Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

Version of Record: https://www.sciencedirect.

com/science/article/pii/S1226086X18301382
Manuscript_587e49ae78218facd6a913b59513032d

Improved photoelectric performance via fabricated


heterojunction g-C3N4/TiO2/HNTs loaded photocatalysts for
photodegradation of ciprofloxacin
Dongyao Wua, Jinze Lia, Jingru Guana, Chongyang Liua, Xiaoxu Zhaoa, Zhi Zhua,

Changchang Mab, Pengwei Huoa*, Chunxiang Lia, Yongsheng Yana*

a
Institute of Green Chemistry and Chemical Technology, School of Chemistry &

Chemical Engineering, Jiangsu University, Zhenjiang 212013, PR China

b
School of Environment, Jiangsu University, Zhenjiang 212013, PR China

* Corresponding author Tel.: +86 511 8879 0187 Fax. : +86 511 8879 1800

E-mail address: huopw@mail.ujs.edu.cn

© 2018 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
Abstract

An Intercalated heterostructural g-C3N4/TiO2/HNTs supported photocatalyst was

successfully prepared via sol-gel and calcination methods. The introduction of HNTs

and the g-C3N4-TiO2 heterojunction effectively enhanced the charge transfer and

separation efficiency of photogenerated electron-hole pairs, which endued the

g-C3N4/TiO2/HNTs hybrid material with an outstanding photoelectric performance

and good stability. And an obviously enhanced photocatalytic activity was exhibited

by photodegrading ciprofloxacin compared with pure TiO2. Furthermore, the main

active species were detected through trapping experiment and ESR spin-trap

technique with DMPO, which confirmed that the •O2− and the h+ were the main active

species in the photocatalytic system.

Keywords: TiO2; g-C3N4; Halloysite nanotubes; Heterojunction ; Photocatalytic

1. Introduction

In recent years, with the rapid development of industry, environmental pollution

problems become more and more serious, especially the water pollution. One of the

reasons of the water pollution is the abuse of antibiotics, which has not only caused the

environmental pollution, but also induces the bacterial resistance, jeopardizing human

health. [1-3] Therefore, the removal of antibiotic from the wastewater is of great

importance to people and environment. Nowadays, different methods are used to

remove the antibiotics from the wastewater, such as adsorption [4], electrochemical

method [5], microbial degradation [6], photocatalytic degradation [7, 8], and membrane

filtration [9]. Among them, semiconductor photocatalytic degradation has been


considered to be the most promising method, which has the advantage of completed

degradation the antibiotics to H2O, CO2 and other non-toxic or less toxic substances.

Up to now, in a variety of semiconductors, such as CdS, TiO2, Cu2O, BiVO4 and so

on [10-13], anatase TiO2 is one of the most studied photocatalytic materials, due to its

high catalytic activity, non-toxic, stability, inexpensive and anti-corrosive [14-18].

However, the application of TiO2 is largely restricted, owing to its wide band gap (about

3.2eV), which makes it less response to visible light and can only absorb the UV light

with the wavelength less than 320 nm and high recombination rate of photo-generated

electrons and holes, However, the utilization of visible light is greatly degraded for

TiO2 [19]. Consequently, various strategies should be devoted to enhancing the

efficiency in visible light utilization. Coupling TiO2 with visible-light catalysts to form

heterostructures has been considered as an effective strategy. As is known to all, the

photo-generated electrons and holes can be transferred to opposite directions by the

build-in electric field of heterojunction, thereby inhibiting their recombination.

