elsayed2020
elsayed2020
A R T I C L E I N F O A B S T R A C T
Article History: This work reports the preparation of Na-doped ZnO (Na-ZnO) as a promising photocatalyst under sunlight
Received 26 July 2020 through a very simple precipitation method. The crystal and electronic structures of the prepared Na-ZnO
Revised 6 October 2020 were characterized by experimental techniques and Density Functional Theory (DFT) calculations. Where the
Accepted 19 October 2020
DFT calculations revealed that the effect of Na-doping on the ZnO electronic structure is consistent with the
Available online xxx
experimental observations and the bandgap value found to be decreasing. Under direct sunlight, the as-pre-
pared Na-ZnO decolorized more than 99.5% of the methylene blue dye (MB) after 1.5 h and led to mineraliza-
Keywords:
tion of 98.3% of it after 3 h. While pure ZnO decolorized only 35% of MB dye at the same conditions. The
Na-doped ZnO
Direct sunlight
prepared Na-ZnO has a promising photocatalytic activity at different pH’s and high photostability reaching
DFT calculations to 7 recycle with the same efficiency. Finally, the primary active species produced during the photocatalytic
Photocatalytic degradation reaction were determined and the photocatalytic mechanism was elucidated.
Precipitation method © 2020 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
https://doi.org/10.1016/j.jtice.2020.10.018
1876-1070/© 2020 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
2 M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111
were mineralized after 3 h. The prepared Na-ZnO has both high pho- structure of pure compounds can be solved through the analysis of
tocatalytic activity and high photostability at different pH’s, where the position and intensity of the diffraction peaks. In this study, X-ray
more than 99% of MB are decolorized and 80.4% are mineralized after diffractograms of the prepared samples were acquired at room tem-
7 recycles. perature using X'Pert PRO PANalyticalNetherland equipment;
measurements were made in scan mode with Cu-Ka radiation
2. Experimental section (λ = 0.1542 nm) and the range of 2Ѳ from 10° to 80° For calculating
the average crystallite size we used the Debye-Scherrer equation.
2.1. Chemical and reagents The functional groups and structural properties of the samples were
analyzed using Fourier transform infrared (FTIR) spectroscopy (Shi-
The following materials were used without any purification: zinc madzu FTIR spectrometer). X-ray photoelectron spectroscopy (XPS)
acetate dihydrate [Zn(CH3CO2)2¢2H2O], Ferrous ammonium sulfate (ULVAC-PHI 5000 VersaProbe II) was used to examine the surface ele-
(FAS) [(NH₄)2Fe(SO₄)₂.6H₂O] were bought from LOBA Chemie. Methy- mental composition. The materials thin films were prepared by spin-
lene Blue dye is a cationic dye with the chemical structure shown in coating on ITO substrates for XPS measurement. The morphology,
Fig. 1 (C16H18N3SCl, molar mass of 319.85 g mol1), it was obtained composition and crystalline information of the photocatalyst were
from Oxford Laboratory Reagents. Sodium hydroxide (NaOH), potas- characterized by high-resolution transmission electron microscopy
sium dichromate (K₂Cr₂O₇), ferrous sulfate (FeSO₄), ethyl alcohol (HRTEM; Tecnai G20, FEI, Netherlands) and scanning electron micros-
(C₂H₅OH), ethylenediaminetetraacetic acid (EDTA) [C10H16N2O8], copy (SEM; Quanta 250 FEG (field emission gun)) at an accelerating
concentrated sulphuric acid (H2SO4) were obtained from the Biotech voltage of 200 kV and 30 kV respectively. The surface properties
Laboratory Chemicals Company. Ascorbic acid (C6H8O6), tert‑butanol were investigated via N2 adsorption at 77.4 K using an ASAP2010
(C4H10O) and 1,10-phenanthroline monohydrate (C12H8N2¢H2O) instrument. Before starting the measurement, the samples were out-
were obtained from Merck company. Nitric acid and sodium hydrox- gassed at 343 K for 3 h under a reduced pressure of 105 Torr. Bruna-
ide diluting solutions (1Molar) were used as a regulator for pH during uer-Emmett-Teller (BET) theory was used for determining the spe-
the experiments. All solutions were prepared with distilled water. cific sample surface area. Inductively coupled plasma optical
emission spectroscopy (ICP-OES, Leeman, USA) was used to detect
2.2. Preparation of Na-doped ZNO nanoparticles and quantify the amount of sodium in the Na-ZnO sample. The optical
properties of the samples were examined by the spectra of ultravio-
As illustrated in Fig. 1, Na-doped ZnO nanoparticles were pre- let-visible absorption (Jasco-V 570 spectrophotometer, Japan) and
pared by a very simple precipitation method using Zn fluorescence spectra (Hitachi F-7000 spectrophotometer) at room
(CH3COO)2¢2H2O as the Zn precursor and NaOH as both a precipitat- temperature.
ing and doping agent. Typically, 18.0 mmol of NaOH solution was
added dropwise into 6.0 mmol of Zn(CH3COO)2¢2H2O solution with 2.4. Computational study
continuous stirring. The produced white Zn(OH)2 precipitate was
maintained at 70 °C overnight. Afterward, the precipitate was The calculations in this work have been done to investigate the
allowed to cool to the room temperature and was washed 5 times doping effect of Na on ZnO using the first-principles calculations
with a (70:30%) water : ethyl alcohol solution, then dried at 70 °C based on DFT as implemented in the Vienna ab initio simulation
overnight. Pure ZnO photocatalyst was prepared in the same way as package (VASP) ) [28]. For the total energy prediction, the exchange-
before but with a different amount of NaOH, where only 12.0 mmol correlation energy function treated by the generalized gradient
of NaOH was added to 6.0 mmol of Zn(CH3COO)2¢2H2O. approximation (GGA) [29] of the Perdew-Burke- Ernzerhof formula-
tion (PBE) [30] has been applied throughout the system with spin
2.3. Characterization of Na-doped ZnO nanoparticles polarization. In these calculations, the plane-wave cutoff energy was
set at 500 eV and the total energy convergence better than 104 eV
X-ray diffraction (XRD) is used to identify and characterize crystal per unit cell was required. The Monkhorst-Pack scheme k-point grid
structures based on characteristic diffractograms. The crystal sampling was set at 9 £ 9 £ 9 for the irreducible Brillouin zone [31].
