Materials Science in Semiconductor Processing: M. Nami, S. Sheibani, F. Rashchi

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Materials Science in Semiconductor Processing 135 (2021) 106083

Contents lists available at ScienceDirect

Materials Science in Semiconductor Processing


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

Photocatalytic performance of coupled semiconductor ZnO–CuO


nanocomposite coating prepared by a facile brass anodization process
M. Nami, S. Sheibani *, F. Rashchi
School of Metallurgy and Materials Engineering, College of Engineering, University of Tehran, Tehran, Iran

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

Keywords: Coupled semiconductor metallic oxides have been largely taken into consideration due to the photocatalytic
Photocatalyst application. In this work, ZnO–CuO coating nanocomposite was fabricated by simple anodic oxidation of alpha
ZnO brass at different anodizing times. The crystalline structure of hexagonal ZnO and monoclinic CuO phases were
CuO
demonstrated by X-ray diffraction (XRD). Field emission scanning electron microscopy (FESEM) and trans­
Anodizing
Coating
mission electron microscopy (TEM) observation indicated the heterostructure formation of ZnO flower-like and
CuO patches-like after 30 min anodizing. Brunauer-Emmett-Teller (BET) method was conducted for specific
surface area measurement of the samples. Cyclic voltammetry (CV) and Pourbiax diagram indicated that the
brass oxidation process starts with the formation of ZnO followed by CuO formation. Diffuse reflectance spec­
troscopy (DRS) demonstrated that the band gap energy of the ZnO–CuO nanocomposite narrowed by raising
anodizing time owing to the formation of the Cu(3d) energy state near the Zn (3d) valence band. Photo­
luminescence (PL) spectra revealed that the lifetime of the photo-excited electron-hole increased by the effective
combination between CuO and ZnO in anodized brass after 30 min. The photocatalytic performance of ZnO–CuO
nanocomposites with different anodization times was estimated by the degradation of 2 mg/L methylene blue
(MB) and colorless phenol solution under visible light irradiation. After 30 min, the anodized brass with an
estimated 3 eV band gap energy exhibited 54 and 49% degradation of MB and phenol, respectively, after 300
min, which was the highest efficiency among all the samples. Also, the highest photo-degradation was obtained
at 8 mm2/mL of surface area to volume of dye solution. Recyclability results showed only a 5% drop-in pho­
tocatalytic activity after three cycles. Finally, the photo-degradation mechanism was proposed based on the role
of holes and hydroxyl radicals, using various scavengers.

1. Introduction Various semiconductors such as TiO2, ZnO, CuO, CdS, V2O5, WO3,
and their nanostructured assemblies have been studied in photocatalytic
Water treatment and elimination or conversion of toxic pollutants applications [11–16]. ZnO is an intrinsic n-type semiconductor and an
into non-toxic compounds from water resources have become more and excellent candidate for being applied as a photocatalyst, due to its
more crucial due to the ever-increasing industrial activities [1–3]. exceptional properties such as environmental durability, low-cost,
Applying low-price and non-toxic semiconductor materials in photo­ non-toxicity, suitable electron mobility, high photo-activity, and fine
catalytic applications is an effective strategy to deal with wastewater redox properties [17–19]. Furthermore, ZnO can be synthesized through
issues [4–6]. In a photocatalytic system, the light energy exposed to the a large number of techniques including sol-gel, hydrothermal, chemical
surface of a semiconductor can be turned into chemical energy, using bath deposition, sol-electrophoretic deposition, molecular beam
photo-excited electron-hole pairs [7]. The photo-generated charge car­ epitaxy, and chemical vapor deposition with various morphologies and
riers participate in several consecutive redox reactions to produce structures [20–25]. However, the photocatalytic properties of ZnO have
reactive oxygen species including hydroxyl radical (•OH), hydrogen two significant weaknesses. First, the photo-generation charge carriers
peroxide (H2O2) and superoxide anion (•O2ˉ). The reactive oxygen are limited to UV light (that approximately contains 5% of solar light)
species are presumed as beneficial oxidizing elements for contamination due to their wide band gap energy (~3.3 eV). Second, the fast
treatment [8–10]. photo-generated charge carrier recombination rate results in the limited

* Corresponding author.
E-mail address: [email protected] (S. Sheibani).

https://doi.org/10.1016/j.mssp.2021.106083
Received 15 February 2021; Received in revised form 1 July 2021; Accepted 8 July 2021
Available online 12 July 2021
1369-8001/© 2021 Elsevier Ltd. All rights reserved.
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

generation of reactive agents in the photocatalyst process [26–28]. 2. Experimental procedure


Hence, through different research approaches, coupling ZnO with
another narrow band gap semiconductor such as CuO is a practical 2.1. Synthesis of the materials
design strategy to enhance light absorption from UV to UV–Visible range
and increase the lifetime of photo-excited charge carriers [29,30]. CuO Alpha brass foil (65Cu–35Zn) was provided with a thickness of 0.5
is an intrinsic p-type semiconductor with a direct band gap energy mm and cut into 50 × 20 mm2 samples. Afterward, 20 × 20 mm2 of brass
(1.2–1.6 eV) and well known as a low-price, green, recyclable and foil and symmetrical stainless steel 314 immersed into the electrolyte as
visible light active photocatalyst semiconductor [31–33]. Fast charge the working electrode and counter electrode material in the anodizing
carrier recombination rate is the major drawback of CuO, because of its process, respectively. Sodium hydroxide (NaOH), ammonium chloride
narrow band gap energy [34]. Therefore, CuO and ZnO are two excellent (NH4Cl), and acetone were purchased from the Merck KGaA and applied
complementary semiconductors to form an appropriate nanocomposite without any additional refining. Moreover, deionized water (DI, 18.2
in photocatalytic application. MΩ) was utilized as the electrolyte in the anodizing process; and the dye
Several studies [35–37] have investigated various methods to syn­ solution in the photocatalyst analysis. Before the anodizing process, the
thesize ZnO–CuO nanocomposite with various costly or time-consuming brass samples and the stainless steel 314 electrode were mechanically
approaches, including thermal and hydrothermal oxidation, ball mill­ polished to eliminate the surface impurities and pollution. This was
ing, and chemical deposition. Saravanan et al. [38] employed a thermal followed by degreasing in an ultrasonic bath (BANDELIN sonorex, 30W)
decomposition method to synthesize coupled ZnO–CuO semiconductor and finally washed with DI to achieve a clean and smooth surface.
which resulted in higher photocatalytic efficiency compared to that the ZnO–CuO nanocomposites were synthesized through a one-step room
pure ZnO. Li et al. [39] have synthesized tree-like ZnO–CuO through a temperature anodizing method. The electrolyte contained 0.100 M of
complex hydrothermal method and indicated that the branch/stem NaOH and 0.025 M of NH4Cl dissolved in DI. A two-electrode cell con­
structure has improved the photocatalytic performance due to the fast taining the stainless steel and the brass foil with a distance of 50 mm was
carrier transport and effective electron-hole separation. From the pros­ employed for anodization. The anodizing method was conducted at
pect of photocatalysis reactors, powder photocatalyst in slurry mode is room temperature (25 ◦ C) under constant applied voltage (12 V) and
very practical to remove organic dyes. Still, this suspension or slurry constant stirring speed at different times of 15, 30, and 45 min. It should
system has made the scaling-up of the photocatalysis process chal­ be noted that for the sake of comparison, pure Cu foil was also anodized
lenging. From the industrial point of view, it is impractical to recycle the under similar conditions for 30 min. Composition and anodizing time for
powdered photocatalysts from an aqueous solution after photocatalytic different samples are shown in Table 1.
reactions. Many efforts have been made to immobilize photocatalysts on
the supports or substrates as a thin film or coating. Therefore, finding a 2.2. Characterization
cost-efficient, green, and industrial fabrication method becomes more
and more challenging. The brass anodizing is an appropriate method to The phase structure and microstructure of the synthesized ZnO–CuO
fabricate ZnO and CuO semiconductors as a photocatalyst material. nanocomposite coating were characterized by XRD analysis (Philips
Dezfoolian et al. [40] studied the effect of different anodizing parame­ with Cu Kα radiation, λ = 1.54 Å) and FESEM (MIRA3 TESCAN)
ters including applied voltage, the concentration of NaOH, time, and equipped with energy-dispersive X-ray spectroscopy (EDX; JEOL, Cen­
additives on the surface structure of anodized alpha brass. They turio), respectively. The microstructure of the ZCt30 sample was
confirmed the formation of ZnO–CuO nanocomposite during the anod­ investigated by TEM (FEI Tecnai G2 F20 Super Twin), with an acceler­
ization process in an alkaline electrolyte. ation voltage of 200 kV. The average specific surface area (SSA) of the
Generally, brass anodizing has been less studied as a useful method sample was determined by nitrogen adsorption at 77 K (BEL-Belsorp II
to prepare ZnO–CuO coating as a photocatalyst on the brass plate. There instrument) employing the BET isotherm equation. To demonstrate the
is a gap study on the photocatalytic activity and effective parameters on oxidation mechanism, the CV was recorded at room temperature using
ZnO–CuO nanocomposite prepared by brass anodizing. To the best of 0.100 M of NaOH +0.025 M of NH4Cl solution with a scan rate of 50
our knowledge, the fabrication of ZnO/CuO nanocomposite using an mV/s. At the very beginning (after 2 min) and at the end of the anodizing
industrial anodizing method and the effect of brass anodization time on process (after 30 min) 10 mL of electrolyte extracted to evaluate con­
the detailed structural and morphological characteristics, oxidation centrations of dissolved Zn and Cu, using inductively coupled plasma
mechanism, optical properties and photocatalytic activity of the fabri­ optical emission spectroscopy, ICP-OES (Varian-735, United States).
cated ZnO/CuO nanocomposite have not been investigated. Therefore, Additionally, the pH value and potential range of the electrolyte during
in the present work, the anodizing method was utilized to fabricate a the anodization process were 12–14 and 0.21–0.36 V, respectively. The
unique nanocomposite of ZnO flower-like decorated with CuO patches- oxidation mechanism during the brass anodizing process was studied
like with excellent interconnectivity for photocatalytic application. using the Pourbiax diagram of Zn and Cu. The UV–Vis spectra were
Although the basics of the synthesis process are well-developed, the recorded by DRS (Shimadzu, MPC-2200) to examine the band gap en­
following factors can be considered for this particular study; i) normally, ergy of the synthesized samples. PL spectroscopy (Varian Cary Eclipse
nanocomposites are fabricated in two-step or using costly synthesis Fluorescence Spectrophotometer) was employed to study the photo-
methods, whereas, in this study, the nanocomposites were synthesized excited charge carrier recombination rate.
in a one-step and low-cost method. ii) the flower-like morphology of
ZnO is usually obtained in powder form, however, the coating form of
2.3. Photocatalytic tests
ZnO flower-like was fabricated in an ambient temperature method
without any further equipment like a vacuum chamber. The as-prepared
To investigate the photocatalytic activity of the synthesized
nanocomposites were characterized by FESEM, XRD, TEM, BET, CV, PL,
and DRS techniques for structural, microstructural, and optical proper­
ties. In addition, the kinetics of photocatalytic activity, surface effect, Table 1
Composition and anodizing time for different samples.
and reusability of the nanocomposite were investigated by MB and
phenol degradation as common organic water pollutants. Sample Anodizing time (min) Composition

