that
increasing the concentration of tartrazine had decreased the
degradation from 100% (10 mg L1) to 92.9% (50 mg L1) and 36.2%
(100 mg L1) in 60 min.
This can be explained in the light of photocatalytic degradation
mechanism; a single layer of adsorbed water molecules exists on
the surface of PVP-CdS. After photon is adsorbed, the holes
generated in valence band transfer to the surface and oxidize water
molecules and hydroxyl ions to produce HO radicals. These
radicals rapidly attack tartrazine molecules and oxidize them to
intermediates which in turn would be mineralized with subse-
quent attacks of HO radicals.
PVP  CdS þ hn ! eCB þ hVBþ ð4Þ

Fig. 6. Effect of different pH values (3.5–11) on adsorption, photolysis and,


photocatalytic degradation of tartrazine over 90 min.
M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396 393

Fig. 7. Photocatalytic degradation of tartrazine by varying the amount of PVP-CdS


nanoparticles (5–50 mg) with Tartrazine concentration (50 mg L1) and pH 11. Fig. 9. Plots of ln (C/C0) versus illumination time for different concentrations of
tartrazine.

þ 3.5. Kinetic study


h þ H2 O ! HO þ Hþ ð5Þ

In order to study the pattern of the photodegradation reaction


 of tartrazine, Langmuir–Hinshelwood (L–H) model was used. This
HO þ TA ! Intermediates ð6Þ
model has been successfully used for heterogeneous photo-
catalysis to describe the exact relationship between the rate of
degradation and the concentration of pollutant in photocatalytic
HO þ Intermediates ! CO2 þ H2 O ð7Þ
reaction [31]. The rate constant (k) of degradation reaction was
Increasing the initial dye concentration means more tartrazine calculated according to the Formula (3). As shown in Fig. 9, a linear
molecules to surround the nanoparticles and decrease the path relation between tartrazine concentrations and illumination time
length of the photons entering the solution resulting in lower has been observed implying a pseudo-first order kinetics of the
photon absorption by the catalyst and reaction sequence is slowed photodegradation reaction. The pseudo-first order rate constant
down. Consequently, rate of degradation is decreased. At this decreased dramatically with increased initial concentration of
juncture, we believe that PVP passivation has an important role in tartrazine. This is because, as aforementioned, with high concen-
the separation of the photogenerated carriers. This is attributed to trations of the pollutant, less hydroxyl radicals in active sites of the
the prohibition of nonradiative decay of carriers by covering catalyst are generated and the majority of photons are absorbed by
dangling bonds that may work as exciton traps. Also, the pollutant molecules rather than catalyst [54].
coordination of nanoparticles surface atoms which minimizes In order to cover the adsorption properties of the substrate on
the trap states lie within the band gap and increases quantum yield the catalyst surface, the experimental data have been rationalized
by providing alternative pathways of electron–hole separation. in terms of the modified form of L–H kinetic model. The modified
More importantly, complexation between Cd and the lone pair of Langmuir–Hinshelwood expression that explains the kinetics of
nitrogen/oxygen makes pyrrolidone moiety of PVP act as a hole heterogeneous catalytic systems is given by:
acceptor that traps the photogenerated holes and transfers them to
dc kr KC
the target objects. It is worth mentioning here that partial r¼ ¼ ð8Þ
dt ð1 þ KCÞ
decomposition of PVP, during the synthesis process, was beneficial
for exposing trapping sites of pyrrolidone to those targeted objects where r represents the rate of reaction that changes with time
to promote the catalytic reaction sequence. (mg L1 min1), kr is the reaction rate constant (mg L1 min1), and
K is the adsorption coefficient of the reactant onto the catalyst
particles (L mg 1).

Fig. 8. Photocatalytic degradation of varying initial concentrations of tartrazine


(10–100 mg L1) with pH 11 and 25 mg amount of PVP-CdS nanoparticles.
Fig. 10. Plot of 1/r0 versus 1/C0 of the photocatalytic degradation of tartrazine.
394 M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396

when compared with pollutant adsorption onto the surface of the


catalyst.