Fig. 1. Schematic diagram of the preparation of Na-ZnO and pure ZnO for sunlight-photocatalytic degradation of Methylene Blue dye.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111 3
To analyze the electron correlations in transition metal oxides, the into a 250 mL refluxing flask. The mixture was refluxed with mag-
standard DFT calculations based on LDA and GGA produced reason- netic stirring for 2 h. Then after cooling to room temperature, the
able results for structural properties but failed to correctly predict excess K2Cr2O7 was titrated against FAS (0.001 N) using ferroin indi-
the band gaps values because the hybridization between the Zn-d cator. The color altered sharply from bluish-green to reddish-brown
states and the O-p states was not taken into account correctly [32]. at the endpoint of the titration. Using the following Eq. (2) the COD of
Given this, the DFT+U method16 was applied to precisely correct the each sample was calculated:
strong on-site Coulomb repulsion of Zn 3d and O 2p states, the value
ðAB ÞX N X 8000
of U for Zn and O respectively was taken to be 9.3 and 18.4 eV, consis- COD as mgO2 =L ¼ ð2Þ
mL sample
tent with that used in the previous report [33]. Calculations were per-
formed in two steps; firstly, the bulk optimizations of the wurtzite where A and B are the volumes (mL) of FAS used for titrating the
ZnO structure (space group, P63mc) have been carried out to deter- blank and the sample respectively, N = the concentration of FAS
mine the equilibrium geometry. As a next step, the supercell model (equivalent/L) and 8000 = mg O2/equivalent.
of 2 £ 2 £ 2 with 32 atoms (Zn16 O16) was adopted for pure ZnO, as
illustrated in Fig. 1a was built to study the doping effects in ZnO 3. Results and discussion
structure and investigate the electronic properties [34]. This model
represents doping processes by replacing one Zn atom with one Na 3.1. Structure, morphology and optical properties of Na-ZnO
atom, corresponding to approximately 1.86% of Na concentration in photocatalysts
the ZnO host matrix (see Fig. 1).
XRD was used to examine the purity and crystallinity of Na-doped
2.5. Photocatalytic degradation experiments ZnO. As shown in Fig. 2a, all the strong sharp diffraction peaks of the
Na-doped ZnO sample were well indexed to the corresponding hex-
Under similar conditions, all photocatalytic experiments were agonal wurtzite polycrystalline structure (space group: P63mc
done at a specific time on sunny days from 10:00 AM to 6:00 PM (186)). The obtained patterns indicate that the prepared Na-doped
(June-October) using both quartz and borosilicate photoreactor under ZnO was highly crystalline and had a single phase without impurities.
a constant range of temperature from 25 °C to 30 °C. In all experi- The inset in Fig. 2a shows a comparison between Na-doped ZnO and
ments, a particular weight of the photocatalyst was added to a spe- the reference pure ZnO (JCPDS Card No. 361451). As shown, the
cific volume of the dye solutions after adjusting the selected pH using 100, 002, 101 peaks of Na-doped ZnO were shifted towards lower
HANA pH meter, all these mixtures were stirred at 120 rpm in the angles compared with those of pure ZnO. These shifts indicate the
dark for 60 min to reach the equilibrium point between the adsorp- expansion of crystal lattice of ZnO by substituting the Zn by Na
tion and desorption of the dye on the surface of the catalyst. Then, because the ionic radius of Na (0.102 nm) is larger than that of Zn
the photocatalytic experiments started by exposing the suspension (0.074 nm) [25] and the bond length of NaO (0.210 nm) is larger
to the light source and different samples were taken at different time than that of ZnO (0.193 nm) [36]. The N2 adsorption-desorption iso-
intervals for analysis. The photocatalyst was separated from the sam- therm of Na-ZnO is shown in Fig. 2b with the corresponding pore size
ples by centrifugation and the samples were periodically moved for distribution curves (inset in Fig. 2b). The obtained isotherms of Na-
analysis. The effect of several parameters, like dye photolysis (with- ZnO correspond to a type II isotherm in the Brunauer classification
out catalyst), initial dye concentration, pH, catalyst dose, time of day [37] which is characterized by a hysteresis loop and does not exhibit
and type of glass (quartz or borosilicate) of the degradation tube, any limiting adsorption at high relative pressures. The hysteresis
were studied in a similar manner. All other parameters were held loop of Na-ZnO (Fig. 2b) starting from p/po = 0.1 to 0.9 indicates that
constant while the target parameter was varied. Additionally, experi- Na-ZnO contains a high percentage of micropores. The specific BET
ments under similar conditions were performed to compare UV light surface area of Na-ZnO was 29.67 m2/g, and the average pore radius
(20 W low-pressure mercury-vapor lamp model VL-120. G (254 nm)) and total pore volume were 4.07 nm and 0.0761 cc/g, respectively.
with sunlight as the light source. To investigate the main active spe- Fig. 2c shows the typical FTIR spectrum of Na-ZnO. The characteristic
cies in the photocatalytic reaction we used some scavengers (under peaks of Na-ZnO including a band at 578 cm1 potentially correspond
similar conditions) such as tert‑butanol (10 mM) as a hydroxyl radical to the ZnO stretching modes [38], a slightly broadband at 15001600
(¢OH) scavenger, ascorbic acid (1.0 mM ) as a superoxide anion radical cm1 corresponding to the Zn-OH bending mode, and bands at
(O2¢) scavenger and EDTA (1.0 mM) as a hole (h+) scavenger. Also, 11001600 cm1 related to Zn-O. The broadband at
we studied the recyclability of Na-doped ZnO photocatalyst to iden- 29713663 cm1 corresponding to a stretching mode of the OH
tify its stability through 7 consecutive rounds. groups in residual physisorbed water molecules. XPS analysis was
performed to examine the surface elemental composition of Na-ZnO
2.6. Analytical procedures and ZnO (Fig. 2d). The binding energy peaks of all elements were cali-
brated according to the information of C 1 s (284.6 eV) as the refer-
To determine the color removal efficiency, we measured the MB ence [39]. The XPS survey of Na-ZnO and pure ZnO shows peaks
concentration in each analytical sample using a related to Zn, O and C elements, the detected carbon comes from the
UVVis spectrophotometer (Perkin Elmer Lambda 35). Then using adsorbed carbon materials on the surface of the sample. Noteworthy,
Eq. (1), we can calculate the efficiency of decolorization. the Na peak was detected only with the XPS spectra of Na-ZnO, indi-
cating that the Na element is apart from the chemical structure of
C0 Ct
Color removal ð%Þ ¼ x 100 ð1Þ Na-ZnO. The percentage of Na element inside the Na-ZnO sample
C0
was measured by XPS and ICP-OES. The result indicates that the Na%
where C0 and Ct are the MB dye concentration initially and after irra- measured by XPS was 0.65%, while ICP-OES revealed 2.82 wt% of Na
diation time t respectively. (Table S1, S2). This result proved the presence of Na element inside
Using the open reflux method (5220) [35] we can determine the the crystal lattice of ZnO, where the XPS measures the Na% only on
chemical oxygen demand (COD) to investigate the efficiency of the the surface the ICP-OES measures the total Na% of the sample. The
catalyst for the mineralization of dye solutions, which involves the high-resolution spectrum of Zn2p, O 1 s, and Na 1 s are shown in
following procedures: Figure S1.