Ct30 30 Pure Cu
ZCt15 15 Brass
ZCt30 30 Brass
ZCt45 45 Brass

2
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

ZnO–CuO nanocomposite coating, MB was utilized as an organic dye. by the fact that dissolved Zn ions can be turned into ZnO because of the
Photo-decomposition of MB with an initial concentration of 2 mg/L in reduction properties of Zn due to the following reactions [41]:
100 mL dye solution (pH = 6.7) over 4 mm2/mL of the synthesized
Cu2 O + Zn→ZnO + 2Cu (1)
ZnO–CuO coating was recorded under visible light exposure (150 W
xenon lamp OSRAM, Germany) with UV filter. The photocatalyst
CuO + Zn = ZnO + Cu (2)
experiment was done for 300 min. First, the dye solution containing the
photocatalyst plate was placed in dark condition for 60 min at ambient To determine the CuO/ZnO phase ratio in the different samples, the
temperature to obtain the equilibrium for absorption and desorption Rietveld refinement technique was applied using XRD patterns. The data
between the dye molecules and photocatalysts. To reveal the photo­ are presented in Table 2 for different samples. It can be concluded that
catalytic mechanism further, some sets of experiment were performed in the anodizing time directly affect the CuO/ZnO phase ratio in the syn­
the presence of methanol (MeOH), disodium ethylenediamine- thesize process. As the anodizing time increases, the ratio of phases in­
tetraacetate (EDTA), and p-benzoquinone (BQ), and cupric nitrate (Cu creases to 0.40 which is equivalent to 28 wt% of CuO in the coating. In
(NO3)2) as the scavengers for •OH, holes, •Oˉ2, and electrons, respec­ other words, anodizing time can count as a reliable factor in the growth
tively. The concentration of MeOH, EDTA, and Cu(NO3)2 was 10 mM, and formation of ZnO/CuO nanocomposite with various CuO/ZnO
and for BQ was 2 mM. The photocatalytic performance over 4 mm2/mL phase ratios.
ZCt30 photocatalyst in the degradation of a colorless organic contami­ FESEM images of the synthesized samples are shown in Fig. 2. Also,
nant, 2 mg/L phenol (>99% purity, Sigma-Aldrich), was also investi­ FESEM images at higher magnification were shown in the top right inset.
gated. Every 30 min, 2 mL of the solution was extracted to determine the The FESEM image of anodized pure Cu foil demonstrates in Fig. 2 (a)
dye concentration by UV–Vis spectrophotometer (PG T80 model) at the showed a flak-like CuO formation. Moreover, it can be inferred from
maximum peaks of 664 nm (for MB) and 270 nm (for phenol). Total Fig. 2 (b) that the ZCt15 sample exhibits a relatively low surface area.
organic carbon (TOC) measurement was performed with a TOC analyzer The FESEM image at higher magnification in Fig. 2 (b) showed the
(Analytik Jena multi-N/C 3100) to determine the degree of MB formation of a hexagonal plate-like ZnO for the anodized brass in 15
mineralization. min. By increasing the anodization time to 30 min, the hexagonal plate-
like ZnO appears as a 3D flower-like morphology. The high concentra­
3. Results and discussion tion of Zn2+ in the early stage of the anodizing process results in ho­
mogeneous nucleation of ZnO and appears hexagonal plate-like
3.1. Structural, microstructural and optical properties morphology. When anodization time goes on, as-formed ZnO tends to
gather together to reduce the surface energy and consequently form the
Fig. 1 represents the XRD patterns of the synthesized samples. ZnO flower-like structure [42]. As is obvious in the higher magnification
Similar peaks were attributed to the brass alloy (as substrate), CuO, and image in Fig. 2 (c), the CuO nanopatches grow on the surface of the
ZnO with different intensities can be observed in all the XRD patterns. as-formed ZnO flower-like. Finally, it can be concluded that a homog­
Two main peaks located at 42.3◦ ((111) planes) and 49.3◦ ((200) planes) enous ZnO–CuO nanocomposite with suitable surface area and inter­
approved the alpha phase (CuZn33) of brass (JCPDS No. 50–1333). connectivity can be formed by brass anodization in 30 min. Besides, by
Three main peaks of wurtzite structure of ZnO (JCPDS No. 36–1451) at further increasing the anodization time to 45 min, the as-formed flow­
angles of 31.74◦ ((100) planes), 34.4◦ ((002) planes), and, 36.3◦ ((101) er-like ZnO in the earlier stage grows more. The overgrowth of the ZnO
planes) are evident. Furthermore, peaks attributed to the monoclinic structure and solubility of ZnO in alkaline electrolyte after 45 min re­
structure of CuO (JCPDS No. 74–1021) located at 32.3◦ ((110) planes), sults in the reduction of the surface area of the structure (see Fig. 2 (d)).
53.7◦ ((020) planes) and, 58.6◦ ((202) planes) are also detected. It It should be noted that as the concentration of Cu2+ increases in the
should be noted that the CuCl2 phase also has formed due to the exis­ alkaline solution by the progress of anodization time, the CuO nano­
tence of NH4Cl additive in the electrolyte. These results indicate that patches start to grow. This can be observed in the FESEM image of the
both ZnO and CuO phases were formed successfully through a simple ZCt45 sample at high magnification.
one-step anodizing process. Also, similar peak patterns of all the samples It is worth mentioning that the morphological structure of the crys­
demonstrate constant phase development during the oxidization pro­ tals contributes to the intrinsic properties and external circumstances. It
cess. In the ZCt15 sample, a peak of Cu2O appeared at 43.3◦ . This peak is
not detectable in ZCt30 and ZCt45 samples, suggesting that as the
Table 2
anodizing process proceeds and Cu2O transforms into CuO. Moreover, as CuO/ZnO phase ratio and BET SSA results for different samples.
shown in Fig. 1 (b), a slight shift in ZnO peak patterns is discernible by
Sample CuO/ZnO phase ratio SSA (m2/g)
increasing the anodizing time. This indicates the copper atoms have
possibly doped in the structure of ZnO through the oxidation process. Ct30 – 16.7
The shift of these peaks is due to the difference between the ionic radius ZCt15 0.14 5.3
ZCt30 0.19 17.9
of Cu and Zn [22]. In the ZCt45 sample, new peaks of ZnO at 47.28 and ZCt45 0.40 11.8
56.14◦ are discernible compared to other samples. This can be explained

Fig. 1. (a) XRD patterns of different samples and (b) an enlarged view of the peaks at 2θ of 31–37◦ .

