You are on page 1of 12

Photocatalytic Performance of ZnO=g-C3N4 for Removal of

Phenol under Simulated Sunlight Irradiation


Noor Izzati Md Rosli 1; Sze-Mun Lam 2; Jin-Chung Sin 3;
Ichikawa Satoshi 4; and Abdul Rahman Mohamed 5
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Abstract: A ZnO=g-C3 N4 photocatalyst was successfully synthesized via an impregnation method. The high-resolution transmission
electron microscope (HRTEM) result shows that g-C3 N4 had coated the surface of ZnO, indicating a sufficient blending between ZnO
and g-C3 N4 during synthesis. Reduced photoluminescence (PL) intensity of the ZnO=g-C3 N4 reveals the synergic effect of the hybrid nano-
particles on the recombination of photoinduced electron–hole pairs. The resultant ZnO=g-C3 N4 photocatalysts show higher degradation
efficiency of phenol than that of pure ZnO under simulated sunlight irradiation. Especially, the ZnO=ð5 wt%Þ g-C3 N4 demonstrate the highest
photodegradation efficiency of phenol because of the high amount of •OH radicals generated in the photocatalytic reaction. Process param-
eters such as photocatalyst amount, initial phenol concentration, and solution pH show a remarkable effect on the photocatalytic activity of
ZnO=g-C3 N4 . The degradation mechanism of phenol is also detailed through identification of product intermediates and detection of reactive
species. Moreover, the ZnO=g-C3 N4 photocatalyst could be reused several times without appreciable loss of activity, showing great potential
to be an excellent candidate for environmental remediation. DOI: 10.1061/(ASCE)EE.1943-7870.0001300. © 2017 American Society of
Civil Engineers.
Author keywords: Zinc oxide (ZnO); Graphitic carbon nitride (g-C3 N4 ); Heterojunction; Photocatalytic; Phenol; Simulated sunlight.

Introduction is only limited to within the ultraviolet (UV) light range and thereby
impedes its practical application under the visible-light portion of
In recent years, the application of solar energy conversion and envi- solar light. In addition, its poor charge separation signifies a major
ronmental remediation by semiconductor photocatalysts has at- loss of energy that results in low quantum efficiency. Therefore,
tracted research attention (Chai et al. 2009). TiO2 is by far the appreciable efforts have been attempted to overcome the recombi-
most widely used photocatalyst for this purpose. As an alternative nation of photogenerated electron–hole pairs such as doping (Qiu
to TiO2 , ZnO has also appeared to be a very promising photoca- et al. 2008b, 2011), semiconductor coupling (Habib et al. 2013;
talyst for degradation of organic and inorganic pollutants in aque- Lam et al. 2014a; Zhang et al. 2004), and noble-metal loading
ous systems. ZnO is an important photocatalyst because of its (Liqiang et al. 2004; Lu et al. 2008).
nontoxic nature, high photocatalytic activity, relatively low cost, In order to improve sunlight harvesting, an effective strategy
and environmentally stable behavior. In some cases, it exhibited that incorporated a ZnO-based photocatalyst and other visible-
higher efficiency than TiO2 in photocatalytic degradation of or- light active photocatalyst has been proposed. Graphitic carbon ni-
ganic pollutants (Wang et al. 2009a, c; Yu and Yu 2008). Regard- tride (g-C3 N4 ) has become a promising photocatalyst because of
less of its great application in photocatalytic uses, the major its good photocatalytic activity, medium band-gap energy
drawback of ZnO comes from its large band gap (∼3.2 eV), which (∼2.7 eV), high resistance to thermal and chemical environments,
and interesting electronic properties (Yan et al. 2009; Wang et al.
1
Ph.D. Student, School of Chemical Engineering, Universiti Sains 2009a, c). Given its π-conjugated properties, g-C3 N4 can act as an
Malaysia, 14300 Nibong Tebal, Pulau Pinang, Malaysia. E-mail: electron sink, leading to suppression of recombination of photo-
izzatirosli90@yahoo.com excited charge carriers. The flexible two-dimensional (2D) struc-
2
Senior Lecturer, Dept. of Environmental Engineering, Faculty of ture of g-C3 N4 allows it to easily coat other materials and possibly
Engineering and Green Technology, Universiti Tunku Abdul Rahman, form a heterojunction. Therefore, forming a heterojunction be-
Jalan Universiti Bandar Barat, 31900 Kampar, Perak, Malaysia. E-mail: tween ZnO and g-C3 N4 could result in an excellent photocatalytic
lamsm@utar.edu.my
3 material with enhanced photogenerated charge separation because
Senior Lecturer, Dept. Petrochemical Engineering, Faculty of Engi-
neering and Green Technology, Universiti Tunku Abdul Rahman, Jalan of the well-matched and overlapping band structures possessed by
Universiti Bandar Barat, 31900 Kampar, Perak, Malaysia. E-mail: sinjc@ both materials.
utar.edu.my Thus, the present study reports on the synthesis of a
4
Professor, Institute for Nanoscience Design, Osaka Univ., 1-3 Machi- ZnO=g-C3 N4 photocatalyst via a simple impregnation method us-
kaneyama, Toyonaka, Osaka, Japan. E-mail: ichikawa@insd.osaka-u.ac.jp ing melamine as a precursor without any surfactant. The as-
5
Professor, School of Chemical Engineering, Universiti Sains Malaysia, prepared photocatalysts are evaluated for the degradation of phenol
14300 Nibong Tebal, Pulau Pinang, Malaysia (corresponding author). under simulated sunlight irradiation. To date, the use of photoca-
E-mail:chrahman@usm.my
talysts (ZnO=g-C3 N4 ), particularly for phenol degradation, has
Note. This manuscript was submitted on February 3, 2017; approved on
July 5, 2017; published online on November 17, 2017. Discussion period scarcely been reported. This study is believed to be the pioneering
open until April 17, 2018; separate discussions must be submitted for work on the catalytic properties, photocatalytic activity, and recy-
individual papers. This paper is part of the Journal of Environmental clability of ZnO=g-C3 N4 in phenol degradation under simulated
Engineering, © ASCE, ISSN 0733-9372. sunlight irradiation.