Furthermore, the heterojunction consisting of two semiconductors photocatalytic can

superpose the photo-response range of the two photocatalytic. Therefore, it is necessary

to extend TiO2 catalysts to the visible-light response by selecting suitable visible light

catalyst. Recently, Graphitic carbon nitride (g-C3N4), with the narrow band gap and the

high thermal and chemical stability, leading to a good photocatalytic for organic

pollutant degradation, H2 evolution from water and photoreduction of CO2, has

attracted enormous attention[20, 21]. It has been reported that many g-C3N4 composite

photocatalysts coupled with different semiconductors photocatalysts have been


synthesized, such as, g-C3N4/BiPO4, g-C3N4/BiOBr, g-C3N4/ZnO, g-C3N4/WO3,

CdS/g-C3N4 and so on [22-26]. Since these composite semiconductors are mostly

doped with narrow band gaps semiconductors, which can enhance their photocatalytic

activity in visible light. Coupling TiO2 with g-C3N4 to form heterostructures has been

shown a great potential application in photocatalysis. The g-C3N4/TiO2 heterojunction

can extend the visible-light response range, promote photogenerated electron-hole pairs

separation, and significantly improve the visible-light photocatalytic activity of TiO2.

Li. et al. [27] have researched the growth of g-C3N4 on mesoporous TiO2 spheres with

high photocatalytic activity under visible light irradiation and high photoelectric

performance. Wei et al. [28] have reported a Z-scheme g-C3N4-TiO2 photocatalyst with

good photoelectric performance and photocatalytic activity. Thus, the introduction of

g-C3N4 to form heterojunctions photocatalysts is beneficial to enhance the photoelectric

properties, improve the utilization of visible light and the photocatalytic activity.

Nevertheless, TiO2 nanoparticles are prone to agglomerate together to form larger

particles, which will reduce the surface area and the active sites, resulting in a decrease

in photocatalytic activity. And therefore, many approaches have been taken to solve the

problem of the aggregation, such as coating technology [29], organic modification [30],

supported technology [31] etc. Among them loading TiO2 with other supported is the

most effective method, which would greatly improve light utilization and

photocatalytic activity with changing the morphology, surface area and other aspects,

like TiO2/GO, TiO2/CNTs, and TiO2/ATP [33-34]. As one kind of natural

aluminosilicate clay minerals supported, Halloysite nanotubes (HNTs) with a


double-layered structure ,and easy to obtain, reusable, and much low cost than other

supported composites, possesses the properties of hollow nanotube structures, large

pore volume and large specific surface areas. Moreover, there are sufficient hydroxyl

groups in the surface of HNTs, which can be used for the support of nanoparticles.

Wang et al. [35] had reported TiO2/HNTs composite by one-step solvothermal

method with and high photocatalytic activity on the degradation of methanol. Li et al.

[36] had investigated the g-C3N4-ZnO/HNT composite photocatalysts, which could

degradate tetracycline under visible light irradiation and possess a good photoelectric

performance. Thus, anchored TiO2 onto HNTs is a promising method to prevent the

aggregation and strengthen the photocatalytic activity in the degradation of pollutants.

In this study, we synthesized g-C3N4/TiO2/HNTs heterojunction photocatalysts by

sol-gel and calcination method. TiO2/HNTs were prepared by sol-gel method,

preferentially. The special morphology of HNTs, beneficial to that of the composite

material, could make it fully contact with pollutants. Next, g-C3N4 and TiO2/HNTs

were coupled to form heterojunctions photocatalysts byalcination method. Moreover,

the g-C3N4/TiO2/HNTs heterojunction photocatalys did not only increase the

photoelectric properties, but also improved the light utilization and photocatalytic

activity. A set of characterizations were used to investigate the structure, morphology,

composition, optical properties, photoelectric performance, stability, and mechanism

of the g-C3N4/TiO2/HNTs heterojunction composite. In addition, the photoelectric

performance was determined by IPCE and EIS. The photocatalytic activity was

confirmed by the degradation of antibiotics under visible light irradiation.


2. Experimental

2.1 Materials

Tetrabutyl titanate (C16H36O4Ti), acetic acid (CH3COOH), ethanol absolute

(C2H5OH), melamine (C3H6N6) were purchased from Sinopharm Chemical Reagent

Co., Ltd. (Shanghai, China). Halloysite nanotubes (HNTs) were supplied by

Zhengzhou Jinyangguang Chinaware Co., Ltd. Ciprofloxacin (CIP) were purchased

from Shanghai Aladdin Bio-Chem Technology Co., Ltd. All the chemicals reactants in

this work were analytical reagent and used without further purification. Deionized

water was used throughout this work without any purification.