3
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 2. FESEM images of (a) Ct30, (b) ZCt15, (c) ZCt30 and (d) ZCt45 samples.

is hard to alter the intrinsic characteristics of the crystals. However, capping agent for the (0001) plane) in the electrolyte because the Cl‾
external circumstances of the material can be changed simply. In this ions tend to chelate to the polar plane of ZnO and suppress the ZnO
study, various external conditions including the anodizing time, alkaline growth along the c-axis [45]. Consequently, the ZnO can grow in the
electrolyte, and NH4Cl (as the additive compound in the electrolyte) lateral plane and form the plate-like ZnO structure. During the formation
were considered to investigate the effect of these parameters on the of the ZnO structure, the CuO nanopatches were formed on the surface of
growth of the flower-like ZnO and patches-like CuO. Firstly, the flower- the plate-like ZnO. By prolonging the anodizing time, CuO nanopatches
like morphology of ZnO can be explained in the alkaline electrolyte [43, can grow in size and form higher content of CuO, as seen in Table 2.
44]. The existence of OHˉ in the alkaline electrolyte results in the for­ Hence, the morphological growth of the ZnO/CuO nanocomposite can
mation of Zn(OH)2−4 and consequently, after a dehydration process, the
be controlled in alkaline solution by chloride base additives and anod­
ZnO nuclei and crystal growth of the ZnO occurs (as shown in Equations izing time.
(5)–(7)). Generally, the basal plane ((0001) plane) of ZnO is more active Additionally, one major factor to evaluate the photocatalyst is SSA.
than the other planes owing to its higher surface energy. Therefore, ZnO The average SSA of the samples was measured using the BET method to
should grow along the c-axis and form a one-dimensional structure [43]. evaluate further structural characterization of the samples. As shown in
According to the FESEM images in Fig. 2, it can be seen that the growth Table 2, the ZCt15 sample represents the lowest SSA (5.3 m2/g) among
of ZnO is suppressed along the c-axis and the crystals were grown along all the samples studied. This could be related to the 15 min anodizing
the a-axis which strongly raises the potential formation of ZnO time, and to the fact that the plate-like ZnO mostly lays on the surface of
flower-like assembled by nanoplates. the sample, and consequently appears relatively low SSA. However, as it
The growth suppression along the basal plane and formation of the was mentioned before, the as-formed plate-like ZnO tends to accumulate
plate-like ZnO can be attributed to the existence of Cl‾ ions (as the and form flower-like morphology. When anodizing time continues in the
ZCt30 sample, accumulation of plate-like ZnO not only provides more

4
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

surface of the brass substance for further oxidation but also increases the polycrystalline structure of the sample. The magnified image of the
SSA to 16.7 m2/g. Likewise, after 45 min anodizing due to the over­ marked rectangular area in Fig. 4 (a) is also presented in Fig. 4 (b). The
growth of the structure of ZnO and further formation of CuO nano­ almost uniform distribution of smaller particles with the size of 30 nm
patches on the pours structure of the sample, the SSA in the ZCt45 can be seen on the surface of larger plates. High resolution transmission
sample reduced to 11.8 m2/g. electron microscopy (HRTEM) observation was also performed on these
Fig. 3 depicts the mapping images of the ZCt30 sample. Mapping areas to more accurately evaluate these particles. The HRTEM image is
images of the ZCt30 sample confirmed the distribution of ZnO flower- shown on the bottom inset of Fig. 4 (b). As can be seen, the sample is
like and CuO patches-like nanocomposite. Therefore, the hetero­ comprised of ZnO and CuO with a measured interplanar spacing of 0.284
structure between ZnO and CuO successfully happened through one step and 0.251 nm, corresponding to the (100) and (002) planes of ZnO and
brass anodization process. Furthermore, a suitable and uniform distri­ CuO, respectively. Therefore, the HRTM observation of the sample de­
bution of ZnO and CuO is detectable which is the most important key to picts the heterostructure between ZnO and CuO phases in the ZnO/CuO
enhance the photocatalytic activity of ZnO and CuO. nanocomposite.
The higher SSA and more favorable morphological structure of the To clarify the oxidation mechanism, it should be considered that in
ZCt30 sample made it more appropriate for photocatalytic activities. ZnO and CuO, the anions have a movability which are multiple orders of
Thus, it was chosen for TEM observation. Fig. 4 (a) exhibits the TEM magnitudes lower than those of the cations. Additionally, ZnO and CuO
image and its corresponding selected area electron diffraction (SAED) exhibit different defect structures [46]. During the brass oxidation
pattern of the ZCt30 sample. According to the blue rectangle in Fig. 4 process, zinc has a higher potential reaction and more mobility in
(a), the relatively small particles have formed on the plate-like structure. comparison with copper. This means that due to the anodic
Based on the XRD results and EDS mapping images of the ZCt30 sample, half-reactions and reductive potential values for zinc and copper (see
it can be noticed that the small particles and the relatively larger plate- Eqs. (3) and (4)), zinc can be further oxidized than copper in the same
like morphologies represent CuO and ZnO, respectively. Hence, an electrochemical environment. Therefore, zinc would diffuse to the sur­
appropriate interconnection between ZnO and CuO is discernible. The face of the brass alloy and give up electrons under a constant applied
indexed SAED pattern in the bottom-right inset indicates the presence of voltage. Hence, zinc cations in brass convert to primary Zn2+ ions and
ZnO, CuO, and brass phases which is in agreement with the XRD results. start to form Zn(OH)2−
4 as the initial phase. A dense layer of Zn(OH)2 (s)
Moreover, the existence of faded ring patterns implies the starts to precipitate on the surface of the sample by raising the

Fig. 3. FESEM image and Zn, Cu, and O mapping images of ZCt30 sample.

5
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 4. (a) TEM and corresponding SAED pattern and (b) the magnified image of the marked rectangle of (a) and HRTEM image of the ZCt30 sample.

concentration of Zn(OH)2−
4 . Afterward, hexagonal wurtzite ZnO film
form on the brass substrate according to the following reactions
[47–49]:

(3)

2CuO ↔ 2Cu2+ + O2 + 4e− E = +0.34 V SHE

(4)

2ZnO ​ ↔ 2Zn2+ + O2 + 4e− E = − 0.76V SHE

Zn2+ + 4OH− →Zn(OH)2−4 (aq) (5)

Zn(OH)2−4 (aq) → Zn(OH)2 (s) + 2OH− (6)

Zn(OH)2 → ZnO ​ + H2 O (7)

As time goes by, Cu species start a two-step oxidation process. Firstly,


a small amount of Cu forms Cu2O. Secondly, as-formed Cu2O transforms
into CuO as the final oxide term of Cu as follows [50,51]: Fig. 5. CV of brass in 0.100 M of NaOH +0.025 M of NH4Cl at ambient tem­

Cu + OH →CuOH + e −
(8) perature and scan rate of 50 mV/s.

2CuOH + H2 O→Cu2 O + 2H+ (9) form of zinc [53]. Therefore, the anodic polarization of brass in alkaline
electrolyte begins with Zn(OH)2 and ZnO formation whereas copper
Cu2 O + 2OH− →2CuO + H2 O + 2e− (10) oxides including Cu2O and CuO appear at higher potentials.
According to the ICP-OES analysis of the anodization bath solution at
While copper oxides formed at the surface of brass alloy, CuCl starts
initial and final stages, Pourbiax diagrams drawn by HSC software to
to form on the surface because of the existence of the Clˉ in the elec­
demonstrate the final formed compounds in the determined pH and
trolyte as follow [52]:
potential range (0.21–0.36 V) during the anodizing process. Fig. 6 (a)
Cu+ + Cl− →CuCl (11) and (b) show the Pourbiax diagram of Cu in 10 μM and 10 mM con­
centrations, respectively. The red box in all diagrams indicates the ob­
2CuCl = Cu + CuCl2 (12) tained pH range (12–14) of electrolytes during the anodizing process. As
seen in Fig. 6 (a), in 10 μM concentration of Cu, Cu(OH)2ˉ is stable be­
To be more specific, the production of the metal oxide layer starts
tween pH = 12 and 14 and for the potential more than ~ − 0.20 V.
with the production of metal hydroxides and then goes on with dehy­
Moreover, in positive potentials, CuO would be stable and in a pH range
dration to metal oxides. Hence, the brass oxidation process follows the
of more than 13.4, copper dissolution would be associated with the
dissolution-precipitation mechanism. To confirm this mechanism, the
formation of [Cu(OH)4]ˉ. However, for a higher concentration of Cu
CV was recorded by a scan rate of 50 mV/s (as shown in Fig. 5). The
(Fig. 6 (b)), the stability of the CuO region is increased. In other words,
voltammogram was launched by sweeping between − 1.5 and 0.3 V/
the formation of CuO is predicted by the Pourbiax diagram in this sys­
SCE. The A1 and A2 anodic peaks are attributed to the ZnO and Zn(OH)2
tem, which is in strong agreement with XRD data. Fig. 6 (c) and (b)
formation, respectively [53]. Also, A3 and A4 peaks are related to the
exhibit the Pourbiax diagram of Zn in 10 μM and 12 mM concentrations,
formation of Cu2O and CuO, respectively [53,54]. In other words, C1 and
respectively. According to Fig. 6 (c), ZnOH+ is stable at potentials higher
C3 peaks indicated the reduction of Cu2+to Cu+ and reduction of Cu+ to
than − 0.76 V and in the range of pH = 7.4 to 8.8. Moreover, Zn(OH)2
the metallic form of copper, respectively. Moreover, the C2 peak is
would be stable in a small range from pH = 8.8 to 9. Additionally,
referred to as the reduction of Zn2+ in Zn(OH)2 and ZnO to the metallic

6
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 6. Pourbiax diagram for copper in (a) 10 μM and (b) 10 mM concentrations and for zinc in (c) 10 μM and (d) 12 mM concentrations in aqueous electrolyte.