3.6. Reusability

Reusability of capped catalyst was investigated in order to


verify its stability and the protective role of PVP in preventing
catalyst degradation. After a first round of catalytic reaction, the
mixture (catalyst+slurry) was filtered and washed with a mixture
of water and ethanol (1:1) then dried in an oven at 65  C for 6 h. The
recovered catalyst was used again to degrade tartrazine under
same conditions used before. A third round was carried out in the
same procedure. Results depicted in Fig. 11 show a very slight
deference in catalytic activity among the first used and reused
catalysts. A small change in the kinetic could be seen but, the
Fig. 11. Reusability of PVP-CdS nanoparticles.
overall degradation in 90 min of study was not importantly
affected. It is our believe that photocorrosion suppression is due to
The applicability of L–H equation for the degradation has been prolonged life time of electron–hole pairs as a consequence of hole
confirmed by the linear plot obtained by plotting the reciprocal of trapping and transferring mediated by PVP.
initial rates (1/r0) against reciprocal of initial concentrations of
tartrazine (1/C0) as shown in Fig. 10. 3.7. Photodegradation and mineralization under visible light
   
1 1 1 irradiation
¼ þ ð9Þ
r0 kr kr KC 0
A 50 mg L1 solution of tartrazine with pH 11 was mixed with
In this study, a reasonable agreement (R2 = 0.969) was obtained 25 mg of PVP-CdS nanoparticles and exposed to visible light
between the experimental results and the linear form of the L–H irradiation in the same experimental setup used under UVC
expression. irradiation. As can be seen in Fig. 12, approximately 85% of the dye
The values of kr and K have been determined from the slope and was degraded and about 69% of the total organic content compared
intercept of this plot and found to be 1.908 and 0.036, respectively, with the blank solution was removed after 9 h.
indicating that photocatalytic degradation is a dominant factor

Fig. 12. Photocatalytic degradation and mineralization of 50 mg L1 tartrazine with pH 11 and 25 mg of PVP-CdS nanoparticles under visible light irradiation.
M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396 395