2.2 Synthesis of TiO2/HNTs composite

The TiO2/HNTs composite was synthesized by sol-gel method [37]. First, 2 mL of

tetrabutyl titanate was added to 12 mL absolute ethanol to form a homogeneous

solution and stirred for 15 min at room temperature. 0.8 mL of Acetic acid was added to

the solution under vigorous constant stirring for 15 min. 0.4 g HNTs was added to the

mixture under continuous magnetic stirring for 1 h, subsequently. A mixture of 1.5 mL

distilled water and 1.5 mL absolute ethanol was slowly added to the suspension, and the

suspension was continuously stirred for 24 h. Afterwards, the sol-gel precursor was

dried at 60°C for 12 h. The sol-gel precursor was calcined in muffle furnace at 500°C

for 2 h at the heating rate of 5°C/min. The samples washed with distilled water and

absolute ethanol several times, were dried in an oven at 60°C for 12 h, and grounded
into a powder.

2.3 Synthesis of g-C3N4/TiO2/HNTs heterojunction composites

The g-C3N4/TiO2/HNTs heterojunction composites were synthesized by a calcination

method. In a typical synthesis, 1g TiO2/HNTs and melamine powder with different

mass ratio of 20%, 50%, 80%, 100% and 150% were mixed in an agate mortar and

grounded for 30 min, then were put in a muffle furnace at 500°C for 4 h with a heating

rate of 5°C min−1. When they were then cooled down to room temperature, they were

washed with distilled water and absolute ethanol several times, and dried in an oven at

60°C for 12 h. A series of g-C3N4/TiO2/HNTs composites with different mass ratios

were prepared, and the obtained samples were denoted CTH-0.2, CTH-0.5, CTH-0.8,

CTH-1.0 and CTH-1.5, respectively.

2.4 Characterization

The phase composition and crystal structure of the as-prepared catalysts were

collected by X-ray diffraction (XRD-6100Lab, Japan) with Ni-filtrated Cu-Kα

radiation. The 2-theta scan range of the sample was 5°-80°at a scanning rate of 5°/min.

The scanning electron microscopy (SEM) images were obtained by a field emission

scanning electron microscopy (FE-SEM; JSM-7001F, Japan) and operated at an

accelerating voltage of 200 kV. The energy dispersive X-ray spectroscopy (EDS) was

performed on a FEI Tecnai G2F20 instrument. The transmission electron microscope

(TEM) and high-resolution transmission electron microscopy (HRTEM) pattern and


the selected area electron diffraction (SAED) images were observed on transmission

electron microscopy (JEM-2010, Japan). The X-ray photoelectron spectroscopy (XPS)

measurements were detected on a PerkinElmer PHI 5300 instrument with Mg Ka

radiation. The diffuse reflectance spectra (DRS) of the as-prepared catalysts were

investigated using a UV–vis spectrophotometer (UV-2450; Shimazu, Japan) with an

integrating sphere in the range of 200–800 nm, in which BaSO4 was the reference

substance. FT-IR spectra were measured using a Fourier transform infrared (FT-IR)

spectrometer, and the samples were dispersed in KBr at a ratio of 1:100. The Raman

spectra of the as-prepared catalysts were determined by laser Raman spectrometer with

an excitation wavelength at 532 nm. Photoluminescence spectra (PL) of the

as-prepared catalysts were performed using a fluorescence spectrometer with an

excitation wavelength at 360 nm. The electron spin resonance (ESR) signals of •OH

and •O2- radicals were obtained on a Brucker A300 ESR spectrometer, and the radical

trap was 5, 5-diamethyl-1-pyrroline N-oxide (DMPO).

2.4 Photoelectrochemical measuremen

The photoelectochemical (PEC) performances of the as-prepared photocatalysts

were measured by an electrochemical workstation (CHI 660B, Shanghai, China) in a

standard three-electrode quartz cells, 0.5 moL/L Na2SO4 solution as electrolyte. The

as-prepared samples were used as the working electrode, platinum gauze was used as

counter electrode, and reference electrode was used as reference electrode.