between pH = 9 to 11.5, ZnO would be stable. Finally, Zn would be between O2− (2p) levels to Zn2+ (3d – 4s) orbitals as the valence band
present in the solution as different anions by raising the pH. By and conduction band, respectively [56]. Moreover, increasing anod­
increasing Zn concentrations in the system and when the potential is ization time and formation of CuO on primary ZnO layers results in the
greater than − 1 V, as it is demonstrated in Fig. 6 (d), ZnO would be the formation of Cu2+ (3d) and Zn2+ (3d – 4s) conduction bands assembled
stable compound. According to the Eh-pH diagrams at given voltages by the sharing of O− valence band [35,56]. Significant improvement in
and between pH = 12 to 14, thermodynamically the CuO and [Cu visible light absorption (red-shift) is detectible since the electron exci­
(OH)4]2ˉ are more likely to form. Among these two species, CuO is more tation between O2− (2p) and Cu2+ (3d) orbitals require lower energy
desired to be formed because of the pH and potential range. Moreover, compare to that of O2− (2p) and Zn2+ (3d – 4s). Furthermore, owing to
the presence of ZnO and zinc anions such as [Zn(OH)4]2ˉ and [Zn(OH)3]ˉ the closely spaced Cu2+ (3d) and Zn2+ (3d – 4s) energy states, electrons
are probable. According to the determined pH and potential range, the can be transmitted between these two energy states (d-d transitions)
presence of ZnO as a more possible thermodynamically stable phase is which indicated another possible reason for light absorption in the
observable. visible spectrum [56].
For more clarification, Fig. 7 represents the schematic of the oxida­ To calculate the optical band gap energy (Eg) of the samples, Tauc
tion process during the anodization process. At the beginning of the plots was used, which is defined as a function between adsorption co­
oxidation process (step 1), OHˉ can be produced through the cathodic efficient (α), and photon energy (hν), according to the following equa­
half-reaction, and Zn2+ ions diffuse to the surface of the brass. Hence, tion [57]:
while copper is losing electrons and turning to Cu2+ ions, ZnO is the first
oxide compound formed on the surface of the brass (step 2). Then, the (αhν)n = A(hν - Eg) (13)
Cu2+ firstly starts to produce Cu2O (step 3). Afterward, because of the where A is constant and n is equal to 2 due to the direct band gap energy
applied potential in the alkaline electrolyte CuO can be formed as the of ZnO and CuO [58,59]. Depending on the Tauc plots in Fig. 8 (b), the
final oxide compound of copper (step 4). However, since the solubility of optical band gap energy values of the samples can be measured by
ZnO in alkaline solution is more than CuO, it is safe to say that the outer extrapolating the linear region of (αhν)2 versus photon energy. The band
layer of the coated layer mostly contains CuO compound. This is in gap energy of anodized pure copper foil (Ct30 sample) is equal to 1.3 eV.
agreement with the previous report [55]. A previous study [32] has indicated that the band gap energy of CuO is
DRS analysis was carried out at ambient temperature to study the equal to 1.2–1.6 eV. This confirms that the band gap energy of the Ct30
optical properties of the synthesized samples. Fig. 8 (a) represents the sample is related to the formation of CuO coating on the surface of the
UV–Vis optical transmittance spectra of the anodized samples. UV–Vis pure copper foil. The estimated band gap values for ZCt15, ZCt30, and
adsorption-desorption spectra of the samples show notably improved ZCt45 samples are 3.1, 3, and 2.9 eV, respectively. The red shift in
adsorption in visible spectra (red-shift) by increasing anodization time. bandgap energy value by the longer anodization process can be
These results suggest that the contact between CuO and ZnO on the explained by the formation of CuO on the ZnO layer in the oxidation
surface of the anodized brass enhanced the absorption in the visible process and confirmed excitations from O2− (2p) to Zn2+ (3d – 4s) and
range and might be an effective factor for superior photocatalytic effi­ Cu2+ (3d) [56]. Previous studies [60,61] indicated that the band gap
ciency under visible light irradiation. In the ZCt15 sample, where the energy of ZnO–CuO nanocomposite synthesized by various methods
surface of the plate mostly covered with a thin ZnO layer, the sharp measured 2.6–3.1 eV which are in agreement with our results.
absorption in ~250 nm is revealed in the UV area due to the excitation PL analysis is based on the recombination rate of photo-induced

7
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 7. Schematic graph of the anodization mechanism of the brass alloy.

charge carriers. Hence, useful information about the optical character­ due to the electron-hole recombination rate, as an effective factor for
istics of semiconductors including trapping and transmission of charge photocatalytic application.
carriers can be extracted from the PL data [62]. Fig. 9 represents the PL
spectra of the anodized samples under an excitation wavelength of 290 3.2. Photocatalytic properties
nm. The results indicated the same peak position with different in­
tensities. Anodized brass samples reveal one UV peak emission located The photocatalytic activity of the ZnO–CuO nanocomposites was
at 385 nm of the characteristic band of ZnO which is in agreement with studied by the degradation rate of MB as an organic dye pollutant in
the literature [63,64]. Owing to the lack of the ZnO content in the wastewaters, under irradiation of visible light for 300 min. Fig. 10 (a)
anodized pure copper foil, the peak emission located at 385 nm does not illustrates the photo-degradation activity of MB as a function of irradi­
appear for this sample. The above-mentioned peak emission can be ation time over different samples with catalyst dosage of 4 mm2/mL (the
ascribed to the band-to-band transition between the conduction band surface area to the solution volume). The ZCt15 sample shows the
constituted by Zn2+ (3d) and O2− (2p) as the valence band [65,66]. minimum photocatalytic activity (degradation of 27%) among all the
Moreover, the reduction of the peak intensities demonstrates a samples. The Ct30 and ZCt45 samples show 35% and 36% degradation
discernible decrease in the recombination rate of photo-induced of MB after 300 min, respectively. Finally, the ZCt30 sample represents
charges. In other words, by comparing the peak emission intensity of 54% photocatalytic efficiency, which is the highest efficiency among all
different CuO–ZnO nanocomposites, a notable decrease in the peak samples. Trends in photocatalytic performance of different samples are
emission intensity of the anodized brass in 30 min compared to anodized in agreement with the structural and optical properties. The photo­
brass in 15 and 45 min can be observed. As it was seen in the FESEM catalytic activity of the ZCt15 sample demonstrates lower photocatalytic
image of the ZCt30 nanocomposite (Fig. 2 (c)), there is an acceptable activity compared with the ZCt30 sample because of the non-uniform
contact between CuO and ZnO. In this way, instead of electron and hole distribution of CuO along with ZnO in the ZCt15 sample. Moreover,
recombination, the photo-excited electrons can be separated by the smaller microstructure, more visible light activity, and more effective
transition between the conduction band of ZnO and the conduction band electron-hole separation can be counted as other reasons for a higher
of CuO [36]. However, further growth of CuO and ZnO phases in the photocatalytic activity of the ZCt30 sample in comparison with the
ZCt45 sample results in reduced contact between them. Hence, the ZCt15 sample. Additionally, the ZCt30 sample also demonstrates higher
possibility of charge transfer decreased for this sample. Moreover, since photocatalytic performance compared to the ZCt45 sample. This can be
the formation of CuO has not happened completely in the ZCt15 sample, attributed to further growth, and consequently decreased surface area of
there is no appropriate contact between ZnO and CuO to obtain ZCt45 during the longer anodization process. Moreover, another reason
adequate charge separation. Finally, the efficient charge carrier sepa­ for the difference between the photocatalytic activity of ZCt30 and
ration may lead to enhanced photocatalytic activity in the ZCt30 sample ZCt45 samples is the less electron-hole recombination rate in the ZCt30

8
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

ln(C0 / C) = ​ kapp t (14)