© ASCE 04017091-1 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Experiment Details centrifugation, washed with deionized water several times, dried
in an oven at 60°C for 17 h, and then added to fresh phenol solution.
Preparation of ZnO =g-C 3 N 4
All reagents were of analytical grade without further purification. Active Species Analysis
The detailed procedures were as follows. First, 1.0 g of commercial The formation of hydroxyl radicals (•OH) throughout the photoca-
ZnO (Acros Organics, Geel, Belgium) and an appropriate amount talytic degradation experiments was investigated by using a tereph-
of melamine (0, 1, 5, and 7 wt%) was added to a methanol solution. thalic acid photoluminescence probing approach. The principle of
Then, the suspension was sonicated for 30 min for uniform this technique was that the terephthalic acid reacted with the •OH to
dispersion and stirred for 24 h. After that, the resultant mixture form 2-hydroxyterepthalic acid (HTA), which showed a signal at
was filtered, washed several times with deionized water, dried 425 nm. The experiment was carried out according to the process
for 17 h at 60°C in air, and then calcined for 2 h at 500°C. described by Sin et al. (2013a). The influence of active species to
the photocatalytic degradation of phenol was also explored by add-
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Material Characterization ing 2 mM of ethanol (EtOH), p-benzoquinone (BQ), and sodium


iodide (NaI) to scavenge •OH, superoxide anion radicals (O•− 2 ), and
The crystal structure of the samples was investigated using X-ray positively charged holes (hþ ), respectively, in a manner similar to
diffraction (XRD) analysis on a Philips PW1820 diffractometer the mentioned former photocatalytic experiment.
with Cu Kα radiation at a scanning rate of 2° min−1 in the range
of 20–80°. The morphological features and crystal microstructure
were acquired by a Leo Supra 50VP field-emission scanning Results and Discussion
electron microscopy (FESEM) (Germany) and Fei Tecnai 20
high-resolution transmission electron microscope (HRTEM). The
thermo scientific Fourier transform infrared spectroscopy (FTIR) Characterization of As-Prepared Photocatalysts
1510 (Belgium) was used to obtain the infrared spectrum at a scan- Fig. 1 shows the XRD patterns of pure ZnO and ZnO=g-C3 N4 pho-
ning range from 400 to 4,000 cm−1 . The specific surface area tocatalysts. All diffraction peaks in the pattern were identified and
was analyzed by a micrometric ASAP 2020 through nitrogen matched well with typical hexagonal wurtzite structures Joint Com-
adsorption-desorption at 77 K on the basis of the Brunauer– mittee on Powder Diffraction (JCPDS 36-1451). Observably, the
Emmett–Teller (BET) equation. Optical properties of the samples sharp intense peaks of ZnO appeared at 31.8°, 34.4°, 36.3°,
were carried out using a PerkinElmer Lambda UV-visible spectro- 47.5°, 56.6°, 62.8°, 66.3°, 67.9°, 69.0°, 72.5°, and 76.9°, which cor-
photometer (Waltham, Massachusetts) with BaSO4 as the refer- responded to (100), (002), (101), (012), (110), (103), (200), (112),
ence. Photoluminescence emission studies (PL) were recorded (201), (004), and (202), which confirms the good crystalline nature
by a PerkinElmer Lambda S55 spectrofluorometer at room temper- of the as-prepared photocatalyst. Meanwhile, the absence of the
ature at 25°C using a Xe lamp at an excitation wavelength of characteristic peaks of g-C3 N4 is probably a result of the relatively
325 nm. The chemical compositions and oxidation states of the low loading of g-C3 N4 used for the synthesis of ZnO=g-C3 N4 pho-
sample were analyzed by X-ray photoelectron spectroscopy tocatalyst (Xing et al. 2015; Liu et al. 2012; Li et al. 2014). In ad-
(XPS) Kratos Analytical Axis-ULTRA photoelectron spectroscopy dition, the enlarged XRD patterns (inset in Fig. 1) show no obvious
(Manchester, U.K.) with Al Kα radiation. In addition, the zeta shift in the diffraction peaks of all the ZnO=g-C3 N4 , indicating that
potential of the photocatalyst was conducted using a Malvern the g-C3 N4 was probably deposited on the surface of the ZnO in-
Zetasizer (Malvern, U.K.) using 2.5 mL aqueous slurry (0.1 g=L). stead of incorporated into its lattice (Hao et al. 2016).
The morphology of the as-prepared photocatalyst was charac-
terized by FESEM and HRTEM as shown Fig. 2. Figs. 2(a–d) show
Photocatalytic Performance Evaluation that the surfaces of the ZnO nanoparticles are densely packed with
Under simulated sunlight irradiation, the photocatalytic perfor-
mance of the as-prepared photocatalysts was evaluated by the deg-
radation of phenol. In each experiment, 100 mL of aqueous phenol
solution (5 mg=L, pH 5.7) containing required photocatalyst
(100 mg) was kept in the dark for 30 min to achieve adsorption
equilibrium. Subsequently, the solution was irradiated under a
simulated sunlight (Philips, 80 W, Merck, Selangor, Malaysia).
The solution was magnetically stirred, and air bubble was set at
a constant flow rate of 10 mL=min throughout the experiment.
At the given irradiation time intervals, 2 mL of the solution was
sampled and centrifuged to remove the particles, and the concen-
tration of phenol was monitored by a high performance liquid
chromatography (HPLC) (PerkinElmer Series 200, Waltham, Mas-
sachusetts). The experimental analysis by HPLC was carried out
according to the process described by Sin et al. (2014).
In the meantime, comparison studies using commercial TiO2
(Merck, 100% anatase, Darmstadt, Germany) were also performed.
In order to evaluate the mineralization of phenol, the total organic
carbon (TOC) was measured by a Shimadzu TOC-VCPH (Japan).
Fig. 1. XRD patterns of (a) pure ZnO; (b) ZnO=ð1 wt%Þ g-C3 N4 ;
The photocatalytic stability testing was conducted in a manner sim-
(c) ZnO=ð5 wt%Þ g-C3 N4 ; (d) ZnO=ð7 wt%Þ g-C3 N4 photocatalysts;
ilar to the photocatalytic test, and it underwent four consecutive
inset figure shows the magnified diffraction peak at (101)
cycles. After each cycle, the photocatalysts were collected by

© ASCE 04017091-2 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Surface morphologies of (a) pure ZnO; (b) ZnO=ð1 wt%Þ g-C3 N4 ; (c) ZnO=ð5 wt%Þ g-C3 N4 ; (d) ZnO=ð7 wt%Þ g-C3 N4 by FESEM;
(e) HRTEM images of ZnO=ð5 wt%Þ g-C3 N4 photocatalyst

© ASCE 04017091-3 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. FTIR spectra of (a) pure ZnO; (b) ZnO=ð1 wt%Þ g-C3 N4 ; Fig. 4. UV-visible diffuse reflectance spectra of as-prepared
(c) ZnO=ð5 wt%Þ g-C3 N4 ; (d) ZnO=ð7 wt%Þ g-C3 N4 ; (e) g-C3 N4 photocatalysts
photocatalysts