Pipet 10 mL of sample or blank (distilled water only), 10 mL of The morphology of Na-doped ZnO was characterized by TEM and
concentrated sulphuric acid and 5 mL of K2Cr2O7 (0.05 N) solution SEM. Typical TEM images (Fig. 3a) of Na-ZnO showed that the sample
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
4 M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111
Fig. 2. (a) XRD pattern of Na-doped ZnO: inset is a comparison between some XRD peaks of Na-ZnO and that of pure ZnO (JCPDS Card No. 361451). (b) N2 adsorption-desorption
isotherms of Na-ZnO: the inset shows the pore size distribution of Na-ZnO. (c) FTIR spectrum of Na-doped ZnO. (d) XPS of Na-ZnO and ZnO.
appeared as nanosheet particles that aggregated together to give Na can be attributed to the structural deformation in Na-doped ZnO
non-specific morphology. All Na-ZnO nanosheet particles contained caused due to the change in the lattice constant and the replacement
many pores on their surfaces with an average pore diameter of of either substitutional or interstitial Zn ions in the ZnO lattice by Na
7.99 nm (Fig. 3b) and an average pore radius of 3.99 nm. The average ion. In addition, the correlation between the structural and the opti-
pore radius value detected by TEM (3.99 nm) well agreed with that cal properties suggests that the crystallite size of the ZnO crystal is
obtained from N2 adsorption (4.07 nm). The selected area electron predominantly influenced by the band gap energy [40]. The Valence
diffraction (SAED) pattern of the Na-doped ZnO nanoparticles bands (EVB) of both samples were obtained from a photoelectron
(Fig. 3c) revealed clear bright spots, which indicated the high crystal- spectrometer method using the plots of yield (cps0.5) versus the inci-
linity of the sample. The interplanar distances calculated from SAED dent photon energy (eV), significantly we can get the values by plot-
were 1.95 A, 1.91 A, 1.60 A, and 1.18 A, which is consistent with that ting the linear portion of the curve to intersect the photon energy x-
obtained from the XRD pattern. Consistent with the TEM observa- axis and all related plots were shown in Fig. 4c and d. The conduction
tions, the morphology of Na-doped ZnO can be characterized by SEM band (ECB) was calculated from EVB Eg and the values are summa-
(Fig. 3d). rized in Table S1.
The optical properties of Na-ZnO and pure ZnO nanoparticles
were examined using a UVVis spectrophotometer. As shown in 3.2. Theoretical calculation
Fig. 4a, the doping of ZnO with Na is slightly red-shifted in the
absorption spectra as compared with those of pure ZnO. The red The bulk optimizations of the wurtzite ZnO structure has been
shifting of the absorption peak of Na-ZnO leads to a slightly narrow carried out before the calculations of the electronic structures to find
bandgap (Eg) which enhances the visible light-harvesting ability of the lattice parameters with the minimum energy (Fig. 5a and b), as a
the semiconductor. The optical bandgap values of Na-ZnO and ZnO general result, the lattice constant values obtained at the PBE level
are calculated via UVVis spectroscopy using Tauc Plots (Fig. 4b) to for the bulk ZnO (a = b = 3.282 A and c = 5.301 A) were slightly higher
be 3.13 and 3.25 eV respectively. Table S3 illustrates the crystallite than the experimental results of a = b = 3.254 A and c = 5.215 A, and
size and lattice parameters of ZnO and Na-ZnO. The lattice constants the deviation is less than 1.6%, but were well compatible with the
of Na-ZnO have larger values than those of pure ZnO because the previously reported first-principles theoretical results [41]. Simulated
ionic radius of Na+ (0.102 nm) is relatively larger than those of Zn+2 XRD data were compared to identify the crystal structure and deter-
(0.074 nm). It is worth mentioning here that the crystallite size for mine differences due to the doping effect. In the XRD pattern of the
ZnO sample decreases with the introduction of Na dopant ions. The Na-doped ZnO sample (Fig. 5d), all peaks observed were related to
band gap of the Na-doped sample is observed to shrink from 3.25 to the lattice planes (100), (002), (101), (102), (110), (103), (200), (112)
3.13 eV as the radius of the dopant ion (0.102 nm) larger than those and (201). These peaks are in good agreement with the experimental
of Zn+2 (0.074 nm). The decrement in the band gap after doping by data (Fig. 2a) and correspond to the hexagonal wurtzite of ZnO [42].
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111 5
Fig. 3. (a) TEM image, (b) high resolution TEM image, (c) SEAD pattern of Na-ZnO. (d) SEM of Na-ZnO. .