where C0, C, and kapp depict the initial concentration of MB, the con­
centration of MB at time t, and the apparent rate constant, respectively.
The linear plot of ln (C0/C) versus irradiation time in the bottom left
inset of Fig. 10 (a) shows that the MB degradation is well fitted with a
pseudo-first-order model. Therefore, the kinetic model applied in this
experimental study agrees with the literature [36,69]. For more clari­
fication of the kinetics, Fig. 10 (b) exhibits the apparent rate constants of
photocatalytic degradation of different photocatalysts. The ZCt30
nanocomposite exhibits the highest rate constant (kapp = 0.00261
1/min) among all the samples. This indicates that the rate constant of
the ZCt30 sample is ~2.5 times greater than that of the ZCt15 sample
and ~1.7 times more than that of the ZCt45 and Ct30 samples.
For additional verification of the photocatalytic performance of the
ZCt30 sample, the photocatalytic experiment was applied on phenol as a
colorless pollutant under a similar condition. The result was compared
with the degradation of MB in Fig. 10 (c). It can be seen that the
degradation rate of phenol was lower than MB and the degradation
efficiency reached 49% after 300 min visible irradiation. Furthermore,
kapp value for phenol degradation was 0.00222 1/min. It has been
confirmed in a recent report [60] that ZnO–CuO nanocomposites
generally exhibit lower photocatalytic efficiency on phenolic pollutants
in comparison with MB. However, the close photo-decomposition effi­
ciency of phenol and MB as two different pollutants over the ZCt30
sample implies that the nanocomposite is an applicable photocatalyst on
the degradation of various pollutants.
As shown in Fig. 10 (d), the TOC test proved the mineralization of MB
during photocatalytic degradation. TOC values were related to the
amount of carbon found in the MB dye that was degraded during the
Fig. 8. (a) Transmittance versus wavelength graphs and (b) Tauc plots of process. TOC was measured for mineralization of MB solution over the
different samples. ZCt30 photocatalyst. TOC removal of 54% was observed after 300 min.
Therefore, both UV–Vis and TOC data represented relatively similar
results indicating that MB experienced mineralization into CO2 and H2O
and that most intermediate products degraded during the process [70].
To clarify the photo-degradation mechanism of MB by the ZnO–CuO
nanocomposite, it should be noted that CuO as a p-type semiconductor
has relatively low visible light activity [71–73]. One possible reason is
that the band structure of CuO is not capable enough to start with the
process of oxidation of H2O to •OH reactive radical and the reduction of
O2 to •O2ˉ reactive radicals [74]. Additionally, the fast electron-hole
recombination rate and lack of active sites in the structure of CuO
may be another possible explanation for the relatively low photo­
catalytic potential of CuO [32]. However, in CuO–ZnO nanocomposite,
CuO plays the dominant role in photocatalytic activity as it absorbs
visible light, which results in maximum electron-hole separation. Be­
sides, heterostructure with ZnO inhibits the recombination of the
photo-generated charge carrier. Based on the previous reports [75–77]
working on powdered and non-powdered ZnO–CuO nanocomposites, it
Fig. 9. PL spectra of different samples. can be claimed that the intrinsic defect sites can be observed in the
as-prepared ZnO–CuO nanocomposites. Hence, the intrinsic defect sites
sample. The results demonstrate that the ZCt30 sample is ~1.5 times in the ZnO structure trapped the photo-excited charges, which is
more active than the Ct30 sample in photocatalytic usage. Besides, the favorable for the photocatalytic activity of the nanocomposite [64]. To
formation of heterostructure between ZnO and CuO in anodized brass is be more specific, it has been confirmed that the electron acceptors (V0),
the possible reason for the enhanced photocatalytic efficiency of ZCt30 hole accepters (Oi), and oxygen vacancy are three major defects in ZnO.
in comparison with the Ct30 sample. Therefore, it can be concluded that Therefore, these defects can act as sub-band gap states whit lower en­
coupling CuO with ZnO in anodized brass can improve the photo- ergy than the real band gap of ZnO, which could be a secondary source
degradation of MB under visible light irradiation. The p-n junction be­ for visible light absorption [66]. Also, these surface defects could inhibit
tween ZnO and CuO, visible light activity, higher surface area and electrons and holes recombination and prolong the lifetime of charge
effective electron-hole separation are effective factors for the enhanced carriers. This increases the reduction and oxidation reaction potentials
photocatalytic performance of ZnO–CuO nanocomposite. This also has of the photocatalytic process. To be more specific, V0 can trap the
been confirmed by the kinetic study of the degradation process. The photo-excited electrons. Whereas, Oi is a shallower trap of photo-excited
photo-degradation of organic pollutants can follow a pseudo-first-order holes. While the degradation process takes place on the surface of V0 and
kinetics model according to the well-known equation [67,68]: Oi defects, oxygen vacancies in ZnO could be further possible active sites
in the photocatalytic process [78]. Hence, the photo-generated charges

9
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 10. (a) Photocatalytic degradation rate and reaction kinetics, (b) kapp values of MB degradation by different samples, (c) comparison between photocatalytic
degradation rate of phenol and MB, and (d) comparison between TOC and UV–Vis results of MB degradation (with the surface area to volume of dye solution being 4
mm2/mL) under visible light radiation.

on the surface of the photocatalyst participate in several reactions with absolute electronegative factor (χCuO = 5.81 eV and χZnO = 5.79 ​ eV)
oxygen and water molecules to generate series of reactive oxygen spe­ and, Ee is assumed to be equal to ~ 4.5 eV as the energy of the free
cies including •OH, •Oˉ2 hydroperoxyl radical (•HO2), and H2O2 ac­ electron in the hydrogen standard. In the same way, the bottom of the
cording to the following reactions [36,74]: conduction band can be calculated as follow:

ZnO + hν→ZnO(e− ) + ZnO(h+ ) (15) EBCB = ETVB − Eg (23)

Calculated band edge positions (see Fig. 11 (c)) are in agreement


CuO + hν→CuO(e− ) + CuO(h+ ) (16)
with DRS results indicating the conduction and valence band edge po­
sitions of CuO locate between ZnO band structures. Previous reports [60,
e− (ZnO / CuO) + O2 →Ȯ−2 (17)
79] over ZnO–CuO nanocomposite have suggested that if the conduction
band and valence band edge position of CuO lay between those of the
˙ − + H+
h+ (ZnO / CuO) + H2 O→OH (18)
wurtzite structure of ZnO, another effective photo-decomposition
2e− + O2 + 2H+ →H2 O2 (19) mechanism can be activated. Therefore, the photo-generated electrons
can be transferred between CuO and ZnO conduction bands. In this
_ + OH− + O2
Ȯ2 ˉ + H2 O→OH (20) scenario, the relocation of electrons between these two states of energy
results in the reduction of Cu2+ to Cu+. Hence, the reaction between Cu+
_2
O2 ˉ + H+ →HO (21) and dissolved oxygen molecules in the dye solution leads to the gener­
ation of further hydrogen peroxide as follows [60]:
Additionally, the band edge position of semiconductors can be
evaluated according to the following equation [60]: CuO / ZnO + hν→(ZnO)h+ + (CuO)e− (24)

ETVB = χ − Ee + 0.5Eg (22) Cu2+ + e− →Cu+ (25)

where ETVB is the energy at the top of the valence band, χ represents the 2Cu+ + 2H+ + O2 →2Cu++ + H2 O2 (26)

10
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 11. (a) Effect of various scavengers on the photocatalytic performance of the ZCt30 sample, (b) the kapp in the presence of different scavengers, and (c)
schematic diagram of photo-degradation mechanism of MB by ZnO–CuO nanocomposite.

Eventually, MB can be oxidized and break down by the generated As can be seen, the role of holes and •OH are discernible in the
reactive oxygen species (ROS) as stated by the following reaction: photo-degradation mechanism of MB over ZnO/CuO nanocomposite.
Therefore, considering all the results, it can be concluded that MB as a
ROS + MB→intermediates ​ →CO2 + H2 O (27)
harmful organic dye can be degraded by the synthesized ZnO–CuO
To indicate the role of electrons, holes, OH, and

O2
• ˉ
in the MB nanocomposite through a simple anodizing method for 30 min.
degradation process, several photocatalysis experiments were done in The photo-durability characteristic of the photocatalyst is one of the
the presence of different scavengers. Cu(NO3)2, EDTA, MeOH, and BQ major concerns for utilizing photocatalyst materials in industrial appli­
were utilized as the representative scavengers of electrons, holes, •OH, cations. Therefore, it is important to evaluate the environmental dura­
and •Oˉ2, respectively [80,81]. Fig. 11 (a) exhibits the photo-degradation bility of the photocatalyst. Reusability of the ZCt30 photocatalyst, as the
of MB as a function of time and the impact of various scavengers on the most efficient sample, has been investigated by consecutively moni­
photocatalytic efficiency of the ZCt30 sample. As illustrated in Fig. 11 toring the photo-degradation of MB under visible light irradiation after
(b), by applying Cu(NO3)2 and BQ, the kapp of photocatalytic perfor­ three cycles. After each cycle, the photocatalyst was washed with
mance exhibits only a slight decrease. This indicates that both electrons deionized water and acetone to eliminate the remnant solution on the
and •Oˉ2 affect the photocatalytic activity of the ZCt30 nanocomposite, surface. After drying, the sample was immersed in a fresh solution with a
insignificantly. On the other hand, the kapp is decreased to 0.00051 and similar concentration of MB for another cycle. As shown in Fig. 12 (a),
0.00043 by applying EDTA and MeOH, respectively. Hence, it can be photo-degradation of ZnO–CuO nanocomposite demonstrates ~54% MB
inferred that the photo-generated holes are more effective in the pho­ degradation in the first cycle. The calculated photocatalytic activity of
tocatalytic performance of the ZnO–CuO nanocomposite compared with the sample was 51% and 49% in the second and third cycles, respec­
the photo-generated electrons. Besides, although •Oˉ2 has a negligible tively. This means that only a 5% drop in photocatalytic efficiency is
effect on the photo-degradation of MB, the role of •OH is greater in the observed after three cycles. The relatively slight reduction in photo­
photocatalytic performance of the sample. To be more specific, the catalytic activity can be referred to the inactivation of some adsorption
photo-generated holes can put a significant impact on the photocatalytic sites on the surface of the photocatalyst and/or the partial phase
activity of ZnO/CuO nanocomposite by directly reacting with MB and transformation of the photocatalyst in the dye solution under visible
decomposing the pollutant into the same byproducts as the product from light. Since no significant decrease in photocatalytic activity of the
the reaction between •OH and MB [82]. Furthermore, the sample is observed during consecutive cycles, it can be claimed that the
photo-generated holes can participate in water oxidation which results ZCt30 sample shows favorable reusability in photocatalytic usage. For
in the production of •OH, as well. Hence, the holes and •OH are both the further clarification, the ZCt30 sample was characterized by XRD after
major species for the photocatalytic mechanism of the ZCt30 sample on three cycles of photocatalytic experiments, which is shown in Fig. 12 (b).
photo-degradation of MB. These results are in agreement with a previous According to this data, the XRD pattern of the ZCt30 sample after three
study [83] on the MB photo-degradation by ZnO/CuO nanocomposite. cycles of photocatalyst experiments confirms the reliable phase stability
For more elucidation, Fig. 11 (c) represents the possible mechanism for of the as-prepared photocatalyst. The reusability of the ZnO–CuO
MB degradation in the ZnO–CuO nanocomposite photocatalytic system. nanocomposite, which is produced by a very simple, inexpensive, and