Table 1. Summary of Band-Gap Energy, BET Surface Area, Phenol


g-C3 N4 . The average diameter of the nanoparticles was determined Adsorption, and Degradation Values of As-Prepared Photocatalysts
to be 200–400 nm. Meanwhile, the addition of g-C3 N4 to ZnO BET Phenol Phenol
did not result in any significant changes of overall morphology Band-gap surface adsorption degradation
because the loading of the g-C3 N4 was very low for altering the energy area over over
morphology. Furthermore, it was difficult to differentiate between Sample (eV) (m2 g−1 ) 30 min (%) 60 min (%)
ZnO particles and g-C3 N4 , which may be attributable to the coat- Pure ZnO 3.30 4.63 8.6 79.8
ing of g-C3 N4 on the surface of ZnO particles. The added g-C3 N4 ZnO=ð1 wt%Þ g-C3 N4 3.28 3.13 7.0 89.1
could be dispersed as a g-C3 N4 nanosheet with an ultrathin single- ZnO=ð5 wt%Þ g-C3 N4 3.28 2.56 1.3 99.5
layered structure on the surface of ZnO by ultrasonic dispersion ZnO=ð7 wt%Þ g-C3 N4 3.28 1.04 8.4 93.7
during the preparation method (Bu and Chen 2014).
In Fig. 2(e), the HRTEM image of ZnO=ð5 wt%Þ g-C3 N4
clearly reveals that ZnO nanoparticles were decorated by g-C3 N4 ZnO=g-C3 N4 photocatalysts, suggesting the existence of g-C3 N4
with a thickness of ∼4.2 nm. The HRTEM observations also indi- and ZnO.
cate that sufficient blending between ZnO and g-C3 N4 was attained Fig. 4 depicts the optical properties of the as-prepared photoca-
during synthesis and formed a ZnO=g-C3 N4 photocatalyst. Despite talysts determined by UV-visible diffuse reflectance spectroscopy.
having undergone ultrasonic treatment before the HRTEM analy- The reflectance band edge of the ZnO sample occurs at 375 nm,
sis, a close interface still existed between ZnO and g-C3 N4 , sug- and the band gap is calculated based on Planck’s equation to be
gesting the possible formation of a heterojunction. The g-C3 N4 approximately 3.3 eV. After the addition of g-C3 N4 , the band edge
acted as a bridge connecting different ZnO nanoparticles, which of ZnO=g-C3 N4 photocatalysts was slightly extended to the longer
could be beneficial for the separation of photogenerated charge car- wavelength region, with the band-gap value calculated as provided
riers. Moreover, the lattice spacing of ZnO was measured to be in Table 1. The result also shows that the coupled-ZnO phototca-
0.28 nm, which corresponded to the (010) lattice planes of the talysts possess similar band-gap energy, which is 3.28 eV. This con-
ZnO wurtzite phase [Fig. 2(e)]. dition may be caused by the small difference in loading contents
In order to get more information about the g-C3 N4 , the FTIR impregnated onto the ZnO. This observation also implies that the
spectrum was taken to investigate the bond structure of addition of g-C3 N4 did not change the band-gap energy of ZnO, but
ZnO=g-C3 N4 photocatalysts. Fig. 3 shows that the characteristics changed the reflectance intensities of ZnO=g-C3 N4 photocatalysts
peaks at 1,246, 1,325, 1,410, and 1,623 cm−1 can be ascribed to (Wang et al. 2011). The previously mentioned UV-visible diffuse
the stretching vibration of C–N heterocycles (Boonprakob et al. reflectance spectra (DRS) results also show that the synthesized
2014; Chai et al. 2012). In addition, the peak detected at samples can absorb in UV and visible-light regions, which can
808 cm−1 was assigned to the characteristic breathing mode of be promising in sunlight photocatalysis.
s-triazine units (Li et al. 2014; Boonprakob et al. 2014). A broad Surface chemical composition and chemical states of as-
band at 508 cm−1 observed in all samples can be ascribed to the prepared photocatalyst were characterized using XPS, and the
stretching vibration of the Zn-O bond, whereas the broad band at results are shown in Fig. 5. Fig. 5(a) shows the total survey spec-
3,000–3,500 cm−1 can be ascribed to the stretching vibration of trum for ZnO and ZnO=ð5 wt%Þ gC3 N4 ; these indicate the pres-
N-H and O-H of physically adsorbed water (Ren et al. 2014; Jo ence of C, N, Zn, and O elements. The Zn 2p XPS spectrum
and Selvam 2015). In addition, the absorbance intensity of as- [Fig. 5(b)] shows two peaks at 1,021.0 and 1,044.1 eV, which can
prepared photocatalyst was stronger than that of g-C3 N4 , which be assigned to Zn 2p3=2 and 2p1=2 , respectively. In addition, at the
indicates the formation of a composite between ZnO and O 1s XPS spectrum [Fig. 5(c)] the peak centered at 530.0 eV can
g-C3 N4 . (Yan and Yang 2011). Furthermore, the absorbance inten- be related to the O2− ions in ZnO, whereas the shoulder peak at
sity at 1,246–1,623 cm−1 was observed to be increasing as the 532.0 eV can be associated to the OH− group absorbed onto the
amount g-C3 N4 loadings increased. This result further shows that photocatalyst’s surface (Ji et al. 2013). In Fig. 5(d), the high
the main characteristic peaks of g-C3 N4 and ZnO appeared in resolution of C 1s spectra for ZnO=ð5 wt%Þ g-C3 N4 sample

© ASCE 04017091-4 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. XPS spectra of the as-prepared ZnO=g-C3 N4 : (a) survey of the sample; (b) Zn 2p; (c) O 1s; (d) C 1s; (e) N 1s

exhibited peaks at 284.6 and 285.0 eV (Bu and Chen 2014), 400.0 eV, which correspond to C─N═C and N─ðCÞ3 , respectively
which correspond to sp2 C─C bonds of graphitic carbon and (Le et al. 2016; Xing et al. 2015). Based on the HRTEM, FTIR,
C─OH bond configurations, respectively. As shown in Fig. 5(e), and XPS results, it can be concluded that ZnO=g-C3 N4 photoca-
weak binding energy peaks can be observed at 398.0 and talyst had been successfully prepared.

© ASCE 04017091-5 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


degradation of phenol was negligible throughout the 60-min of si-
mulated sunlight irradiation, indicating the photocatalysis experi-
ments were conducted in nearly pure photocatalytic conditions.
Interestingly, under the same experimental conditions, the decrease
in phenol concentration is more prominent with ZnO=g-C3 N4 pho-
tocatalysts than with pure ZnO and commercial TiO2 . The photo-
catalytic degradation of phenol improved remarkably with
increased g-C3 N4 loading. When the loading reached 5 wt%,
the photocatalyst exhibited the highest photocatalytic activity, with
99.5% of phenol removal achieved after 60 min of irradiation. This
implies that g-C3 N4 plays a vital role in the improvement of photo-
catalytic activity. However, when the g-C3 N4 loading was further
increased to 7 wt%, the photocatalytic performance decreased. This
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

result shows that the optimal g-C3 N4 loading was at 5 wt%.