However, the estimated peaks deviated slightly from the experimen- the bandgap which agrees with the reported for alkali-doped ZnO nano-
tal data by 1.4%. It was found that the correlation coefficient (R2) structures [48, 49]. As a result of Na doping, the bandgap was signifi-
between the position of the calculated peaks and those experimen- cantly reduced by around 0.16 eV in agreement with our experimental
tally observed was 0.99997. The observed and calculated XRD peaks observations (0.12 eV). The detailed analysis of the DOS for the pure
for ZnO and Na-doped ZnO listed in Table S4. Fig. 5c and d, show the ZnO model is shown in Fig. 6c to understand the change in the elec-
XRD patterns for the simulated ZnO and Na-doped ZnO. It is worthy tronic structure. The calculated PDOS of orbital contribution for each
to mention that the diffraction peaks of Na-doped ZnO shifted to element (Zn and O) indicates that the valence band (VB) mainly consists
lower angles than those in pure ZnO (Table S4), which is in good of Zn-3d and O-2p orbital levels in the range of (8 to 6.7 eV) and
agreement with our experimental results. The explanation for the left (3.9 to 0 eV) respectively, which is consistent with the other report
shift in Fig. 5d is due to the substitution of Na+1 (having higher ionic [50] with a small contribution of the Zn-3p orbital level. On the other
radii) at the Zn2+lattice [43, 44]. hand, the CB is mainly composed of the O-2p and Zn-4 s hybrid states
In order to investigate the effect of Na doping on the electronic struc- [51] with minimal Zn-3p orbital contribution. Because 3d states are
ture of ZnO and to study the doping mechanism, we calculated the total involved in the valence band, O and Zn have a strong bonding character
density of states (TDOS), and the partial density of states (PDOS) of Na- [52]. Moreover, the partial DOS for the Na-doped ZnO model, as shown
doped ZnO in comparison with pure ZnO sample, see Figs. 6a and b. The in Fig. 6d, reveals significant variations in both VB and CB. According to
GGA bandgap energy (Eg) values for undoped ZnO was about 0.71 eV, in these results, the VB of Na-doped ZnO was dominated by the state of
accordance with the previously reported results of 0.73 and 0.74 eV in Zn-3d orbital along with the state of O-2p. Besides, it can be clearly seen
Ref. [45, 46], but still lower compared to our experimental value from Fig. 6d that the Na atom contributed strongly in the VB with the
(3.11 eV). The exchange-correlation functional GGA+U has been used to 2p orbital level, as well as the hybridization of O-2p and Na-2p is domi-
avoid the underestimation of the bandgap energy. The calculated GGA nant near the Fermi level. Furthermore, the Fermi level shifts downward
+U bandgap energy values for ZnO raised to 2.49 eV (only 0.76 eV less to the valence band compared to the pure ZnO, indicating it is a p-type
than the experimental value), which significantly improved the pre- dopant [53].
dicted bandgap concerning the GGA result and was in good agreement
with the reported value (2.49 eV) from the hybrid functional calculation 3.3. Photocatalytic activity
[47]. However, our discussions regarding the energy gap will not be
affected, as only relative changes in the bandgap energy (ΔEg) were The photocatalytic efficiency of the prepared samples was esti-
caused for concern. In our simulation, the calculated bandgap (Eg) of mated by photocatalytic degradation of MB dye under direct sunlight.
pure ZnO was 2.49 eV (Fig. 6a), whereas for Na-doped ZnO the band gap Firstly, the hydrolysis (in the dark without catalyst) and photolysis
was 2.33 eV (Fig. 6b), suggesting that the Na dopant in ZnO can reduce (under light without catalyst) of MB dye was investigated at different
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
6 M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111
Fig. 4. Optical absorbance spectra (a) and optical band gap energy of (b) pure ZnO and Na-doped ZnO. The valence bands of Na-ZnO (c) and ZnO (d) were determined by the photo-
electron spectrometer.
pH values [5, 79]. After keeping the MB solutions in the dark for to 9. An increase in the rate of direct photolysis of MB in alkaline
30 min, the results indicate that the removal of MB by hydrolysis or medium was also reported [54]. Fig. 7b shows a comparison between
adsorption to the glass walls was negligible at all selected pH values pure ZnO, commercial TiO2 (P25) (as a popular standard), and Na-
(Fig. 7a). Thereafter, the dye solutions were exposed to direct sunlight ZnO for the photocatalytic degradation of 30 ppm MB at pH 8 under
irradiation. As shown in Fig. 7a, the photolysis of MB in acidic sunlight irradiation. To reach the adsorption equilibrium of the dye
medium (pH 5) was minimal while it slightly increased at pH from 7 onto the photocatalyst surface, the experiments were maintained in
Fig. 5. Crystal structures of (a) ZnO and (b) Na-doped ZnO (2 £ 2 £ 2) supercells. Simulated XRD for (c) ZnO and (d) Na-doped ZnO.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111 7
Fig. 6. Total density of states (TDOS) of ZnO (a) and Na-doped ZnO (b). Partial density of states (PDOS) of ZnO (c) and Na-doped ZnO (d), the dashed vertical line represents the posi-
tion of the Fermi level at E = 0.
the dark for 60 min. The removal of MB by dark adsorption was only catalyst), photolysis and catalyst-dye samples were exposed to direct
7.5% by ZnO, 15.4% by TiO2, and 9.7% by Na-ZnO. After dark adsorp- sunlight. After 90 min of the photocatalytic experiments, the blank
tion equilibrium was reached, the blank (in the dark without sample was approximately stable, while 16%, 35%, 79.8%, and 99.5%
Fig. 7. (a) Hydrolysis and photolysis of MB dye at different pH values under sunlight irradiation ([MB] = 30 ppm). (b) adsorption and photocatalytic degradation of MB over the two
photocatalysts (catalyst dose = 0.5 g/L, pH = 8, [MB] = 30 ppm). (c) Photocurrent transient response spectra, and (d) Photoluminescence emission spectra of ZnO and Na-ZnO.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
8 M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111
of MB were removed by photolysis, pure ZnO, TiO2, and Na-doped photocatalytic degradation reaction of MB using Na-ZnO were well
ZnO respectively. The results indicated that the doping of ZnO by Na represented by the pseudo-first-order model, with correlation coeffi-
element enhanced the photocatalytic efficiency of ZnO and made it a cients R2 0.990 for all initial MB concentrations (Table S6). Further-
promising photocatalyst for the removal of organic pollutants from more, there was a good agreement between the experimental and
wastewater and better than the commercial TiO2 (P25) [55]. To calculated values of the initial concentrations (Co). From Table S6, the
understand more about the photocatalytic efficiency enhancement of reaction rate constant decreased with an increasing initial dye con-
Na-ZnO, the photoresponse ability of ZnO and Na-ZnO were mea- centration.