11
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Fig. 12. (a) Reusability of the ZCt30 sample (with the surface area to volume of dye solution of 4 mm2/mL) on the degradation of 2 mg/L MB solution under visible
light for three cycles and (b) XRD patterns of the ZCt30 sample before and after three cycles.

short time process, can confirm the industrial view of this research. It
can be claimed that the ZnO–CuO nanocomposite coating synthesized by
brass anodizing can be a reliable method to overcome the recyclability
issues of powder photocatalysts in a continuous reactor.
The photocatalyst dosage in photocatalytic applications is an effec­
tive factor in degradation efficiency. In plate form materials, the effect of
photocatalyst dosage is evaluated by the proportion of the surface area
to volume of dye solution [34]. Hence, the surface effect of the syn­
thesized plates for degradation of MB was investigated at three values of
the surface area to volume of dye solution; 4, 8, and 12 mm2/mL. Fig. 13
shows the MB degradation efficiency by ZCt30 sample at the three
photocatalyst dosage. Higher photocatalytic efficiency is obtained at 8
mm2/mL compared with 4 mm2/mL. This can be attributed to the more
available active sites and increasing in the charge carrier generation. On
the contrary, applying extra dosage of the ZCt30 photocatalyst plates at Fig. 13. The photocatalytic degradation rate of MB by ZCt30 sample under
12 mm2/mL reduces the photo-degradation efficiency. First of all, this visible light irradiation at three ratios of the surface area to volume of
may because that the MB adsorption to the surface of the nanocomposite dye solution.
is dominant. Secondly, using excessive photocatalyst dosage may pre­
vent light to reach the surface of the samples. results indicated a red-shift in light absorption and narrower band gap
Various factors including the dye concentration, the dosage of the energy of anodized brass at the longer anodization time. PL analysis
photocatalyst, and the light intensity affect the performance and the showed that the anodized brass at 30 min has a lower electron-hole
photocatalytic degradation rate. In the current research, the photo- recombination rate due to the appropriate contact between ZnO and
degradation of MB in 100 mL of 2 mg/L dye solution under 300 min CuO. Photo-degradation of MB showed that the ZnO–CuO nano­
visible light reached 54%, using a 150 W tungsten–halogen light source composite prepared by brass anodizing at 30 min had the highest pho­
over 4 mm2/mL of the ZCt30 sample. To investigate the industrial tocatalytic efficiency. Surface area, visible light activity and, electron-
capability of the ZCt30 sample, the photocatalytic performance of this hole recombination rate are discussed as the possible reasons for the
sample was compared with the TiO2 (P25) as a well-known photo­ difference between the photocatalytic activity of anodized brass at
catalyst material. The previous report [84] showed 60% degradation of various anodization times. The holes and •OH radicals were proposed as
10 mg/L MB solution after 120 min irradiation under similar experi­ the important elements for the photo-degradation mechanism over the
mental conditions over P25 photocatalyst with the dosage of 0.2 g/L. synthesized photocatalyst. The ZnO–CuO nanocomposite represents
The relatively lower photocatalytic performance of the nanocomposite favorable reusability after three cycles. The maximum photocatalytic
coating sample prepared in the present study is due to its non-powder efficiency was obtained with the photocatalyst dosage of 8 mm2/mL and
form, which usually has a much lower specific surface area than the decreased at the higher dosage. It seems that utilizing brass oxidation as
powder samples. Although the photocatalytic efficiency of the ZCt30 a cost-efficient and applicable method to produce ZnO–CuO nano­
sample is lower than that of the P25, still utilizing the ZCt30 sample as a composite can be a reliable approach for the industrialization of this
photocatalyst might be practical because of the non-powder form nanocomposite in photocatalytic applications.
advantage in recycling and notable stability of the sample.
Author statement
4. Conclusions
M. Nami: Conceptualization, Methodology, Data curation, Writing –
ZnO–CuO nanocomposite was synthesized through simple one-step
original draft, S. Sheibani: Visualization, Validation, Supervision,
brass anodizing in alkaline aqueous solution at different anodization
Writing- Reviewing and Editing, F. Rashchi: Visualization, Validation,
times. XRD and FESEM characterizations showed the formation of
Supervision, Writing- Reviewing and Editing.
wurtzite ZnO flower-like and monoclinic CuO patches-like
(morphology) phases. Moreover, HRTEM analysis revealed the hetero­
structure between ZnO and CuO phases. SSA measurements using the Declaration of competing interest
BET method revealed that the ZCt30 sample exhibits the highest SSA
among all the samples. Pourbiax diagram and CV analysis suggested the The authors declare that they have no known competing financial
formation mechanism of ZnO and CuO during brass anodization. DRS interests or personal relationships that could have appeared to influence
the work reported in this paper.