Electrochemical impedance spectroscopy (EIS) was carried out in 0.5 moL/L Na2SO4
solution by a ZENNIUM electrochemical workstation (Zahner Instruments, Germany),

and the range of frequency was from 0.1 Hz to 100 kHz. Moreover, the

electrochemical signals of the as-prepared samples were measured by a CHI660 B

electrochemical analyzer (Chen Hua Instruments, Shanghai, China).

2.5 Photocatalytic measurements

The photocatalytic activities of as-prepared photocatalysts were evaluated by the

degradation effect of CIP under simulated solar irradiation. The light source of the

photocatalytic reaction was a 300 W Xe lamp. In order to maintain the temperature at

room temperature, the system was connected to the equipment of circulating water. In

the typical experiments, 80 mg of as-prepared photocatalysts was dispersed in 100 mL

of CIP solution (15 mg/L). The suspension was stirred in the dark for 30 min to

guarantee the adsorption-desorption balance of CIP on the photocatalysts. At a certain

time interval, 5 mL of the suspension was taken and centrifuged. The concentration of

CIP was detected by a UV/vis spectroscopy, and the maximum absorption wavelength

was 272 nm.

3. Results and discussion

3.1 XRD

The XRD patterns of as-prepared photocatalysts were shown in Fig. 1(a), which

provided the information about the phase composition and crystal structure. As

identified in Fig. 1(a), it could be found that the characteristic diffraction peaks at 2θ =
25.4°, 37.9°, 48.2°, 53.9°, 55.1°, 62.7°, 68.9°, 70.2° and 75.2°corresponding to the

(101), (004), (200), (105), (211), (204), (116), (220) and (215) crystal planes of anatase

TiO2(JCPDS No. 21-1272), respectively [16,28]. Halloysite was conformed with

Halloysite 7 Å and Halloysite 10 Å, and a small amount of quartz (JCPDS file number

46-1045) was also present in HNTs, which observed the two peaks at 2θ = 47.8° and

53.9°, and a stronger peak at 2θ = 25.3°. The characteristic diffraction peaks of

Halloysite 10 Å (JCPDS No. 291489) were appeared at 2θ = 20.1° and 26.6°, the peaks

of Halloysite 7Å (JCPDS No. 29-1487) observed at 2θ = 12.4°, 34.9° and 62.8° [38].

Different from halloysite, g-C3N4/TiO2/HNTs had no significant peaks at 2θ =12.4°,

20.1°, 34.9° and 62.8°, showing that the high temperature calcination partially

destroyed the HNTs structure [39]. The XRD pattern of pure g-C3N4 (JCPDS No.

87-1526) presented two characteristic diffraction peaks at 2θ = 13.1° (100) and 27.5°

(002), which were caused by the present of tri-s-triazine units and the interlayer

stacking of the conjugated aromatic system and the (100) plane [40,41], respectively.

As for g-C3N4/TiO2/HNTs composites, the characteristic diffraction peaks of anatase

TiO2, g-C3N4 and HNTs were observed, indicating that the g-C3N4/TiO2/HNTs

composites was successfully prepared. However, the peak of 13° were disappeared in

all composites, this observation could be attributed to low content or evenly dispersion

of g-C3N4, as well as the poor crystallization due to the presence of TiO2/HNTs during

the process of preparation [42, 43]. In addition, compared with pure TiO2, the

characteristic peaks of TiO2 were wider, suggesting that the average crystallite size of

TiO2 decreased [43, 44]. Moreover, from Fig. 1(b), g-C3N4/TiO2/HNTs composites
with different mass ratios of g-C3N4 were observed. As the mass ratios of g-C3N4

increased, the intensity of g-C3N4 at (002) plant slowly increased, the intensity of

anatase TiO2 decreases. It was worthwhile to note that the characteristic diffraction

peaks of CTH-0.2 composites cannot be observed, which was due to the low content or

evenly dispersion of g-C3N4 [42].