4. Conclusion [13] N. Soltani, E. Saion, M.Z. Hussein, R.B. Yunus, M. Navaseri, Characterization of
CdS nanoparticles synthesized using microwave-assisted polyol method, Adv.
Mater. Res. 667 (2013) 122–127, doi:http://dx.doi.org/10.4028/www.scientific.
Polyol synthesis with microwave irradiation and calcination net/amr.667.122.
have been used as an efficient method to fabricate cadmium sulfide [14] Q. Xia, X. Chen, K. Zhao, J. Liu, Synthesis and characterizations of
nanoparticles in an average crystallite size of about 48 nm when polycrystalline walnut-like CdS nanoparticle by solvothermal method with
PVP as stabilizer, Mater. Chem. Phys. 111 (1) (2008) 98–105, doi:http://dx.doi.
PVP was used as a capping agent and about 115 nm without org/10.1016/j.matchemphys.2008.03.020.
capping. The prepared nanoparticles have been effectively used as [15] Z. Jing, L. Tan, F. Li, J. Wang, Y. Fu, Q. Li, Photocatalytic and antibacterial
a photocatalyst for the degradation of tartrazine in aqueous activities of CdS nanoparticles prepared by solvothermal method, Indian J.
Chem. A 52 (1) (2013) 57–62.
solution. Total degradation of tartrazine was obtained after 90 min [16] S.H. Liu, X.F. Qian, J. Yin, X.D. Ma, J.Y. Yuan, Z.K. Zhu, Preparation and
using 25 mg of the PVP-CdS with initial dye concentration of characterization of polymer-capped CdS nanocrystals, J. Phys. Chem. Solids 64
50 mg L1 and pH 11 under UVC irradiation. The role of PVP in (3) (2003) 455–458, doi:http://dx.doi.org/10.1016/s0022-3697(02) 00333-5.
[17] M. Pattabi, B. Saraswathi Amma, K. Manzoor, Photoluminescence study of PVP
controlling the growth and stabilizing the nanoparticles was
capped CdS nanoparticles embedded in PVA matrix, Mater. Res. Bull. 42 (5)
demonstrated by the compassion with the uncapped product (2007) 828–835, doi:http://dx.doi.org/10.1016/j.materresbull.2006.08.029.
which has shown more growing and less catalytic activity. [18] R. Pawar, J.-Y. Lee, E.-J. Kim, H. Kim, C.S. Lee, Synthesis of CdS with graphene by
The Langmuir–Hinshelwood models indicated that tartrazine CBD (chemical bath deposition) method and its photocatalytic activity, Korean
J. Mater. Res. 22 (10) (2012) 504–507, doi:http://dx.doi.org/10.3740/
underwent a pseudo first order kinetics of the reaction and that MRSK.2012.22.10.504.
photocatalytic degradation was prevailing compared with pollut- [19] F. Fiévet, R. Brayner, The polyol process, in: R. Brayner, F. Fiévet, T. Coradin
ant adsorption on PVP-CdS surface. In addition, results has shown (Eds.), Nanomaterials: A Danger or a Promise?, Springer, London, 2013, pp. 1–
25, doi:http://dx.doi.org/10.1007/978-1-4471-4213-3_1.
that our nanoparticles can work sufficiently under visible light [20] G.G. Couto, J.J. Klein, W.H. Schreiner, D.H. Mosca, A.J.A. de Oliveira, A.J.G.
with about 85% abatement of initial tartrazine concentration and Zarbin, Nickel nanoparticles obtained by a modified polyol process: synthesis,
69% mineralization after 9 h of irradiation. characterization, and magnetic properties, J. Colloid Interface Sci. 311 (2)
(2007) 461–468, doi:http://dx.doi.org/10.1016/j.jcis.2007.03.045.
In conclusion, PVP-CdS nanoparticles synthesized in this work [21] Z. Poormohammadi-Ahandani, A. Habibi-Yangjeh, Fast, green and template-
might be used effectively in environmental treatment processes to free method for preparation of Zn1xCdxS nanoparticles using microwave