12
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

Acknowledgments [25] C.G. Núñez, F. Liu, W.T. Navaraj, A. Christou, D. Shakthivel, R. Dahiya,
Heterogeneous integration of contact-printed semiconductor nanowires for high-
performance devices on large areas, Microsystems Nanoeng 4 (2018) 1–15.
The authors would like to acknowledge the support of the University [26] M. Wang, F. Ren, J. Zhou, G. Cai, L. Cai, Y. Hu, D. Wang, Y. Liu, L. Guo, S. Shen,
of Tehran and the Iran Nanotechnology Initiative Council for this N doping to ZnO nanorods for photoelectrochemical water splitting under visible
research. light: engineered impurity distribution and terraced band structure, Sci. Rep. 5
(2015) 1–13.
[27] S. Le, T. Jiang, Y. Li, Q. Zhao, Y. Li, W. Fang, M. Gong, Highly efficient visible-light-
References driven mesoporous graphitic carbon nitride/ZnO nanocomposite photocatalysts,
Appl. Catal. B Environ. 200 (2017) 601–610.
[28] E.T.D. Kumar, S. Easwaramoorthi, J.R. Rao, Fluorinated reduced graphene oxide-
[1] P. Ghasemipour, M. Fattahi, B. Rasekh, F. Yazdian, Developing the ternary ZnO
encapsulated ZnO hollow sphere composite as an efficient photocatalyst with
doped MoS 2 nanostructures grafted on CNT and reduced graphene oxide (RGO)
increased charge-carrier mobility, Langmuir 35 (2019) 8681–8691.
for photocatalytic degradation of aniline, Sci. Rep. 10 (2020) 1–16.
[29] M. Pirhashemi, A. Habibi-Yangjeh, S.R. Pouran, Review on the criteria anticipated
[2] M.E.M. Ali, E.A. Assirey, S.M. Abdel-Moniem, H.S. Ibrahim, Low temperature-
for the fabrication of highly efficient ZnO-based visible-light-driven photocatalysts,
calcined TiO 2 for visible light assisted decontamination of 4-nitrophenol and
J. Ind. Eng. Chem. 62 (2018) 1–25.
hexavalent chromium from wastewater, Sci. Rep. 9 (2019) 1–9.
[30] S. Chabri, A. Dhara, B. Show, D. Adak, A. Sinha, N. Mukherje, Mesoporous CuO-
[3] Z. Fan, T. Wu, X. Xu, Synthesis of reduced graphene oxide as a platform for loading
ZnO pn heterojunction with high specific surface area for enhanced photocatalysis
β-NaYF 4: Ho 3+@ TiO 2 based on an advanced visible light-driven photocatalyst,
and electrochemical sensing, Catal, Sci. Technol. 6 (2016) 3238–3252.
Sci. Rep. 7 (2017) 1–15.
[31] J. Yang, W. Yin, B. Zhou, A. Cui, L. Xu, D. Zhang, W. Li, Z. Hu, J. Chu, Composition
[4] H. Wang, G. Wang, Y. Zhang, Y. Ma, Z. Wu, D. Gao, R. Yang, B. Wang, X. Qi,
dependence of optical properties and band structures in p-type Ni-doped CuO films:
J. Yang, Preparation of RGO/TiO 2/Ag aerogel and its photodegradation
spectroscopic experiment and first-principles calculation, J. Phys. Chem. C 123
performance in gas phase formaldehyde, Sci. Rep. 9 (2019) 1–12.
(2019) 27165–27171.
[5] G. Mano, S. Harinee, S. Sridhar, M. Ashok, A. Viswanathan, Microwave assisted
[32] M.B. Gawande, A. Goswami, F.-X. Felpin, T. Asefa, X. Huang, R. Silva, X. Zou,
synthesis of ZnO-PbS heterojuction for degradation of organic pollutants under
R. Zboril, R.S. Varma, Cu and Cu-based nanoparticles: synthesis and applications in
visible light, Sci. Rep. 10 (2020) 1–14.
catalysis, Chem. Rev. 116 (2016) 3722–3811.
[6] Y. Nosaka, A.Y. Nosaka, Generation and detection of reactive oxygen species in
[33] G. Hu, C.-X. Hu, Z.-Y. Zhu, L. Zhang, Q. Wang, H.-L. Zhang, Construction of Au/
photocatalysis, Chem. Rev. 117 (2017) 11302–11336.
CuO/Co3O4 tricomponent heterojunction nanotubes for enhanced photocatalytic
[7] G. Mishra, M. Mukhopadhyay, TiO 2 decorated functionalized halloysite nanotubes
oxygen evolution under visible light irradiation, ACS Sustain. Chem. Eng. 6 (2018)
(TiO 2@ HNTs) and photocatalytic PVC membranes synthesis, characterization and
8801–8808.
its application in water treatment, Sci. Rep. 9 (2019) 1–17.
[34] S. Behjati, S. Sheibani, J. Herritsch, J.M. Gottfried, Photodegradation of dyes in
[8] M.A. Moosavi, M. Sharifi, S.M. Ghafary, Z. Mohammadalipour, A. Khataee,
batch and continuous reactors by Cu2O-CuO nano-photocatalyst on Cu foils
M. Rahmati, S. Hajjaran, M.J. Łos, T. Klonisch, S. Ghavami, Photodynamic N-TiO 2
prepared by chemical-thermal oxidation, Mater. Res. Bull. (2020) 110920.
nanoparticle treatment induces controlled ROS-mediated autophagy and terminal
[35] M. Mansournia, L. Ghaderi, CuO@ ZnO core-shell nanocomposites: novel
differentiation of leukemia cells, Sci. Rep. 6 (2016) 1–16.
hydrothermal synthesis and enhancement in photocatalytic property, J. Alloys
[9] B. Liu, D. Yin, F. Zhao, K.K. Khaing, T. Chen, C. Wu, L. Deng, L. Li, K. Huang,
Compd. 691 (2017) 171–177.
Y. Zhang, Construction of a novel Z-scheme heterojunction with molecular grafted
[36] S. Chabri, A. Dhara, B. Show, D. Adak, A. Sinha, N. Mukherjee, Mesoporous
carbon nitride nanosheets and V2O5 for highly efficient photocatalysis, J. Phys.
CuO–ZnO p–n heterojunction based nanocomposites with high specific surface area
Chem. C 123 (2019) 4193–4203.
for enhanced photocatalysis and electrochemical sensing, Catal. Sci. Technol. 6
[10] H. Yin, K. Yu, C. Song, R. Huang, Z. Zhu, Synthesis of Au-decorated V2O5@ ZnO
(2016) 3238–3252.
heteronanostructures and enhanced plasmonic photocatalytic activity, ACS Appl.
[37] L. Zhu, H. Li, Z. Liu, P. Xia, Y. Xie, D. Xiong, Synthesis of the 0D/3D CuO/ZnO
Mater. Interfaces 6 (2014) 14851–14860.
heterojunction with enhanced photocatalytic activity, J. Phys. Chem. C 122 (2018)
[11] H. Maleki, V. Bertola, TiO2 nanofilms on polymeric substrates for the
9531–9539.
photocatalytic degradation of methylene blue, ACS Appl. Nano Mater. 2 (2019)
[38] R. Saravanan, S. Karthikeyan, V.K. Gupta, G. Sekaran, V. Narayanan, A. Stephen,
7237–7244.
Enhanced photocatalytic activity of ZnO/CuO nanocomposite for the degradation
[12] C.B. Ong, L.Y. Ng, A.W. Mohammad, A review of ZnO nanoparticles as solar
of textile dye on visible light illumination, Mater. Sci. Eng. C 33 (2013) 91–98.
photocatalysts: synthesis, mechanisms and applications, Renew. Sustain. Energy
[39] Z. Li, M. Jia, B. Abraham, J.C. Blake, D. Bodine, J.T. Newberg, L. Gundlach,
Rev. 81 (2018) 536–551.
Synthesis and characterization of ZnO/CuO vertically aligned hierarchical tree-like
[13] C.L. Londoño-Calderón, S. Menchaca-Nal, L.G. Pampillo, P. Froimowicz, Cupric
nanostructure, Langmuir 34 (2018) 961–969.
oxide nanoleaves for the oxidative degradation of methyl orange without heating
[40] M. Dezfoolian, F. Rashchi, R.K. Nekouei, Synthesis of copper and zinc oxides
or light, N.J. Franç ois, ACS Appl. Nano Mater. 3 (2020) 2987–2996.
nanostructures by brass anodization in alkaline media, Surf. Coating. Technol. 275
[14] H. Yu, X. Huang, P. Wang, J. Yu, Enhanced photoinduced-stability and
(2015) 245–251.
photocatalytic activity of CdS by dual amorphous cocatalysts: synergistic effect of
[41] E. Beucher, M. Lenglet, S. Weber, S. Scherrer, Oxidation mechanisms on α-brasses
Ti (IV)-hole cocatalyst and Ni (II)-electron cocatalyst, J. Phys. Chem. C 120 (2016)
(10 and 30 at.% Zn) by optical methods and SIMS, Surf. Interface Anal. 18 (1992)
3722–3730.
673–678.
[15] M. Jabeen, M. Ishaq, W. Song, L. Xu, I. Maqsood, Q. Deng, UV-assisted
[42] B. Li, Y. Wang, Facile synthesis and photocatalytic activity of ZnO–CuO
photocatalytic synthesis of ZnO-reduced graphene oxide nanocomposites with
nanocomposite, Superlattice. Microst. 47 (2010) 615–623.
enhanced photocatalytic performance in degradation of methylene blue, ECS J.
[43] R.S. Ganesh, G.K. Mani, R. Elayaraja, E. Durgadevi, M. Navaneethan,
Solid State Sci. Technol. 6 (2017) M36.
S. Ponnusamy, K. Tsuchiya, C. Muthamizhchelvan, Y. Hayakawa, ZnO hierarchical
[16] O. Tomita, T. Otsubo, M. Higashi, B. Ohtani, R. Abe, Partial oxidation of alcohols
3D-flower like architectures and their gas sensing properties at room temperature,
on visible-light-responsive WO3 photocatalysts loaded with palladium oxide
Appl. Surf. Sci. 449 (2018) 314–321.
cocatalyst, ACS Catal. 6 (2016) 1134–1144.
[44] H. Liu, M. Li, Y. Wei, Z. Liu, Y. Hu, H. Ma, A facile surfactant-free synthesis of
[17] H. Liang, Y.-C. Hu, Y. Tao, B. Wu, Y. Wu, J. Cao, Existence of ligands within
flower-like ZnO hierarchical structure at room temperature, Mater. Lett. 137
sol–gel-derived ZnO films and their effect on perovskite solar cells, ACS Appl.
(2014) 300–303.
Mater. Interfaces 11 (2019) 43116–43121.
[45] D. Pradhan, K.T. Leung, Vertical growth of two-dimensional zinc oxide
[18] L. Zhu, H. Li, P. Xia, Z. Liu, D. Xiong, Hierarchical ZnO decorated with CeO2
nanostructures on ITO-coated glass: effects of deposition temperature and
nanoparticles as the direct Z-scheme heterojunction for enhanced photocatalytic
deposition time, J. Phys. Chem. C 112 (2008) 1357–1364.
activity, ACS Appl. Mater. Interfaces 10 (2018) 39679–39687.
[46] X. Shi, X. Yang, X. Gu, H. Su, CuO–ZnO heterometallic hollow spheres: morphology
[19] A. Shafi, N. Ahmad, S. Sultana, S. Sabir, M.Z. Khan, Ag2S-Sensitized NiO–ZnO
and defect structure, J. Solid State Chem. 186 (2012) 76–80.
heterostructures with enhanced visible light photocatalytic activity and acetone
[47] F.H. Assaf, S.S.A. El-Rehim, A.M. Zaky, Cyclic voltammetric behaviour of–brass in
sensing property, ACS Omega 4 (2019) 12905–12918.
alkaline media, Br. Corrosion J. 36 (2001) 143–150.
[20] A. Pickett, A.A. Mohapatra, S. Ray, Q. Lu, G. Bian, K. Ghosh, S. Patil, S. Guha,
[48] S. Gilani, M. Ghorbanpour, A.P. Jadid, Antibacterial activity of ZnO films prepared
UV–Ozone modified sol–gel processed ZnO for improved diketopyrrolopyrrole-
by anodizing, J. Nanostructure Chem. 6 (2016) 183–189.
based hybrid photodetectors, ACS Appl. Electron. Mater. 1 (2019) 2455–2462.
[49] S. Xu, Z.L. Wang, One-dimensional ZnO nanostructures: solution growth and
[21] W. Shi, M.M. Ahmed, S. Li, Y. Shang, R. Liu, T. Guo, R. Zhao, J. Li, J. Du,
functional properties, Nano Res 4 (2011) 1013–1098.
Regulating the sensitivity and operating temperatures by morphology engineering
[50] X. Shu, H. Zheng, G. Xu, J. Zhao, L. Cui, J. Cui, Y. Qin, Y. Wang, Y. Zhang, Y. Wu,
of 2D ZnO nanostructures and 3D ZnO microstructures for the detection of organic-
The anodization synthesis of copper oxide nanosheet arrays and their
amines, ACS Appl. Nano Mater. 2 (2019) 5430–5439.
photoelectrochemical properties, Appl. Surf. Sci. 412 (2017) 505–516.
[22] A.H. Rakhsha, H. Abdizadeh, E. Pourshaban, M.R. Golobostanfard, V.R. Mastelaro,
[51] B.-S. Kim, T. Piao, S.N. Hoier, S.-M. Park, In situ spectro-electrochemical studies on
M. Montazerian, Ag and Cu doped ZnO nanowires: a pH-Controlled synthesis via
the oxidation mechanism of brass, Corrosion Sci. 37 (1995) 557–570.
chemical bath deposition, Materialia 5 (2019) 100212.
[52] M.M. Antonijevic, G.D. Bogdanovic, M.B. Radovanovic, M.B. Petrovic, A.
[23] A.M. Hallajzadeh, H. Abdizadeh, M. Taheri, M.R. Golobostanfard, Hierarchical
T. Stamenkovic, Influence of pH and chloride ions on electrochemical behavior of
porous Ga doped ZnO films synthesized by sol-electrophoretic deposition, Ceram.
brass in alkaline solution, Int. J. Electrochem. Sci. 4 (2009) 654–661.
Int. 46 (8, Part B) (2020) 12665–12674.
[53] N. Bellakhal, K. Draou, J.-L. Brisset, Plasma and wet oxidation of (63Cu37Zn)
[24] M. Suja, S.B. Bashar, M.M. Morshed, J. Liu, Realization of Cu-doped p-type ZnO
brass, Mater. Chem. Phys. 73 (2002) 235–241.
thin films by molecular beam epitaxy, ACS Appl. Mater. Interfaces 7 (2015)
8894–8899.