Fig. 1. XRD pattern of as-prepared (a) TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2,

g-C3N4/TiO2/HNTs composites; (b) Different mass ratio of g-C3N4/TiO2/HNTs

composites.

3.2 SEM and TEM

The morphology and microstructure of the as-prepared samples were investigated by

FESEM, EDS. The SEM and EDS images of pure TiO2, pure g-C3N4, HNTs,

TiO2/HNTs, g-C3N4/TiO2 and CTH-0.5 were given in Fig. 2. As can be seen from Fig.

2(a), TiO2 presented irregular block, due to the particles gathering together. Fig. 2(b)

displayed that g-C3N4 appeared irregular layers structure and aggregated together. Fig.

2(c) shown that HNTs present the smooth cylindrical tubes of empty lumen structure. In

comparison with pure TiO2 and HNTs, after loading TiO2 on HNTs as shown in Fig.

2(d), the TiO2 nanoparticles were irregularly dispersed on the surface of HNTs, which

significantly reduced the aggregation of TiO2 and made the HNTs become rougher. Fig.

2(e) shown the SEM images of g-C3N4/TiO2, it could be observed the surface of g-C3N4

became rougher after combining with TiO2. Fig. 2(f) shown the SEM images of
CTH-0.5, it could be observed that the HNTs with TiO2 nanoparticles was deposited on

the g-C3N4, and the CTH-0.5 was intercalated structure. The inset of Fig. 2(f) was the

EDS, displayed that the CTH-0.5 are composed of the elements of Ti, O, C, N, Si, Al. In

addition, The mass and atomic ratios of C, N, O, Si, Al and Ti element of as-obtained

samples were given in Table 1.

In order to further observe the morphology and composites, the TEM analysis was

ere shown in Fig.3. From Fig. 3(a) and 3(b), it could be seen that the pure TiO2 present

about 10~30 nm spherical nanoparticles, and the g-C3N4 showed the layers structure,

respectively. As shown in Fig. 3(c), the TiO2 were surely deposited on the surface of

HNTs. Fig. 3(d) showed that the TiO2 nanoparticles appeared on the surface of g-C3N4.

For g-C3N4/TiO2/HNTs composite in Fig. 3(e), TiO2 and HNTs were embedded in the

layers structure of g-C3N4 and formed the g-C3N4/TiO2/HNTs intercalated

heterojunction. The heterojunction structure would promote photogenerated

electron-hole pairs separation and the visible light photocatalytic activity. To further

investigate the heterojunction structure between g-C3N4 and TiO2, the HR-TEM is

shown in Fig. 3(f). It could be observed that the interface between g-C3N4 and TiO2 are

tightly combined together. In addition, some lattice fringes could also be observed, the

lattice fringes of 0.35 nm in agreement well with the (101) lattice plane of the anatase

TiO2 [46, 47]. And the lattice fringes of 0.19 nm could be assigned to the (200) lattice

plane of the anatase TiO2 [48, 49]. The inset of Fig. 3(f) showed a typical SEAD

pattern of the composites. These results demonstrate that TiO2 and HNTs were

successfully combined with g-C3N4, and formed the heterojunction between g-C3N4
and TiO2, which was consistent with XRD result.

Fig. 2. SEM images of (a) TiO2, (b) g-C3N4, (c) HNTs, (d) TiO2/HNTs, (e)

g-C3N4/TiO2, (f) CTH-0.5 and the insets of (f) is the EDS of CTH-0.5 composite.

Fig. 3. TEM images of (a) TiO2, (b) g-C3N4, (c) TiO2/HNTs, (d) g-C3N4/TiO2, (e)

CTH-0.5; HR-TEM image of (f) CTH-0.5 and the insets of (f) is the SEAD of

CTH-0.5 composite.

Table 1 The mass and atomic ratios of CTH-0.5 composite.