It is well known that the photocatalytic degradation is mainly
influenced by the phase structure, surface area, and effective sep-
Fig. 6. Phenol degradation using various photocatalysts (½phenol ¼ aration of photogenerated electron and holes. The SEM and XRD
5 mg=L; ½ZnO=g-C3 N4  ¼ 1 g=L; pHphenol ¼ 5.7) analyses show that the crystallinity of ZnO does not change after
modification by g-C3 N4. As indicated in Table 1, the corresponding
BET surface areas of ZnO=g-C3 N4 photocatalysts did not have no-
ticeable differences, which implies that the surface area was not the
Evaluation of Photocatalytic Activity main governing factor of photocatalytic activity. It is believed that
A series of experiments on phenol degradation under simulated the trapping sites of charge carriers increased with the increase of
sunlight irradiation was carried out to assess the photocatalytic ac- g-C3 N4 up to 5wt%, which can prolong the lifetime of carriers and
tivity of the as-prepared photocatalysts. Phenol is known as an promote photocatalytic activity. On the other hand, further increas-
endocrine-disruptor chemical and can be found widely in various ing the g-C3 N4 loading could favor the recombination of electron
industries like petrochemicals and oil refineries, and products like and hole instead of providing an electron pathway (Liu et al. 2012;
disinfectants and pesticides (Yavuz et al. 2008). Because of its ex- Chen et al. 2008). Hence, ZnO=ð5 wt%Þ g-C3 N4 displayed the
tensive use and recalcitrance in nature, phenol has become the big- highest photocatalytic activity for phenol degradation. These
gest threat to human health and the environment. Hazardous effects charge carriers’ recombination behaviors have been verified using
caused by phenol can be both acute and chronic because of its tox- PL analysis.
icity, persistence in the environment, and bioaccumulation (Dang The HPLC signals in Fig. 7 showed that the peak appeared at a
et al. 2016). Exposure to phenol by humans can result in the irri- retention time (RT) of 5.6 min, corresponding to phenol. This peak
tation of the skin, eyes, and mucous membranes. Animal studies gradually decreased with increasing reaction time. In addition, the
with phenol solutions have reported fetal bodyweight reductions, peaks corresponding to benzoquinone, resorcinol, and hydroqui-
growth retardation, and abnormal development in offspring none compared with standard chemicals were identified at 3.3,
(Villegas et al. 2016). Thus, it is necessary to degrade phenol in 2.5, and 2.2 min, respectively. The same findings were also re-
industrial waste. ported in other studies (Górska et al. 2009; Guo et al. 2006). Based
The photocatalytic performance of the as-prepared photocata- on the identification of the degradation intermediates, the proposed
lyst, pure ZnO, and P25 TiO2 were studied as shown in Fig. 6. degradation pathway is shown in Fig. 8. At the initial stage of deg-
In the absence of photocatalyst and in dark conditions, the radation, several aromatic intermediates like hydroquinone and

Fig. 7. HPLC signals of phenol treated by ZnO=ð5 wt%Þ g-C3 N4 photocatalyst as a function of time

© ASCE 04017091-6 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Proposed phenol photodegradation pathway over the ZnO=g-C3 N4 photocatalysts

resorcinol were formed from the reaction process between active ZnO (Fig. S1) and the pKa of phenol, which were 8.9 and 9.9,
species and phenol. The presence of benzoquinone was attributed respectively (Lam et al. 2012).
to the dehydrogenation of hydroquinone (Villaseñor et al. 2002; It was found that the degradation efficiency increases from
Sin et al. 2014). After undergoing ring cleavage reactions, these pH 3.0 to 5.7 and decreases at pH 10. The result for the degradation
aromatic intermediates would form aliphatic acid and ultimately of phenol for different solution pH values can be seen in Fig. 9(c).
mineralize to form CO2 and H2 O (Sin et al. 2014). According to point of zero charges of ZnO, the expected surface
reaction at different pH values is shown in Eqs. (1) and (2) (Sin
Effect of Photocatalyst Amount et al. 2013a)

The effect of photocatalyst amount on the degradation of phenol pH < pHpzc ∶ZnOH þ Hþ → ZnOHþ
2 ð1Þ
was investigated over ZnO/(5 wt%), and the result is shown in
Fig. 9(a). It can be seen that with an increase in photocatalyst pH > pHpzc ∶ZnOH þ OH− → ZnO− þ H2 O ð2Þ
amount, the photocatalytic degradation of phenol increased from
0.028 min−1 at 0.25 g=L to 0.076 min−1 at 1.0 g=L and then de- where ZnOH = metal hydroxyl group. Hence, it is reasonable to
creased to 0.042 min−1 at 2.0 g=L. Increasing the photocatalyst expect that the different electrostatic interaction between photoca-
amount increases the available active sites for phenol degradation talyst and phenol will affect the photocatalytic activity of the photo-
(Barzgari et al. 2016). Above 1.0 g=L ZnO=ð5 wt%Þ g-C3 N4 , catalyst. In an acidic medium, the ZnO surface was exhibited as
lower photocatalytic degradation was observed because of excess positively charged and phenol usually acted as neutral in its
suspension of the photocatalysts that intercept light (Yusoff un-ionized form. Thus, at this condition the Cl− anions possibly
et al. 2014). competing with phenol molecules on available active sites, thereby
reducing the photocatalytic activity (Lam et al. 2010). On the other
hand, at the natural pH of 5.7, no pH adjustment is needed. Hence,
Effect of Initial Phenol Concentration the electrostatic attraction force occurring between phenol mole-
The effect of initial phenol concentration was evaluated over cules and positively charged ZnO surface lead to high photocata-
1.0 g=L ZnO=ð5 wt%Þ g-C3 N4 . The result in Fig. 9(b) clearly lytic activity.
shows that photocatalytic degradation of phenol decreases when On the other hand, phenol molecules undergo deprotonation and
the initial phenol concentration increases from 5 to 30 mg=L. become negatively charged ions at high pH. Electrostatic repulsion
Increasing the phenol concentration can increase the available phe- occurring between phenolate molecules and the negatively charged
nol molecules adsorbed on the photocatalyst surface. Thus, more photocatalyst surface of ZnO causes a reduction in phenol adsorp-
reactive species are required to degrade phenol. At constant light tion onto the photocatalyst surface, resulting in a drop of photoca-
intensity, photocatalyst amount, and duration of irradiation, the talytic efficiency. In addition, the lower photocatalytic activity in an
amount of reactive species practically remained the same. Hence, alkaline medium is also influenced by competition among adsorp-
at higher concentrations, the available reactive species were inad- tion sites between ions from NaOH salts and phenolate molecules
equate to degrade phenol (Habib et al. 2013; Yang and Tian 2012). (Lam et al. 2012). Therefore, pH 5.7 was identified as an optimum
In addition, low degradation efficiency in higher concentration is pH solution for the degradation of phenol.
also attributed to the decrease in path length of photon entering the In summary, influencing factors like g-C3 N4 loadings (5 wt%),
phenol solution (Sin et al. 2013b). ZnO=g-C3 N4 amount (1 g=L), phenol concentration (5 mg=L), and
solution pH (pH 5.7) can be controlled to achieve a highest perfor-
mance for degradation of phenol.
Effect of Solution pH
The photocatalytic degradation of phenol under different pH
TOC Removal and Photocatalyst Recycling
medium was studied with initial pH ranging from 3.0 to 10.0
and adjusted by an equimolar NaOH or HCl aqueous solution while In order to evaluate the extent of mineralization of phenol, TOC
keeping all other experimental conditions constant. The range of monitoring was conducted for the irradiated solution of phenol.
solution pH covered the pH at point of zero charge (pzc) of Fig. 10(a) shows that only 56.0% of TOC removal achieved almost