remove organic pollutants from water. These nanoparticles are irradiation and their photocatalytic activities, Phys. E 43 (1) (2010) 216–223,
doi:http://dx.doi.org/10.1016/j.physe.2010.07.009.
promising to work under visible light sources to reduce greatly the
[22] R. Shahid, M.S. Toprak, M. Muhammed, Microwave-assisted low temperature
expense of treatment. synthesis of wurtzite ZnS quantum dots, J. Solid State Chem. 187 (0) (2012)
130–133, doi:http://dx.doi.org/10.1016/j.jssc.2012.01.007.
[23] S.K. Choubey, K. Tiwary, Microwave assisted synthesis of CdS nanoparticles for
Acknowledgment
structural and optical characterization, IJIRSET 3 (3) (2014) 10670–10674.
[24] T. Tanaka, Reproductive and neurobehavioural toxicity study of tartrazine
The authors wish to thank Tehran University of Medical administered to mice in the diet, Food Chem. Toxicol. 44 (2) (2006) 179–187,
Sciences for the financial and instrumental support of this doi:http://dx.doi.org/10.1016/j.fct.2005.06.011.
[25] E.S. Beach, R.T. Malecky, R.R. Gil, C.P. Horwitz, T.J. Collins, Fe-TAML/hydrogen
research. peroxide degradation of concentrated solutions of the commercial azo dye
tartrazine, Catal. Sci. Technol. 1 (3) (2011) 437–443, doi:http://dx.doi.org/
References 10.1039/c0cy00070a.
[26] H.M. El-Wahab, G.S. Moram, Toxic effects of some synthetic food colorants
and/or flavor additives on male rats, Toxicol. Ind. Health 29 (2) (2013) 224–232,
[1] H. Zhang, G. Chen, D.W. Bahnemann, Photoelectrocatalytic materials for doi:http://dx.doi.org/10.1177/0748233711433935.
environmental applications, J. Mater. Chem. 19 (29) (2009) 5089–5121, doi: [27] S.K. Al-Dawery, Photo-catalyst degradation of tartrazine compound in
http://dx.doi.org/10.1039/b821991e. wastewater using TiO2 and UV light, J. Eng. Sci. Technol. 8 (6) (2013) 683–691.
[2] C.C. Kaan, A.A. Aziz, S. Ibrahim, M. Matheswaran, P. Saravanan, Heterogeneous [28] A. Mittal, L. Kurup, J. Mittal, Freundlich and Langmuir adsorption isotherms
photocatalytic oxidation an effective tool for wastewater treatment–a review, and kinetics for the removal of tartrazine from aqueous solutions using hen
Studies on Water Management Issues, InTech, 2012, pp. 219–236, doi:http:// feathers, J. Hazard. Mater. 146 (1–2) (2007) 243–248, doi:http://dx.doi.org/
dx.doi.org/10.5772/30134. 10.1016/j.jhazmat.2006.12.012.
[3] M.I. Litter, Heterogeneous photocatalysis: transition metal ions in [29] A. Mittal, J. Mittal, L. Kurup, Adsorption isotherms, kinetics and column
photocatalytic systems, Appl. Catal. B: Environ. 23 (2–3) (1999) 89–114, doi: operations for the removal of hazardous dye, tartrazine from aqueous
http://dx.doi.org/10.1016/s0926-3373(99) 00069-7. solutions using waste materials-bottom ash and de-oiled soya, as adsorbents,
[4] H.R. Pouretedal, M.H. Keshavarz, M.H. Yosefi, A. Shokrollahi, A. Zali, J. Hazard. Mater. 136 (3) (2006) 567–578, doi:http://dx.doi.org/10.1016/j.
Photodegradation of HMX and RDX in the presence of nanocatalyst of zinc jhazmat.2005.12.037.
sulfide doped with copper, Iran J. Chem. Chem Eng. 28 (2009) 13–19. [30] N. Modirshahla, M.A. Behnajady, S. Kooshaiian, Investigation of the effect of
[5] A. Mercy, R.S. Selvaraj, B.M. Boaz, A. Anandhi, R. Kanagadurai, Synthesis, different electrode connections on the removal efficiency of tartrazine from