13
M. Nami et al. Materials Science in Semiconductor Processing 135 (2021) 106083

[54] F. Caballero-Briones, A. Palacios-Padrós, O. Calzadilla, F. Sanz, Evidence and based MOF-199 and their applications in visible-light-driven photocatalytic
analysis of parallel growth mechanisms in Cu2O films prepared by Cu anodization, degradation of dyes, J. Nanomater. 2019 (2019).
Electrochim. Acta 55 (2010) 4353–4358. [71] Y.-C. Chang, J.-Y. Guo, C.-M. Chen, H.-W. Di, C.-C. Hsu, Construction of CuO/In 2 S
[55] K. Ryczek, M. Kozieł, E. Wiercigroch, K. Małek, M. Jarosz, G.D. Sulka, L. Zaraska, 3/ZnO heterostructure arrays for enhanced photocatalytic efficiency, Nanoscale 9
Fast fabrication of nanostructured semiconducting oxides by anodic oxidation of (2017) 13235–13244.
brass, Mater. Sci. Semicond. Process. 113 (2020) 105035. [72] H. Chen, W. Leng, Y. Xu, Enhanced visible-light photoactivity of CuWO4 through a
[56] M.T. Qamar, M. Aslam, I.M.I. Ismail, N. Salah, A. Hameed, Synthesis, surface-deposited CuO, J. Phys. Chem. C 118 (2014) 9982–9989.
characterization, and sunlight mediated photocatalytic activity of CuO coated ZnO [73] H. Chen, Y. Xu, Photocatalytic organic degradation over W-rich and Cu-rich CuWO
for the removal of nitrophenols, ACS Appl. Mater. Interfaces 7 (2015) 8757–8769. 4 under UV and visible light, RSC Adv. 5 (2015) 8108–8113.
[57] J. Tauc, R. Grigorovici, A. Vancu, Optical properties and electronic structure of [74] G. Zhao, J. Zou, X. Chen, T. Zhang, J. Yu, S. Zhou, C. Li, F. Jiao, Integration of
amorphous germanium, Phys. Status Solidi 15 (1966) 627–637. microfiltration and visible-light-driven photocatalysis on a ZnWO4 nanoparticle/
[58] S. Das, T.L. Alford, Structural and optical properties of Ag-doped copper oxide thin nickel–aluminum-layered double hydroxide membrane for enhanced water
films on polyethylene napthalate substrate prepared by low temperature purification, Ind. Eng. Chem. Res. 59 (2020) 6479–6487.
microwave annealing, J. Appl. Phys. 113 (2013) 244905. [75] M. Nami, A. Rakhsha, S. Sheibani, H. Abdizadeh, The enhanced photocatalytic
[59] P. Sathishkumar, R. Sweena, J.J. Wu, S. Anandan, Synthesis of CuO-ZnO activity of ZnO nanorods/CuO nanourchins composite prepared by chemical bath
nanophotocatalyst for visible light assisted degradation of a textile dye in aqueous precipitation, Mater. Sci. Eng. B 271 (2021) 115262.
solution, Chem. Eng. J. 171 (2011) 136–140. [76] M.T. Qamar, M. Aslam, I.M.I. Ismail, N. Salah, A. Hameed, Synthesis,
[60] A. Naseri, M. Samadi, N.M. Mahmoodi, A. Pourjavadi, H. Mehdipour, A. characterization, and sunlight mediated photocatalytic activity of CuO coated ZnO
Z. Moshfegh, Tuning composition of electrospun ZnO/CuO nanofibers: toward for the removal of nitrophenols, ACS Appl. Mater. Interfaces 7 (2015) 8757–8769,
controllable and efficient solar photocatalytic degradation of organic pollutants, https://doi.org/10.1021/acsami.5b01273.
J. Phys. Chem. C 121 (2017) 3327–3338. [77] D.S. Eissa, S.S. El-Hagar, E.A. Ashour, N.K. Allam, Electrochemical nano-patterning
[61] F. Cao, T. Wang, X. Ji, Enhanced visible photocatalytic activity of tree-like ZnO/ of brass for stable and visible light-induced photoelectrochemical water splitting,
CuO nanostructure on Cu foam, Appl. Surf. Sci. 471 (2019) 417–424. Int. J. Hydrogen Energy 44 (2019) 14588–14595.
[62] X. Liu, B. Du, Y. Sun, M. Yu, Y. Yin, W. Tang, C. Chen, L. Sun, B. Yang, W. Cao, [78] E. Erdem, Microwave power, temperature, atmospheric and light dependence of
Sensitive room temperature photoluminescence-based sensing of H2S with novel intrinsic defects in ZnO nanoparticles: a study of electron paramagnetic resonance
CuO–ZnO nanorods, ACS Appl. Mater. Interfaces 8 (2016) 16379–16385. (EPR) spectroscopy, J. Alloys Compd. 605 (2014) 34–44.
[63] M. Aslam, I.M.I. Ismail, T. Almeelbi, N. Salah, S. Chandrasekaran, A. Hameed, [79] H. Irie, K. Kamiya, T. Shibanuma, S. Miura, D.A. Tryk, T. Yokoyama, K. Hashimoto,
Enhanced photocatalytic activity of V2O5–ZnO composites for the mineralization Visible light-sensitive Cu (II)-grafted TiO2 photocatalysts: activities and X-ray
of nitrophenols, Chemosphere 117 (2014) 115–123. absorption fine structure analyses, J. Phys. Chem. C 113 (2009) 10761–10766.
[64] C. Florica, A. Costas, N. Preda, M. Beregoi, A. Kuncser, N. Apostol, C. Popa, [80] R. Wang, J. Shen, W. Zhang, Q. Liu, M. Zhang, H. Tang, Build-in electric field
G. Socol, V. Diculescu, I. Enculescu, Core-shell nanowire arrays based on ZnO and induced step-scheme TiO2/W18O49 heterojunction for enhanced photocatalytic
Cu x O for water stable photocatalysts, Sci. Rep. 9 (2019) 1–15. activity under visible-light irradiation, Ceram. Int. 46 (2020) 23–30.
[65] A.B. Djurišić, Y.H. Leung, Optical properties of ZnO nanostructures, Small 2 (2006) [81] K. Sahu, A. Bisht, S. Kuriakose, S. Mohapatra, Two-dimensional CuO-ZnO
944–961. nanohybrids with enhanced photocatalytic performance for removal of pollutants,
[66] N. Abdullayeva, C.T. Altaf, M. Mintas, A. Ozer, M. Sankir, H. Kurt, N.D. Sankir, J. Phys. Chem. Solid. 137 (2020) 109223.
Investigation of strain effects on photoelectrochemical performance of flexible ZnO [82] Z. Liu, H. Bai, S. Xu, D.D. Sun, Hierarchical CuO/ZnO “corn-like” architecture for
electrodes, Sci. Rep. 9 (2019) 1–14. photocatalytic hydrogen generation, Int. J. Hydrogen Energy 36 (2011)
[67] F. Ansari, S. Sheibani, M. Fernández-García, Characterization and performance of 13473–13480.
Cu 2 O nanostructures on Cu wire photocatalyst synthesized in-situ by chemical [83] A.G. Acedo-Mendoza, A. Infantes-Molina, D. Vargas-Hernández, C.A. Chávez-
and thermal oxidation, J. Mater. Sci. Mater. Electron. 30 (2019) 13675–13689. Sánchez, E. Rodríguez-Castellón, J.C. Tánori-Córdova, Photodegradation of
[68] P. Du, A. Bueno-Lopez, M. Verbaas, A.R. Almeida, M. Makkee, J.A. Moulijn, G. Mul, methylene blue and methyl orange with CuO supported on ZnO photocatalysts: the
The effect of surface OH-population on the photocatalytic activity of rare earth- effect of copper loading and reaction temperature, Mater. Sci. Semicond. Process.
doped P25-TiO2 in methylene blue degradation, J. Catal. 260 (2008) 75–80. 119 (2020) 105257.
[69] K. Mageshwari, D. Nataraj, T. Pal, R. Sathyamoorthy, J. Park, Improved [84] F. Bekena, D.-H. Kuo, 10 nm sized visible light TiO2 photocatalyst in the presence
photocatalytic activity of ZnO coupled CuO nanocomposites synthesized by reflux of MgO for degradation of methylene blue, Mater. Sci. Semicond. Process. 116
condensation method, J. Alloys Compd. 625 (2015) 362–370. (2020) 105152.
[70] T.T. Minh, N.T.T. Tu, T.T. Van Thi, L.T. Hoa, H.T. Long, N.H. Phong, T.L.M. Pham,
D.Q. Khieu, Synthesis of porous octahedral ZnO/CuO composites from Zn/Cu-

14

You might also like