sured using the transient photocurrent response under simulated The UVvis spectra of MB in the aqueous solution obtained in
sunlight irradiation at 1.2 V (vs. Ag/AgCl). The results shown in Fig. 7c scanning mode (from 200 to 800 nm) at different time points during
indicate that the doping of ZnO by Na element enhanced the photo- the photocatalytic degradation reaction is shown in Fig. 8a. The
current response many times than pure ZnO. This is evidence that absorption peaks between 600 and 700 nm can be attributed to a
the Na-ZnO has high and good sensitivity to the light illumination chromophore containing a long conjugated p system, while the
more than pure ZnO, which corresponds to the result of photocata- absorption peaks at 292 and 246 nm can be related to the aromatic
lytic degradation of MB. Furthermore, the photoluminescence spectra rings [54]. According to Fig. 8a, the abovementioned peaks decrease
obtained from ZnO strongly quenched after doping by Na element, as considerably over the course of irradiation and finally disappear
shown in Fig. 7d. This indicates that the Na doping process sup- before 60 min in the presence of Na-ZnO under sunlight irradiation.
presses the electron-hole recombination during the photocatalytic These results only suggested the decolorization of the MB dye during
reaction, because of the new energy level generated by Na element the photocatalytic reaction, while the full degradation (mineraliza-
which enhances the electrons transfer before the recombination hap- tion) of the dye solution was measured by determining the chemical
pens [56]. In addition, the comparison of photodegradation of the oxygen demand (COD) of the MB dye during the photocatalytic reac-
methylene blue dye (MB) using Na-doped ZnO photocatalyst with tion. Fig. 8b shows that COD% of the dye solution decreased from
other recently reported photocatalysts are illustrated in Table S5. 100% to 1.7% over 3.0 h of photocatalytic reaction using Na-ZnO
Different factors were studied to investigate the optimum condi- under sunlight. This result indicated the high photocatalytic activity
tions of the photocatalytic degradation experiments using Na-ZnO, of Na-ZnO for mineralization of MB dye under sunlight within a short
for example, pH, catalyst dose and daytime. As shown in Figure S2, irradiation period. The high stability of the prepared photocatalyst
after 60 min of the photocatalytic experiment approximately 98% of and the high photocatalytic degradation efficiency over multiple
10 ppm of MB were degraded at all selected pH values ([5, 78] and cycles increased the economic feasibility of its application. To investi-
[9]), which indicates the promising photocatalytic activity of Na-ZnO gate the recyclability of Na-ZnO, the following experiment was con-
at different pH mediums (acidic, neutral and alkaline). Further ducted. First, 20 mL of MB solution at pH 8 containing 10 mg of
experiments were performed at pH 8 to simulate the pH of aqueous catalyst was exposed to sunlight for 3.0 h. Then, the photocatalyst
solutions under environmental conditions. The photocatalyst dose is was collected by centrifugation and reused without washing or dry-
an important parameter that has been extensively studied. Increasing ing in a new 20 mL of MB solution. This process was repeated up to
the catalyst dose from 0.25 to 0.5 g/L was accompanied by a sharp 7 times. While studying the reuse of the photocatalyst, all parame-
increase of the degradation of MB from 49% to 81.6% after 15 min ters, including the irradiation time, initial MB concentration and pH
(Figure S3a), this was due to the available active sites increased with of the medium, were kept constant. The results in Fig. 8c show that
increasing the amount of the photocatalyst [57]. However, an MB dye was completely decolorized (more than 99%) with all 7
increase in the catalyst dose from 0.5 to 3.0 g/L slightly raised the per- recycles. Fig. 8d shows the mineralization of MB in each recycles. In
centage of dye degradation, because the large amount of Na-ZnO acts the first recycle, the%COD remaining was 1.7% in the reaction period.
as a shield and blocks the penetration of light into some catalyst mol- Increasing the number of recycles to 7 was companied by a slight
ecules. Therefore, the optimal dose of Na-ZnO was determined as increase for the%COD remaining to 19.6%. Increasing the% COD
0.5 g/L. The photocatalytic degradation of MB using Na-ZnO was per- remaining may be due to the Na ion leaching of Na-ZnO or the oxida-
formed under sunlight at four different times: 9:40 am, 2:45 pm, tive intermediates formed by the decomposition of MB dye in the
4:45 pm, and 6:20 pm to investigate the effect of daytime on the pho- solution which could be adsorbed on the surface of the photocatalyst
tocatalytic reaction. As shown in Figure S3b, the best performance of that influenced the light penetration to the surface thus resulted rela-
Na-ZnO was detected at 2:45 pm where the sunlight was vertical. tively decrease of Na-ZnO activity and increase in the%COD remain-
While the efficiency decreased gradually with increasing the daytime ing [58]. These results indicate that the prepared Na-ZnO
until it reached about 30% at 6:20 pm (at sunset). Figure S4 illustrates photocatalyst has high efficiency and stability for the mineralization
the difference between the quartz photoreactors and the borosilicate of organic pollutants in the water.
used in the photocatalytic reaction, where the borosilicate glass pre- The separation of electrons and holes is the initial step of the
vented the UV portion of the sunlight from reaching the photocata- mechanism of the photocatalytic degradation reaction. The primary
lytic reaction in contrast to the quartz glass which allowed its active species formed as a result of the photoexcitation of the elec-
passage. The rates of the photocatalytic degradation of MB are trons of the semiconductors are hydroxyl radical ( ¢ OH) and superox-
approximately the same with the two different photoreactors, which ide anion radicals ( ¢ O2 ) [59]. The effect of different scavengers was
indicates that the UV portion of the sunlight is not the main reason studied to determine the active spices produced during the photoca-
for the photocatalytic reaction. talytic degradation reaction. The selected scavengers in this study
Non-linear regression (using origin 8.6 software) of a pseudo- were ascorbic acid (1 mM) for ¢ O2 elimination, EDTA (1 mM) for
first-order model was used to investigate the kinetics of the photoca- hole (h+) elimination, and tert‑butyl alcohol (t-BuOH) (10 mM) for ¢
talytic degradation reaction of MB dye using Na-ZnO at different dye OH elimination. Fig. 9a shows that the removal of ¢ OH by t-BuOH
concentrations (8, 14, 31 and 44 mg/L), as shown in Figure S5a. The slightly inhibits the photocatalytic degradation reaction compared to
non-linear form for the pseudo-first-order model is represented by that without adding scavengers, while the removal of ¢ O2 by ascorbic
Eq. (3): acid was accompanied by a considerable inhibition of the photocata-
lytic degradation reaction compared to that without adding scav-
C ¼ Co eKt ð3Þ
engers. The photocatalytic removal of MB in the presence of ascorbic
where Co is the initial dye concentration, C is the dye concentration at acid was strongly inhibited during the first 30 min of the reaction
time t of the reaction and K is the rate constant of the first-order reac- while it increased after 30 min because the ascorbic acid itself was
tion (min1). The results shown in Figure S5b indicated that the degraded by Na-ZnO, as shown from UVvis absorption spectra of
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111 9
Fig. 8. (a) UVvis spectra of MB and (b) COD% of the MB dye during the photocatalytic reaction. (c) Decolorization and (d) mineralization of MB dye during the recyclability of Na-
ZnO photocatalyst ([MB] = 10 mg/L, pH 8, dose = 0.5 g/L and reaction time = 3 h).