3.3 XPS

The surface chemical compositions and valence states of the as-prepared CTH-0.5

were characterized by XPS. The survey scan spectrum in Fig. 4 indicated the presence

of Ti, O, C, N, Si and Al in the sample, which proved the chemical compositions of

CTH-0.5. Among them, the Si and Al were mainly derived from the HNTs support. As

depicted in Fig. 5(a), four overlapping peaks could be obtained after deconvolution of

C 1s. The peak at 284.8eV was ascribed to the adventitious carbon. The peaks at 286.4

eV, 288.2 eV and 288.9 eV could be fitted to C-N bands, (N) 2-C=N bands, and

O-C=O bands, respectively [50, 51]. From Fig. 5(b), the N 1s peak could be

deconvolved into four peaks. The peaks at 398.7 eV, 399.5 eV, 401.1 eV and 404.4 eV

was attached to the C-N=C bands, N-(C) 3 bands, H-N-H bands, and the charging

effects, respectively [52, 53]. In the O 1s spectrum (Fig. 5(c)), two peaks could be

separated at 532.1 eV and 530.0 eV, which could be assigned to Ti-O bands, and the
surface hydroxyl groups [54, 55]. As shown in Fig. 5(d), the peaks at 458.7 eV and

464.4 eV were attributed to the Ti 2p3/2 and Ti 2p1/2 of the Ti4+ in TiO2 [46, 54]. The

high-resolution scanning of Si 2p displayed in Fig. 5(e) shown a single peak at 103.1 eV,

corresponding to the Si-O bond of HNTs. In the Al 2p spectrum (Fig. 5(f)), one peak

could be separated at 75.1 eV, which could be assigned to Al-O bands of HNTs[56, 57].

The XPS results confirmed that the g-C3N4/TiO2/HNTs composite was composed of

g-C3N4, TiO2 and HNTs.

Fig. 4. XPS whole-range spectra of the CTH-0.5 composite.

Fig. 5. XPS spectra of (a) C 1s, (b) N 1s, (c) O 1s, (d) Ti 2p (e) Si 2p and (f) Al 2p of

CTH-0.5 composite.

3.4 FT-IR

The compositions of the as-prepared samples were further confirmed by the FT-IR

spectra. It could be seen in Fig. 6, in the FT-IR spectrum of pure TiO2, the absorption

peak at 460-1180 cm-1 corresponded to the vibration of Ti-O-Ti [58]. For the HNTs, the

double peaks at 3696 cm-1 and 3624 cm-1 were assigned to the stretching vibrations of

Al-OH group at the inner-surface of HNTs [39, 54]. After forming the composites, the

double peaks at 3696 cm-1 and 3624 cm-1 disappeared, which further confirmed that the

structure of HNTs was destroyed. The peak at around 1632 cm-1 was due to the

deformation vibration of the interlayer water [59]. The peaks at 1096 cm-1, 1033 cm-1

and 468 cm-1 are ascribed to the stretching vibration of Si-O, asymmetrical stretching
vibration of Si-O-Si and the bending vibration of Si-O-Si, respectively [60, 61]. The

peaks at 910 cm-1 and 539 cm-1 were owing to the deformation vibration of Al-O-Si of

surface hydroxyl groups and O-H deformation [58]. In composites, the peaks of HNTs

from 500 cm-1 to 1100 cm-1 could not been observed, and may be overlapped by the

peaks of TiO2 [58]. As for pure g-C3N4, the broad peak at 2500-3500 cm-1 corresponded

to the stretching vibration modes of N-H groups [62]. The peak at 805 cm−1 was

associated with the typical breathing mode of s-triazine. The three peaks at 1554 cm−1,

1406 cm−1, 1406 cm−1, 1320 cm−1 and 1242 cm−1 were ascribed to the stretching

vibration of C-N heterocycle [63]. As for g-C3N4/TiO2 and g-C3N4/TiO2/HNTs

composite, the peaks of TiO2, g-C3N4 could be observed, demonstrating that the

g-C3N4/TiO2/HNTs composites are successfully prepared.