© ASCE 04017091-7 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. (a) Effect of photocatalyst amount on photocatalytic degradation of phenol over ZnO=ð5 wt%Þ g-C3 N4 (½phenol ¼ 5 mg=L; pH ¼ 5.7);
(b) effect of initial phenol concentration on photocatalytic degradation of phenol over ZnO=ð5 wt%Þ g-C3 N4 (photocatalyst amount ¼ 1 g=L;
pH ¼ 5.7); (c) effect of solution pH on photocatalytic degradation of phenol over ZnO=ð5 wt%Þ g-C3 N4 (½phenol ¼ 5 mg=L; photocatalyst
amount ¼ 1 g=L)

complete phenol degradation. The findings indicated that a com- suggesting that the ZnO=g-C3 N4 photocatalyst has high stability.
plete removal of phenol does not necessarily mean complete min- Thus, ZnO=g-C3 N4 can be used as a potential photocatalyst in
eralization, but rather a transient transformation into another practical applications for environmental purification.
compound occurred, as revealed by the results of HPLC chromato-
gram in Fig. 7. Thus, it appears that a longer irradiation time is Mechanism Considerations
required to achieve complete mineralization of phenol.
Furthermore, the stability of a photocatalyst is important for Photoluminescence Spectra
its assessment in practical applications. The cycling runs for the Photoluminescence is generally related to the electronic, optical,
degradation of phenol using ZnO=ð5 wt%Þ g-C3 N4 photocatalyst and photoelectric properties of materials. Generally, the recombi-
were performed under the same experimental conditions to evaluate nation of photogenerated charge carriers in a semiconductor can
its stability. As shown in Fig. 10(b), there is no obvious changed release energy in the form of PL emissions. In particular, the lowest
in the photodegradation of phenol after four consecutives cycles. emission intensity indicates a lower recombination of charge car-
The slightly decreased photodegradation could be attributed to riers, leading to higher photocatalytic activity (Le et al. 2017). The
the photocatalyst loss, which reduced the active sites. The FESEM PL spectra in Fig. 11 show a significant quenching of PL intensity
image of reused photocatalyst in Fig. 10(c) shows that the morphol- for all ZnO=g-C3 N4 photocatalysts compared with pure ZnO. Such
ogy of the ZnO=ð5 wt%Þ g-C3 N4 photocatalyst was preserved a decrease in intensity might be ascribed to the efficient charge-
without distinct changes after four cycles of experiments, injection process between ZnO and g-C3 N4 .

© ASCE 04017091-8 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. (a) Photocatalytic degradation and TOC removal of phenol over ZnO=ð5 wt%Þ g-C3 N4 (½phenol ¼ 5 mg=L; ½ZnO=g-C3 N4  ¼ 1 g=L;
pHphenol ¼ 5.7); (b) reusability test of phenol over ZnO=ð5 wt%Þ g-C3 N4 ; (c) SEM image of ZnO=ð5 wt%Þ g-C3 N4 photocatalyst after being
recycled four times

It can also be observed that the PL intensity of ZnO=g-C3 N4


photocatalyst changed when varying the g-C3 N4 loading. In par-
ticular, the PL intensity of ZnO=ð5 wt%Þ g-C3 N4 exhibited the
lowest PL intensity, which is highly consistent with the photoca-
talytic activity result (Fig. 6). This result implies that the charge
carriers have a longer lifetime and generate more-active •OH rad-
icals in the photocatalytic reaction.

Hydroxyl Radical Test


The formation of •OH radicals on the simulated sunlight irradiated
ZnO products was validated using terephthalic acid–photolumines-
cence (TA-PL) technique (Zhang et al. 2009; Qiu et al. 2008a, b).
Fig. 12 displays the pure ZnO and ZnO=g-C3 N4 photocatalysts’ PL
spectra under simulated sunlight irradiation in the basic solution of
TA. Usually, the amount of •OH radicals was determined from the
PL intensity at approximately 425 nm. The figure reveals the for-
mation of •OH radicals, which indeed participate in the photoca-
talytic degradation process. The amount of •OH radicals produced
Fig. 11. PL spectra of (a) pure ZnO; (b) ZnO=ð1 wt%Þ
by the ZnO=ð5 wt%Þ g-C3 N4 photocatalyst was higher than that of
g-C3 N4 ; (c) ZnO=ð5 wt%Þ g-C3 N4 ; (d) ZnO=ð7 wt%Þ g-C3 N4
other samples, which implies higher photocatalytic activity of the
photocatalysts
former compared with the latter. Moreover, the higher PL intensity

© ASCE 04017091-9 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


shown by all the ZnO=g-C3 N4 photocatalysts compared with pure
ZnO suggests that the charge transfer was accelerated. Likewise,
the coupling of g-C3 N4 with ZnO inhibited the recombination
of e− -hþ pairs, resulting in the increase of •OH formation to con-
tribute in the photocatalytic degradation of phenol (Cui et al. 2012).