ascorbic acid during the photocatalytic reaction (Fig. 9b). This result Na-ZnO happened through ¢ O2 and ¢ OH active spices produced by
indicated that the superoxide anion radical ( ¢ O2 Þwas the main active photoexcitation of electrons of Na-ZnO under sunlight.
species, while the ¢ OH partially participated in the photocatalytic From the results of our study and the literature, we suggest a
degradation of MB. Similar results were reported previously by Tang mechanism for the photocatalytic degradation of MB using Na-ZnO
et al. [60]. The addition of EDTA during the photocatalytic degrada- photocatalyst under sunlight irradiation (Fig. 10). Under sunlight
tion reaction as a hole (h+) scavenger increased the rate of the reac- irradiation, when the photocatalyst is illuminated by the energy of
tion, as shown in Fig. 9a. Because the elimination of holes (h+) by photons (hn) higher than (or equal to) its band-gap, electrons excited
EDTA enhanced the electron-hole charge separation and quenched from the valence band (VB) to the conduction band (CB) resulting in a
the electron-hole recombination of the photocatalyst. It can be con- hole in the VB and electron in the CB [61]. The electron in CB is
cluded that the photocatalytic degradation of MB using the prepared trapped by Na dopant which reduces the recombination rate of the
Fig. 9. (a) Photocatalytic degradation of MB dye using Na-ZnO under sunlight in the presence of various scavengers ([MB] = 10 mg/L, [ascorbic acid] = [EDTA] = 1 Mm, [t-
BuOH] = 10 mM and pH of reaction = 8) and (b) UVvis spectrum of MB dye samples produced from photodegradation in the presence of ascorbic acid.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
10 M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111
Fig. 10. The mechanism of the photocatalytic degradation reaction of MB dye using Na-doped ZnO under sunlight.
photo-generated electrons and holes. The electrons in CB or those [3] Jayakumar J, Chou HH. Recent advances in visiblelightdriven hydrogen evolu-
trapped by Na dopant can be transferred to electronic acceptors like tion from water using polymer photocatalysts. Chem. Cat. Chem. 2020;12
(3):689–704.
adsorbed O2 leading to the production of superoxide anion radicals [4] Chang CL, Lin WC, Jia CY, Ting LY, Jayakumar J, Elsayed MH, Yang YQ, Chan YH,
that can produce hydroxyl radical or react directly with the organic Wang WS, Lu CY, Chen PY, Chou HH. Low-toxic cycloplatinated polymer dots
dye. On the other hand, the hole can trap electronic donors such as with rational design of acceptor co-monomers for enhanced photocatalytic effi-
ciency and stability. Appl. Catal. B. 2020;268:118436.
H2O or HO generating reactive hydroxyl radicals that react with the [5] Ting LY, Jayakumar J, Chang CL, Lin WC, Elsayed MH, Chou HH. Effect of control-
dye. All the produced radicals convert the organic dye to dye radicals ling the number of fused rings on polymer photocatalysts for visible-light-driven
and so on until the organic compound degrades to CO2 and H2O. hydrogen evolution. J. Mater. Chem. A. 2019;7(40):22924–9.
[6] Qamar MT, Aslam M, Ismail IMI, Salah N, Hameed A. Synthesis, characterization,
and sunlight mediated photocatalytic activity of CuO Coated ZnO for the removal
4. Conclusion of nitrophenols. ACS Appl. Mater. Interfaces. 2015;7(16):8757–69.
[7] Yang T, Peng J, Zheng Y, He X, Hou Y, Wu L, Fu X. Enhanced photocatalytic ozona-
tion degradation of organic pollutants by ZnO modified TiO2 nanocomposites.
In conclusion Na-ZnO photocatalyst was developed, characterized Appl. Catal. B. 2018;221:223–34.
and its catalytic activity towards decolorization and mineralization of [8] Hu C, Hu X, Li R, Xing Y. MOF derived ZnO/C nanocomposite with enhanced
MB under sunlight was determined. The prepared Na-ZnO showed a adsorption capacity and photocatalytic performance under sunlight. J. Hazard.
Mater. 2020;385:121599.
morphology of nanosheet particles. The results showed that degrada- [9] Zhang G, Chen D, Li N, Xu Q, Li H, He J, Lu J. Fabrication of Bi2MoO6/ZnO hierarchi-
tion of MB occurred at a sufficient level in alkaline, acidic, and neutral cal heterostructures with enhanced visible-light photocatalytic activity. Appl.
media with a catalyst dose of 0.5 g/L. The photodegradation process Catal. B. 2019;250:313–24.
[10] Elmorsi T, Elsayed M, Bakr MJAJC. Na doped ZnO nanoparticles assisted photoca-
followed first-order reaction kinetics. In recyclability tests, Na-ZnO talytic degradation of congo red dye using solar light. J. Chem Eng & Process Tech.
was very stable and could mineralize 80.4% of the MB dye after 7 2017;7:48–57.
recycles. [11] Kim K-J, Kreider PB, Choi C, Chang C-H, Ahn H-G. Visible-light-sensitive Na-doped
p-type flower-like ZnO photocatalysts synthesized via a continuous flow micro-
reactor. RSC Adv 2013;3(31):12702–10.
Declaration of Competing Interest [12] Davis K, Yarbrough R, Froeschle M, White J, Rathnayake H. Band gap engineered
zinc oxide nanostructures via a solgel synthesis of solvent driven shape-con-
trolled crystal growth. RSC Adv 2019;9(26):14638–48.
The authors declare no conflict of interest [13] Kamarulzaman N, Kasim MF, Rusdi R. Band gap narrowing and Widening of ZnO
nanostructures and doped materials. Nanoscale Res. Lett. 2015;10(1):346.