structural and optical characterisation of cadmium sulphide nanoparticles, aqueous solutions by electrocoagulation, Dyes Pigment 74 (2) (2007) 249–257,
Indian J. Pure Appl. Phys. 51 (6) (2013) 448–452. doi:http://dx.doi.org/10.1016/j.dyepig.2006.02.006.
[6] H.R. Pouretedal, H. Eskandari, M.H. Keshavarz, A. Semnani, Photodegradation [31] C. Aydiner, Y. Kaya, G. Beril, Z. onder, I. Vergili, Evaluation of membrane fouling
of organic dyes using nanoparticles of cadmium sulfide doped with and flux decline related with mass transport in nanofiltration of tartrazine
manganese, nickel and copper as nanophotocatalyst, Acta Chim. Slov. 56 solution, J. Chem. Technol. Biotechnol. 85 (9) (2010) 1229–1240, doi:http://dx.
(2009) 353–361. doi.org/10.1002/jctb.2422.
[7] R. Chauhan, A. Kumar, R.P. Chaudhary, Visible-light photocatalytic degradation [32] K. Rajeshwar, M.E. Osugi, W. Chanmanee, C.R. Chenthamarakshan, M.V.B.
of methylene blue with Fe doped CdS nanoparticles, Appl. Surf. Sci. 270 (0) Zanoni, P. Kajitvichyanukul, R. Krishnan-Ayer, Heterogeneous photocatalytic
(2013) 655–660, doi:http://dx.doi.org/10.1016/j.apsusc.2013.01.110. treatment of organic dyes in air and aqueous media, J. Photochem. Photobiol. 9
[8] Y. Yu, Y. Ding, S. Zuo, J. Liu, Photocatalytic activity of nanosized cadmium (4) (2008) 171–192, doi:http://dx.doi.org/10.1016/j.
sulfides synthesized by complex compound thermolysis, Int. J. Photoenergy jphotochemrev.2008.09.001.
2011 (5) (2011) , doi:http://dx.doi.org/10.1155/2011/762929. [33] H.-Y. Shu, C.-R. Huang, Degradation of commercial azo dyes in water using
[9] S. Peretz, D.F. Anghel, E. Teodor, G. Stanciu, C. Stoian, G. Zgherea, M. Florea- ozonation and UV enhanced ozonation process, Chemosphere 31 (8) (1995)
Spiroiu, Improving the properties of CdS nanoparticles by adding polymers, 3813–3825, doi:http://dx.doi.org/10.1016/0045-6535(95) 00255-7.
Part. Sci. Technol. 29 (3) (2011) 229–241, doi:http://dx.doi.org/10.1080/ [34] R. Jain, M. Bhargava, N. Sharma, Electrochemical studies on a pharmaceutical
02726351.2010.494707. azo dye: tartrazine, Ind. Eng. Chem. Res. 42 (2) (2003) 243–247, doi:http://dx.
[10] G. Ghosh, M. Kanti Naskar, A. Patra, M. Chatterjee, Synthesis and doi.org/10.1021/ie020228q.
characterization of PVP-encapsulated ZnS nanoparticles, Opt. Mater. 28 (8–9) [35] P. Oancea, V. Meltzer, Photo-Fenton process for the degradation of tartrazine
(2006) 1047–1053, doi:http://dx.doi.org/10.1016/j.optmat.2005.06.003. (E102) in aqueous medium, J. Taiwan Inst. Chem. Eng. 44 (6) (2013) 990–994,
[11] A.F. Hezinger, J. Tessmar, A. Gopferich, Polymer coating of quantum dots-a doi:http://dx.doi.org/10.1016/j.jtice.2013.03.014.
powerful tool toward diagnostics and sensorics, Eur. J. Pharm. Biopharm. 68 (1) [36] F. Parolin, U.M. Nascimento, E.B. Azevedo, Microwave-enhanced UV/H2O2
(2008) 138–152, doi:http://dx.doi.org/10.1016/j.ejpb.2007.05.013. degradation of an azo dye (tartrazine): optimization, colour removal,
[12] L. Saravanan, S. Diwakar, A. Pandurangan, R. Jayavel, Synthesis, structural and mineralization and ecotoxicity, Environ. Technol. 34 (2013) 1247–1253, doi:
optical properties of PVP encapsulated CdS nanoparticles, Nanomater. http://dx.doi.org/10.1080/09593330.2012.7444319-12).
Nanotechnol. 1 (2) (2011) 42–48, doi:http://dx.doi.org/10.5772/50959.
396 M. Darwish et al. / Materials Research Bulletin 74 (2016) 387–396