Free-Radical and Hole-Scavenging Experiments


To understand the mechanism for the photodegradation of phenol,
individual influences of active species such as hydroxyl radicals
(•OH), superoxide radicals (O•− þ
2 ), and holes (h ) were explored.
Different scavengers were used individually in the photocatalytic
reaction to quench a specific reactive species. The scavenger used
in this study are ethanol for •OH scavenger (Lam et al. 2014b),
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

sodium iodide (NaI) to quench hþ (Lam et al. 2014b), and p-ben-


zoquinone (BQ) for O•− 2 scavenger (Ji et al. 2009). The effect of
scavenger on the photocatalytic reaction is shown in the Fig. 13.
As can be seen, the addition of ethanol, NaI, and BQ caused
the photocatalytic degradation of phenol decreased from 99.5 to
Fig. 12. TA-PL spectra of (a) pure ZnO; (b) ZnO=ð1 wt%Þ g-C3 N4 ;
5.0, 6.0, and 10.9%, respectively. These results show the participa-
(c) ZnO=ð5 wt%Þ g-C3 N4 ; (d) ZnO=ð7 wt%Þ g-C3 N4 photocatalysts in
tion of all radicals in the photocatalytic activity. Among the radi-
a basic TA solution for 60 min of simulated sunlight irradiation
cals, •OH played a more important role than hþ and O•− 2 in the
degradation of phenol.

Possible Photocatalytic Mechanism


On the basis of the conducted experiments’ results and related lit-
erature (Liu et al. 2012, 2013; Chen et al. 2014), a proposed mecha-
nism for the enhanced photocatalysis of ZnO=g-C3 N4 is shown in
Fig. 14. In the present study, both ZnO and g-C3 N4 can be excited
under simulated sunlight and produced electron and hole pairs
(e− -hþ ) as shown in Eq. (3). The photogenerated e− can migrate
from the conduction band (CB) of g-C3 N4 to the conduction band
of ZnO because the CB edge of ZnO [−0.5 eV versus normal hy-
drogen electrode (NHE)] was lower than CB edge of g-C3 N4
(−1.13 eV versus NHE) (Sun et al. 2012). Meanwhile, the O2 ad-
sorbed on the surface of the photocatalyst was catalyzed by the
accumulated e− in the CB of ZnO to yield O•− 2 radicals and then
converted to •OH radicals according to Eqs. (4)–(7)
þ
ðZnO=g-C3 N4 Þ þ hv → ðZnO=g-C3 N4 Þðe−
CB þ hVB Þ ð3Þ
Fig. 13. Effect of different radical scavengers toward the degradation
of phenol over ZnO=ð5 wt%Þ g-C3 N4 photocatalyst (½ZnO=g-C3 N4  ¼
1.0 g=L; ½phenol ¼ 5 mg=L; pHphenol ¼ 5.7) e− þ O2 → O•− ð4Þ
2

Fig. 14. Schematic illustration of energy-band structure and electron–hole transfer in the ZnO=g-C3 N4 photocatalyst

© ASCE 04017091-10 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


Table 2. Comparison of the Previous Data on Phenol Degradation with Results of Present Study
Experimental
condition
Catalyst
Phenol loading
Photocatalyst (mg L−1 ) (g L−1 ) Source of light Degradation rate Reference
ZnO 75 2.500 Solar light Complete degradation achieved within 480 Pardeshi and Patil (2008)
Pt=ZnO-SiO2 100 1.000 150 W mercury lamp Complete degradation achieved within 60 min Mohamed and Barakat (2012)
ZnO 50 2.000 6 W mercury UV lamp Complete degradation took place within 80 min Lathasree et al. (2004)
Fe3 O4 -ZnO N/A 0.325 757 W metal halide lamp 80% removal achieved within 150 min Feng et al. (2014)
Ag/ZnO 25 1.000 UV-C lamp Complete degradation took place in 150 min Wang et al. (2004)
ZnO=ð5 wt%ÞC3 N4 5 1.000 80 W simulated sunlight Almost complete degradation within 60 min Current study
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