Acknowledgments [14] Ta QTH, Cho E, Sreedhar A, Noh JS. Mixed-dimensional, three-level hierarchical
nanostructures of silver and zinc oxide for fast photocatalytic degradation of mul-
tiple dyes. J Catal 2019;371:1–9.
The authors gratefully acknowledge the financial support of the [15] Xie H, Ding F, Mu H. Effects of Au nanoparticles and ZnO morphology on the pho-
Ministry of Science and Technology of Taiwan (MOST 1092636-E- tocatalytic performance of Au doped ZnO/TiO2 films. Nanotechnology 2019;30
(8):085708.
007021 and MOST 10826228007016). And the authors [16] Height MJ, Pratsinis SE, Mekasuwandumrong O, Praserthdam P. Ag-ZnO catalysts
appreciate the Precision Instrument Support Center of National Tsing for UV-photodegradation of methylene blue. Appl. Catal. B. 2006;63(3):305–12.
Hua University in providing the analysis and measurement facilities. [17] Hsu M-H, Chang C-J. Ag-doped ZnO nanorods coated metal wire meshes as hier-
archical photocatalysts with high visible-light driven photoactivity and photo-
stability. J. Hazard. Mater. 2014;278:444–53.
Supplementary materials [18] Qi K, Xing X, Zada A, Li M, Wang Q, S-y Liu, Lin H, Wang G. Transition metal doped
ZnO nanoparticles with enhanced photocatalytic and antibacterial performances:
experimental and DFT studies. Ceram. Int. 2020;46(2):1494–502.
Supplementary material associated with this article can be found [19] Sukriti Chand P, Singh V. Enhanced visible-light photocatalytic activity of samar-
in the online version at doi:10.1016/j.jtice.2020.10.018. ium-doped zinc oxide nanostructures. J. Rare Earths. 2020;38(1):29–38.
[20] Zhang Q, Liu J-K, Wang J-D, Luo H-X, Lu Y, Yang X-H. Atmospheric Self-induction
Synthesis and Enhanced Visible Light Photocatalytic Performance of Fe3+ Doped
References Ag-ZnO Mesocrystals. Ind. Eng. Chem. Res. 2014;53(34):13236–46.
[21] Vaiano V, Iervolino G, Rizzo L. Cu-doped ZnO as efficient photocatalyst for
[1] Anjaneyulu Y, Sreedhara Chary N, Samuel Suman Raj D. Decolourization of indus- the oxidation of arsenite to arsenate under visible light. Appl. Catal. B.
trial effluents - available methods and emerging technologies - a review. Rev. 2018;238:471–9.
Environ. Sci. Biotechnol. 2005;4(4):245–73. [22] Basyooni MA, Shaban M, El Sayed AM. Enhanced gas sensing properties of spin-
[2] Reddy IN, Reddy CV, Shim J, Akkinepally B, Cho M, Yoo K, Kim D. Excellent visible- coated Na-doped ZnO nanostructured films. Sci Rep 2017;7(1):41716.
light driven photocatalyst of (Al, Ni) co-doped ZnO structures for organic dye deg- [23] Liu W, Xiu F, Sun K, Xie Y-H, Wang KL, Wang Y, Zou J, Yang Z, Liu J. Na-Doped
radation. Catal. Today 2020;340:277–85. p-Type ZnO Microwires. J. Am. Chem. Soc 2010;132(8):2498–9.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018
JID: JTICE
ARTICLE IN PRESS [m5G;November 4, 2020;17:59]
M.H. Elsayed et al. / Journal of the Taiwan Institute of Chemical Engineers 00 (2020) 111 11
[24] Wu C, Huang Q. Synthesis of Na-doped ZnO nanowires and their photocatalytic [43] Laurenti M, Castellino M, Perrone D, Asvarov A, Canavese G, Chiolerio A. Lead-free
properties. J. Lumin. 2010;130(11):2136–41. piezoelectrics: v 3+ to V 5+ ion conversion promoting the performances of V-doped
[25] Ye Z, Wang T, Wu S, Ji X, Zhang Q. Na-doped ZnO nanorods fabricated by chemical Zinc Oxide. Sci. Rep. 2017;7:41957.
vapor deposition and their optoelectrical properties. J. Alloys Compd. [44] Ghosh M, Raychaudhuri A. Structure and optical properties of Cd-substituted ZnO
2017;690:189–94. (Zn1 xCdxO) nanostructures synthesized by the high-pressure solution route.
[26] Akcan D, Gungor A, Arda L. Structural and optical properties of Na-doped ZnO Nanotechnology 2007;18(11):115618.
films. J. Mol. Struct. 2018;1161:299–305. [45] Schleife A, Fuchs F, Furthmu € ller J, Bechstedt F. First-principles study of ground-
[27] Ghosh S, Khan GG, Varma S, Mandal K. Influence of film thickness and oxygen and excited-state properties of MgO, ZnO, and CdO polymorphs. Phys. Rev. B.
partial pressure on cation-defect-induced intrinsic ferromagnetic behavior in 2006;73(24):245212.
luminescent p-Type Na-Doped ZnO thin films. ACS Appl. Mater. Interfaces. [46] Xia C, Wang F, Hu C. Theoretical and experimental studies on electronic structure
2013;5(7):2455–61. and optical properties of Cu-doped ZnO. J. Alloys Compd. 2014;589:604–8.
[28] Kresse G, Furthmu € ller J. Efficient iterative schemes for ab initio total-energy cal- [47] Luo JH, Liu Q, Yang LN, Sun ZZ, Li ZS. First-principles study of electronic structure
culations using a plane-wave basis set. Phys. Rev. B. 1996;54(16):11169–86. and optical properties of (ZrAl)-codoped ZnO. Comput. Mater. Sci.
[29] Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made sim- 2014;82:70–5.
ple. Phys. Rev. Lett. 1996;77(18):3865–8. [48] Peyghan AA, Noei M. The alkali and alkaline earth metal doped ZnO nanotubes:
[30] Ernzerhof M, Scuseria GE. Assessment of the Perdew-Burke-Ernzerhof exchange- DFT studies. Phys. B Condens. Matter. 2014;432:105–10.
correlation functional. J. Chem. Phys. 1999;110(11):5029–36. [49] Baei MT, Peyghan AA, Bagheri Z. First Principles Study on Encapsulation of Alkali
[31] Monkhorst HJ, Pack JD. Special points for Brillouin-zone integrations. Phys. Rev. B. Metals into ZnO Nanocage. Chinese J. Chem. Phys. 2012;25(6):671–5.