[37] P. Oancea, V. Meltzer, Kinetics of tartrazine photodegradation by UV/H2O2 in [46] N. Arabzadeh, A. Khosravi, A. Mohammadi, N.M. Mahmoodi, Enhanced
aqueous solution, Chem. Papers 68 (1) (2014) 105–111, doi:http://dx.doi.org/ photodegradation of hazardous tartrazine by composite of nanomolecularly
10.2478/s11696-013-0426-5. imprinted polymer-nanophotocatalyst with high efficiency, Desalin. Water
[38] T.C. dos Santos, G.J. Zocolo, D.A. Morales, A. Umbuzeiro Gde, M.V. Zanoni, Treat. (2014) 1–10, doi:http://dx.doi.org/10.1080/19443994.2014.989414.
Assessment of the breakdown products of solar/UV induced photolytic [47] M. Umar, H.A. Aziz, Photocatalytic degradation of organic pollutants in water,
degradation of food dye, Food Chem. Toxicol. 68 (2014) 307–315, doi:http://dx. Organic Pollutants—Monitoring, Risk and Treatment, InTech, 2013, doi:http://
doi.org/10.1016/j.fct.2014.03.025. dx.doi.org/10.5772/53699.
[39] V.K. Gupta, R. Jain, A. Nayak, S. Agarwal, M. Shrivastava, Removal of the [48] S.M. Gupta, M. Tripathi, A review of TiO2 nanoparticles, Chin. Sci. Bull. 56 (16)
hazardous dye—tartrazine by photodegradation on titanium dioxide surface, (2011) 1639–1657, doi:http://dx.doi.org/10.1007/s11434-011-4476-1.
Mater. Sci. Eng. C 31 (5) (2011) 1062–1067, doi:http://dx.doi.org/10.1016/j. [49] M.I. Loría-Bastarrachea, W. Herrera-Kao, J.V. Cauich-Rodríguez, J.M. Cervantes-
msec.2011.03.006. Uc, H. Vázquez-Torres, A. Ávila-Ortega, A TG/FTIR study on the thermal
[40] N. Modirshahla, A. Hassani, M.A. Behnajady, R. Rahbarfam, Effect of degradation of poly(vinyl pyrrolidone), J. Therm. Anal. Calorim. 104 (2) (2011)
operational parameters on decolorization of acid yellow 23 from wastewater 737–742, doi:http://dx.doi.org/10.1007/s10973-010-1061-9.
by UV irradiation using ZnO and ZnO/SnO2 photocatalysts, Desalination 271 [50] A.M. Abdelghany, E.M. Abdelrazek, D.S. Rashad, Impact of in situ preparation of
(1–3) (2011) 187–192, doi:http://dx.doi.org/10.1016/j.desal.2010.12.027. CdS filled PVP nano-composite, Spectrochim. Acta A Mol. Biomol. Spectrosc.
[41] M.A. Behnajady, N. Modirshahla, R. Hamzavi, Kinetic study on photocatalytic 130 (2014) 302–308, doi:http://dx.doi.org/10.1016/j.saa.2014.04.049.
degradation of C.I. acid yellow 23 by ZnO photocatalyst, J. Hazard. Mater. 133 [51] M. Mrowetz, E. Selli, Photocatalytic degradation of formic and benzoic acids
(1–3) (2006) 226–232, doi:http://dx.doi.org/10.1016/j.jhazmat.2005.10.022. and hydrogen peroxide evolution in TiO2 and ZnO water suspensions, J.
[42] K. Tanaka, K. Padermpole, T. Hisanaga, Photocatalytic degradation of Photochem. Photobiol. 180 (1–2) (2006) 15–22, doi:http://dx.doi.org/10.1016/j.
commercial azo dyes, Water Res. 34 (1) (2000) 327–333, doi:http://dx.doi.org/ jphotochem.2005.09.009.
10.1016/s0043-1354(99)00093-7. [52] T.A. Gad-Allah, M.E.M. Ali, M.I. Badawy, Photocatalytic oxidation of
[43] C. Andriantsiferana, E.F. Mohamed, H. Delmas, Photocatalytic degradation of ciprofloxacin under simulated sunlight, J. Hazard. Mater. 186 (1) (2011)
an azo-dye on TiO2/activated carbon composite material, Environ. Technol. 35 751–755, doi:http://dx.doi.org/10.1016/j.jhazmat.2010.11.066.
(3) (2014) 355–363, doi:http://dx.doi.org/10.1080/09593330.2013.828094. [53] N. Assi, A. Mohammadi, Q. Sadr Manuchehri, R.B. Walker, Synthesis and
[44] A. Jodat, A. Jodat, Photocatalytic degradation of chloramphenicol and characterization of ZnO nanoparticle synthesized by a microwave-assisted
tartrazine using Ag/TiO2 nanoparticles, Desalin. Water Treat. 52 (13–15) (2014) combustion method and catalytic activity for the removal of ortho-
2668–2677, doi:http://dx.doi.org/10.1080/19443994.2013.794115. nitrophenol, Desalin. Water Treat. 54 (7) (2014) 1939–1948, doi:http://dx.doi.
[45] M. Teimouri, P. Aberoomand, S. Moradi, M. Zhalechin, S. Piramoon, Synthesis of org/10.1080/19443994.2014.891083.
nickel-doped TiO2 nano crystalline by the sol–gel method and influence of [54] H.R. Pouretedal, A. Kadkhodaie, Synthetic CeO2 nanoparticle catalysis of
ultrasonic irradiation for the photo catalytic degradation of Tartrazine dye, The methylene blue photodegradation: kinetics and mechanism, Chin. J. Catal. 31
4th International Conference on Nanostructures (ICNS4): 12–14 March (2012) . (11–12) (2010) 1328–1334, doi:http://dx.doi.org/10.1016/s1872-2067(10)
60121-0.

Potrebbero piacerti anche