O•− þ
2 þ H → HOO• ð5Þ Acknowledgments
This work was supported by a Research Universiti Grant
HOO• þ Hþ þ e− → H2 O2 ð6Þ
(No. 814176) and International Research Collaboration Fund
(No. 910404) from Universiti Sains Malaysia as well as a scholar-
H2 O2 þ e− → •OH þ OH− ð7Þ ship given by the Ministry of Higher Education Malaysia.
Simultaneously, the photogenerated hþ in the valence band
(VB) of ZnO transported to VB of g-C3 N4 . Considering the VB
of g-C3 N4 (1.57 V versus NHE) was more negative than the stan- Supplemental Data
dard reduction potential of E°ð•OH=OH− Þ ¼ 2.72 eV versus NHE
Fig. S1 is available online in the ASCE Library (www
and E°ð•OH=H2 OÞ ¼ 1.99 eV versus NHE, the transported hþ
.ascelibrary.org).
cannot oxidize the H2 O and OH− to •OH radicals (Liu et al. 2013;
Ji et al. 2009). For the case of ZnO, the photogenerated hþ on the
photocatalyst surface experiences charge transfer with adsorbed
water molecules or bounded hydroxide species to produce •OH References
radicals according to Eqs. (8) and (9)
Barzgari, Z., Ghazizadeh, A., and Askari, S. Z. (2016). “Preparation of
þ þ Mn-doped ZnO nanostructured for photocatalytic degradation of orange
H2 O þ h → H þ •OH ð8Þ G under solar light.” Res. Chem. Intermed., 42(5), 4303–4315.
Boonprakob, N., et al. (2014). “Enhanced visible-light photocatalytic ac-
OH− þ hþ → •OH ð9Þ tivity of g-C3 N4 =TiO2 films.” J. Colloid Interface Sci., 417, 402–409.
Bu, Y., and Chen, Z. (2014). “Highly efficient photoelectrochemical anti-
Thus, the improved photocatalytic activity of ZnO=g-C3 N4 can corrosion performance of C3 N4 @ZnO composite with quasi-shell–core
be attributed to the synergistic effect arising from the heterojunc- structure on 304 stainless steel.” RSC Adv., 4(85), 45397–45406.
tion between ZnO and g-C3 N4 , as well as the matched energy-band Chai, B., Peng, T., Mao, J., Li, K., and Zan, L. (2012). “Graphitic carbon
structure that improves the separation efficiency of photogenerated nitride ðg-C3 N4 Þ-Pt-TiO2 nanocomposite as an efficient photocatalyst
for hydrogen production under visible light irradiation.” Phys. Chem.
charge carriers. Ultimately, Table 2 provides a comparison among
Chem. Phys., 14(48), 16745–16752.
the current work and other reported studies on phenol degradation Chai, S. Y., Kim, Y. J., Jung, M. H., Chakraborty, A. K., Jung, D., and
with other ZnO photocatalysts. Lee, W. I. (2009). “Heterojunctioned BiOCl=Bi2 O3 , a new visible light
photocatalyst.” J. Catal., 262(1), 144–149.
Chen, D., et al. (2014). “Significantly enhancement of photocatalytic per-
Conclusions formances via core-shell structure of ZnO@mpg-C3 N4 .” Appl. Catal. B
Environ., 147, 554–561.
A series of ZnO and ZnO=g-C3 N4 photocatalysts were successfully Chen, S., Zhao, W., Liu, W., and Zhang, S. (2008). “Preparation, charac-
synthesized by a surfactant-free impregnation method using terization and activity evaluation of p-n junction photocatalyst
melamine as g-C3 N4 precursor. Compared with pure ZnO, the p-ZnO=n-TiO2 .” Appl. Surf. Sci., 255(5), 2478–2484.
photocatalytic degradation of phenol by ZnO=g-C3 N4 was en- Cui, Y., Huang, J., Fu, X., and Wang, X. (2012). “Metal-free photocatalytic
hanced as a result of the effective separation of photoinduced degradation of 4-chlorophenol in water by mesoporous carbon nitride
charge carriers and high formation of •OH radicals as verified semiconductors.” Catal. Sci. Technol., 2(7), 1396–1402.
by PL measurements. The ZnO=g-C3 N4 photocatalyst prepared Dang, T. T. T., Le, S. T. T., Channei, D., Khanitchaidecha, W., and
at 5 wt% exhibited the highest photocatalytic activity under simu- Nakaruk, A. (2016). “Photodegradation mechanisms of phenol in the
lated sunlight irradiation. The results from process parameters such photocatalytic process.” Res. Chem. Intermed., 42(6), 5961–5974.
as g-C3 N4 loading, photocatalyst dosage, initial concentration, and Feng, X., Guo, H., Patel, K., Zhou, H., and Lou, X. (2014). “High perfor-
mance, recoverable Fe3 O4 -ZnO nanoparticles for enhanced photocata-
solution pH significantly influenced the degradation of phenol. In
lytic degradation of phenol.” Chem. Eng. J., 244, 327–334.
addition, the ZnO=g-C3 N4 photocatalyst showed high stability and
Górska, P., Zaleska, A., and Hupka, J. (2009). “Photodegradation of phenol
can be reused four times with insignificant loss in photoactivity. by UV=TiO2 and Vis/N, C-TiO2 processes: Comparative mechanistic
The improved photocatalytic activity of ZnO=g-C3 N4 photocata- and kinetic studies.” Sep. Purif. Technol., 68(1), 90–96.
lyst was mainly attributed to the synergistic effects between Guo, Z., Ma, R., and Li, G. (2006). “Degradation of phenol by nanomaterial
ZnO and g-C3 N4 . The findings in this work can likely offer sug- TiO2 in wastewater.” Chem. Eng. J., 119(1), 55–59.
gestions for developing a new type of photocatalyst for environ- Habib, M. A., Shahadat, M. T., Bahadur, N. M., Ismail, I. M. I., and
mental purification. Mahmood, A. J. (2013). “Synthesis and characterization of

© ASCE 04017091-11 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091


ZnO-TiO2 nanocomposites and their application as photocatalysts.” Qiu, X., et al. (2008b). “Origin of the enhanced photocatalytic activities of
Int. Nano Lett., 3(1), 1–8. semiconductors: A case study of ZnO doped with Mg2þ .” J. Phys.
Hao, R., Wang, G., Tang, H., Sun, L., Xu, C., and Han, D. (2016). Chem. C, 112(32), 12242–12248.
“Template-free preparation of macro/mesoporous g-C3 N4 =TiO2 heter- Ren, H., et al. (2014). “Improved photochemical reactivities of
ojunction photocatalysts with enhanced visible light photocatalytic Ag2 O=g-C3 N4 in phenol degradation under UV and visible light.”
activity.” Appl. Catal. B: Environ., 187, 47–58. Ind. Eng. Chem. Res., 53(45), 17645–17653.
Ji, H., Chang, F., Hu, X., Qin, W., and Shen, J. (2013). “Photocatalytic Sin, J. C., Lam, S. M., Lee, K. T., and Mohamed, A. R. (2013a). “Photo-
degradation of 2, 4, 6-trichlorophenol over g-C3 N4 under visible light catalytic performance of novel samarium-doped spherical-like ZnO
irradiation.” Chem. Eng. J., 218, 183–190. hierarchical nanostructures under visible light irradiation for 2,
Ji, P., Zhang, J., Chen, F., and Anpo, M. (2009). “Study of adsorption and 4-dichlorophenol degradation.” J. Colloid Interface Sci., 401, 40–49.
degradation of acid orange 7 on the surface of CeO2 under visible light Sin, J. C., Lam, S. M., Lee, K. T., and Mohamed, A. R. (2013b). “Prepa-
irradiation.” Appl. Catal. B Environ., 85(3–4), 148–154. ration and photocatalytic properties of visible light-driven samarium-
Jo, W., and Selvam, N. C. S. (2015). “Enhanced visible light-driven photo- doped ZnO nanorods.” Ceram. Int., 39(5), 5833–5843.
catalytic performance of ZnO-g-C3 N4 coupled with graphene oxide as a Sin, J. C., Lam, S. M., Satoshi, I., Lee, K. T., and Mohamed, A. R. (2014).
Downloaded from ascelibrary.org by Glasgow University Library on 11/17/17. Copyright ASCE. For personal use only; all rights reserved.