1976;13(12):5188–92. [50] Lee Y-S, Peng Y-C, Lu J-H, Zhu Y-R, Wu H-C. Electronic and optical properties of
[32] Janotti A, Segev D, Van de Walle CG. Effects of cation d states on the structural and Ga-doped ZnO. Thin Solid Films 2014;570:464–70.
electronic properties of III-nitride and II-oxide wide-band-gap semiconductors. [51] Li P, Deng SH, Li YB, Huang J, Liu GH, Zhang L. Aluminum and nitrogen impurities
Phys. Rev. B. 2006;74(4):045202. in Wurtzite ZnO: first-principles studies. Phys. B Condens. Matter. 2011;406
[33] Bashyal K, Pyles CK, Afroosheh S, Lamichhane A, Zayak AT. Empirical optimization (17):3125–9.
of DFT + U and HSE for the band structure of ZnO. J. Phys. Condens. Matter. [52] La Porta FA, Andre s J, Vismara MVG, Graeff CFO, Sambrano JR, Li MS, Varela JA,
2018;30(6):065501. Longo E. Correlation between structural and electronic orderdisorder effects and
[34] Deng S, Duan M, Xu M, He L. Effect of La doping on the electronic structure and optical properties in ZnO nanocrystals. J. Mater. Chem. C. 2014;2(47):10164–74.
optical properties of ZnO. Physica B Condens. Matter. 2011;406(11):2314–8. [53] Bousmaha M, Bezzerrouk M, Baghdad R, Chebbah K, Kharroubi B, Bouhafs B. Real-
[35] Elmorsi TM, Riyad YM, Mohamed ZH, Abd El Bary HMH. Decolorization of Mor- ization of p-Type Conductivity in ZnO via Potassium Doping. Acta Phys. Polonica
dant red 73 azo dye in water using H2O2/UV and photo-Fenton treatment. J. Haz- A 2016;129(6):1155–8.
ard. Mater. 2010;174(1):352–8. [54] Soltani T, Entezari MH. Photolysis and photocatalysis of methylene blue by ferrite
[36] Qiu Z, Yang X, Han J, Zhang P, Cao B, Dai Z, Duan G, Cai W. SodiumDoped ZnO bismuth nanoparticles under sunlight irradiation. J Mol Catal A Chem 2013;377:
nanowires grown by highpressure PLD and their acceptorrelated optical prop- 197–203.
erties. J. Am. Ceram. Soc. 2014;97(7):2177–84. [55] Lee KM, Lai CW, Ngai KS, Juan JC. Recent developments of zinc oxide based
[37] Do A-TT, Giang HT, Do TT, Pham NQ, Ho GT. Effects of palladium on the optical photocatalyst in water treatment technology: a review. Water Res
and hydrogen sensing characteristics of Pd-doped ZnO nanoparticles. Beilstein J. 2016;88:428–48.
Nanotechnol. 2014;5(1):1261–7. [56] Carey JJ, McKenna KP. Screening doping strategies to mitigate electron trapping at
[38] Djaja NF, Montja DA, Saleh R. The effect of Co incorporation into ZnO nanopar- anatase TiO2 surfaces. J. Phys. Chem. C. 2019;123(36):22358–67.
ticles. Adv. Mater. - Phys. Chem. 2013;3:33–41. [57] Yang L, Liya EY, Ray MB. Degradation of paracetamol in aqueous solutions by TiO
[39] Han D, Li B, Yang S, Wang X, Gao W, Si Z, Zuo Q, Li Y, Li Y, Duan QJN. Engineering 2 photocatalysis. Water Res 2008;42(13):3480–8.
charge transfer characteristics in hierarchical Cu2S QDs@ ZnO Nanoneedles with [58] Reddy NL, Emin S, Valant M, Shankar MV. Nanostructured Bi2O3@ TiO2 photocata-
pn heterojunctions: towards highly efficient and recyclable photocatalysts. lyst for enhanced hydrogen production. Int. J. Hydrog. Energy. 2017;42(10):
Nanomaterials 2019;9(1):16. 6627–36.
[40] Basyooni MA, Shaban M, El Sayed AM. Enhanced Gas Sensing Properties of Spin- [59] Upreti AR, Li Y, Khadgi N, Naraginti S, Zhang C. Efficient visible light photocata-
coated Na-doped ZnO Nanostructured Films. Sci. Rep. 2017;7:41716. lytic degradation of 17a-ethinyl estradiol by a multifunctional AgAgCl/ZnFe2O4
[41] Fonseca AFVd, Siqueira RL, Landers R, Ferrari JL, Marana NL, Sambrano JR, La Porta magnetic nanocomposite. RSC Adv 2016;6(39):32761–9.
FdA, Schiavon MA. A theoretical and experimental investigation of Eu-doped ZnO [60] Tang CY, Yang MK, Wu FY, Zhao H, Pang YJ, Yang RW, Lu GH, Yang YH. Identifica-
nanorods and its application on dye sensitized solar cells. J. Alloys Compd. tion of miRNAs and their targets in transgenic Brassica napus and its acceptor
2018;739:939–47. (Westar) by high-throughput sequencing and degradome analysis. RSC Adv.
[42] Radha B, Rathi R, Lalithambika KC, Thayumanavan A, Ravichandran K, Sriram S. 2015;5(104):85383–94.
Effect of Fe doping on the photocatalytic activity of ZnO nanoparticles: experi- [61] Mota A, Albuquerque L, Beltrame LC, Chiavone-Filho O, Machulek Jr A, Nasci-
mental and theoretical investigations. J. Mater. Sci. Mater. Electron. 2018;29 mento C. Advanced oxidation processes and their application in the petroleum
(16):13474–82. industry: a review. Braz J. Petrol Gas 2009;2(3):122–42.
Please cite this article as: M.H. Elsayed et al., Direct sunlight-active Na-doped ZnO photocatalyst for the mineralization of organic pollutants at
different pH mediums, Journal of the Taiwan Institute of Chemical Engineers (2020), https://doi.org/10.1016/j.jtice.2020.10.018