novel ternary nanocomposites.” J. Hazard. Mater., 299, 462–470. “Sunlight photocatalytic activity enhancement and mechanism of novel
Lam, S. M., Sin, J. C., Abdullah, A. Z., and Mohamed, A. R. (2010). europium-doped ZnO hierarchical micro/nanospheres for degradation
“Parameter effect on photocatalytic degradation of phenol using of phenol.” Appl. Catal. B Environ., 148–149, 258–268.
TiO2 -P25/activated carbon (AC).” Korean J. Chem. Eng., 27(4), Sun, L., et al. (2012). “Enhanced visible-light photocatalytic activity of
1109–1116. g-C3 N4 -ZnWO4 by fabricating a heterojunction: Investigation based
Lam, S. M., Sin, J. C., Abdullah, A. Z., and Mohamed, A. R. (2012). on experimental and theoretical studies.” J. Mater. Chem., 22(44),
“Degradation of wastewaters containing organic dyes photocatalysed 23428–23438.
by zinc oxide: A review.” Desalin. Water Treat., 41(1–3), 131–169. Villaseñor, J., Reyes, P., and Pecchi, G. (2002). “Catalytic and photocata-
Lam, S. M., Sin, J. C., Abdullah, A. Z., and Mohamed, A. R. (2014a). lytic ozonation of phenol on MnO2 supported catalysts.” Catal. Today,
“Transition metal oxide loaded ZnO nanorods: Preparation, characteri- 76(2–4), 121–131.
zation and their UV-vis photocatalytic activities.” Sep. Purif. Technol., Villegas, L. G. C., et al. (2016). “A short review of techniques for phenol
132, 378–387. removal from wastewater.” Curr. Pollut. Reports, 2(3), 157–167.
Lam, S. M., Sin, J. C., Satoshi, I., Abdullah, A. Z., and Mohamed, A. R. Wang, J., et al. (2009a). “Relationship between oxygen defects and the
(2014b). “Enhanced sunlight photocatalytic performance over photocatalytic property of ZnO nanocrystals in nafion membranes.”
Nb2 O5 =ZnO nanorod composites and the mechanism study.” Appl. Langmuir, 25(2), 1218–1223.
Catal. A Gen., 471, 126–135. Wang, R., et al. (2004). “The characteristics and photocatalytic activities
of silver doped ZnO nanocrystallites.” Appl. Surf. Sci., 227(1–4),
Lathasree, S., Rao, A. N., SivaSankar, B., Sadasivam, V., and Rengaraj, K.
312–317.
(2004). “Heterogeneous photocatalytic mineralisation of phenols in
Wang, X., Maeda, K., Chen, X., Takanabe, K., and Domen, K. (2009c).
aqueous solutions.” J. Mol. Catal. A Chem., 223(1–2), 101–105.
“Polymer semiconductors for artificial photosynthesis: Hydrogen evo-
Le, S., et al. (2016). “Cu-doped mesoporous graphitic carbon nitride for
lution by mesoporous graphitic carbon nitride with visible light.” J. Am.
enhanced visible-light driven photocatalysis.” RSC. Adv., 6(45),
Chem. Soc., 131(5), 1680–1681.
38811–38819.
Wang, Y., Shi, R., Lin, J., and Zhu, Y. (2011). “ Enhancement of photo-
Le, S., et al. (2017). “Highly efficient visible-light-driven mesoporous gra-
current and photocatalytic activity of ZnO hybridized with graphite-like
phitic carbon nitride/ZnO nanocomposites photocatalysts.” Appl. Catal.
C3 N4 .” Energy Environ. Sci., 4(8), 2922–2929.
B Env., 200, 601–610.
Xing, H., et al. (2015). “Preparation of g-C3 N4 =ZnO composites and their
Li, X., et al. (2014). “Synergistic effect of efficient adsorption g-C3 N4 =ZnO enhanced photocatalytic activity.” Mater. Technol. Adv. Perform.
composite for photocatalytic property.” J. Phys. Chem. Solids, 75(3), Mater., 30(1), 122–127.
441–446. Yan, H., and Yang, H. (2011). “TiO2 -g-C3 N4 composite materials for pho-
Liqiang, J., et al. (2004). “Investigations on the surface modification of tocatalytic H2 evolution under visible light irradiation.” J. Alloy.
ZnO nanoparticle photocatalyst by depositing Pd.” J. Solid State Chem., Compd., 509(4), 126–129.
177(11), 4221–4227. Yan, S. C., Li, Z. S., and Zou, Z. G. (2009). “Photodegradation performance
Liu, W., Wang, M., Xu, C., and Chen, S. (2012). “Facile synthesis of of g-C3 N4 fabricated by directly heating melamine.” Langmuir, 25(17),
g-C3 N4 =ZnO composite with enhanced visible light photooxidation 10397–10401.
and photoreduction properties.” Chem. Eng. J., 209, 386–393. Yang, Y., and Tian, C. (2012). “Photocatalytic degradation of methylene
Liu, W., Wang, W., Xu, C., Chen, S., and Fu, X. (2013). “Significantly blue and phenol on Fe-doped sulfated titania.” Res. Chem. Intermed.,
enhanced visible-light photocatalytic activity of g − C3 N4 via ZnO 38(3–5), 693–703.
modification and the mechanism study.” J. Mol. Catal. A Chem., Yavuz, Y., Koparal, A. S., and Ogutveren, U. B. (2008). “Phenol degrada-
368–369, 9–15. tion in a bipolar trickle tower reactor using borron-doped diamond elec-
Lu, W., Gao, S., and Wang, J. (2008). “One-pot synthesis of Ag/ZnO self- trode.” J. Environ. Eng., 10.1061/(ASCE)0733-9372(2008)134:1(24),
assembled 3D hollow microspheres with enhanced photocatalytic per- 24–31.
formance.” J. Phys. Chem. C, 112(43), 16792–16800. Yu, J., and Yu, X. (2008). “Hydrothermal synthesis and photocatalytic ac-
Mohamed, R. M., and Barakat, M. A. (2012). “Enhancement of photoca- tivity of zinc oxide hollow spheres.”Environ. Sci. Technol., 42(13),
talytic activity of ZnO=SiO2 by nanosized Pt for photocatalytic 4902–4907.
degradation of phenol in wastewater.” Int. J. Photoenergy, 2012, 1–8. Yusoff, N., Ong, S. A., Ho, L.-N., Wong, Y. S., and Khalik, W. (2014).
Pardeshi, S. K., and Patil, A. B. (2008). “A simple route for photocatalytic “Degradation of phenol through solar-photocatalytic treatment by zinc
degradation of phenol in aqueous zinc oxide suspension using solar oxide in aqueous solution.” Desalin. Water Treat., 54(6), 1621–1628.
energy.” Sol. Energy, 82(8), 700–705. Zhang, D., et al. (2009). “Role of oxygen active species in the photocata-
Qin, H., Li, W., Xia, Y., and He, T. (2011). “Photocatalytic activity of lytic degradation of phenol using polymer sensitized TiO2 under visible
heterostructures based on ZnO and N-doped ZnO.” Appl. Mater. Inter- light irradiation.” J. Hazard. Mater., 163(2–3), 843–847.
faces, 3(8), 3152–3156. Zhang, M., An, T., Hu, X., Wang, C., Sheng, G., and Fu, J. (2004). “Prepa-
Qiu, R., et al. (2008a). “Photocatalytic activity of polymer-modified ZnO ration and photocatalytic properties of a nanometer ZnO-SnO2 coupled
under visible light irradiation.” J. Hazard. Mater., 156(1–3), 80–85. oxide.” Appl. Catal. A Gen., 260(2), 215–222.

© ASCE 04017091-12 J. Environ. Eng.

J. Environ. Eng., 2018, 144(2): 04017091

You might also like