Fig. 6. FT-IR spectra of TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2, CTH-0.5

composites.

3.5 BET

The specific surface areas of pure TiO2, pure g-C3N4 and CTH-0.5 composite were

determined by nitrogen adsorption-desorption strategy. The higher surface area of

photocatalysts are, the better the photocatalytic performance is. This was because it

would produce more active sites. The N2 adsorption-desorption isotherms and pore

size distributions pattern were presented in Fig. 7. It could be observed that the pure

TiO2 displayed a type IV adsorption-desorption isotherm and H2 hysteresis loops,


implying the existence of mesoporous. The pure HNTs displayed a type IV

adsorption-desorption isotherm, indicating both mesoporous and micropores (IUPAC

classification). The pure g-C3N4 and CTH-0.5 composite showed type IV

adsorption-desorption isotherm and H3 hysteresis loops, suggesting that the present of

macropores [64]. In addition, the BET surface area, pore volume and pore size were

shown in Table 2. The BET surface areas of pure TiO2, pure g-C3N4, pure HNTs and

CTH-0.5 were 44.862 m2/g, 9.772 m2/g, 58.45 m2/g and 39.648 m2/g, respectively.

The BET surface area of HNTs was the highest when compared to TiO2, g-C3N4 and

CTH-0.5 composite. In CTH-0.5 composite, TiO2 and g-C3N4 covered the cavities of

HNTs resulting in a decrease in the BET surface area[35]. The BET surface area of

CTH-0.5 was higher than pure g-C3N4, indicating that the formation of CTH-0.5

composite enhanced the photocatalytic performance of g-C3N4. In addition, the BET

surface area of CTH-0.5 was smaller than pure TiO2, which might be because some

TiO2 could be wrapped in the g-C3N4, and the particle sizes of g-C3N4 and HNTs were

much higher than TiO2, thus after forming composite, while the surface area of

CTH-0.5 was smaller than TiO2[65].

Fig. 7. N2 adsorption-desorption isotherms of TiO2, g-C3N4, CTH-0.5 composite and

the inset is corresponding pore-size distribution curves.

Table 2 BET surface area, pore size and pore volume of pure TiO2, pure g-C3N4, HNTs
and g-C3N4-TiO2/HNTs samples

3.6 Raman
The Raman spectra of TiO2, g-C3N4, and CTH-0.5 samples are shown in Fig. 8,

respectively. Symmetry groups analysis indicated that there should be 15 fundamental

optical modes of anatase TiO2 in Raman: Γ= 1A1g+1A2u+2B1g+1B2u+3Eg+2Eu. The

modes of A1g, B1g, Eg were the Raman typical active modes. Therefore, it was expected

that there are six fundamental transitions of the anatase in Raman spectroscopy [65]. In

the Raman spectra of TiO2, the peaks at 142 and 195 cm−1 were ascribed to the Ti-Ti

vibrational modes, the peaks at 395, 515 and 638 cm−1 attributed to the Ti-O vibrational

modes [54, 67]. For pure g-C3N4, the peaks at 478, 705, 749 and 1231 cm−1 were

assigned to the vibration modes of the triazine ring, the peak at 980 cm−1 ascribed to the

breathing modes of the triazine ring [58,59]. In addition, the characteristic peaks are

weak due to the interference of fluorescence [68]. In the Raman spectra of CTH-0.5

composites, the characteristic peaks of TiO2 and g-C3N4 could be observed, and the

peak at 142 cm−1 of the pure TiO2 was red shift to 145 cm−1 (inset of Fig. 5), which

meant that a close interaction existed between TiO2 and g-C3N4[66].

Fig. 8. Raman spectra of TiO2, g-C3N4, CTH-0.5 composite.

3.7. UV-vis diffuse reflection spectra

An ultraviolet-visible (UV–vis) diffuse reflectance spectroscopy was used to study

the optical absorption properties of pure TiO2, pure g-C3N4, HNTs and CTH-0.5

composite. As can be seen from Fig. 9(a), pure TiO2 presents an absorption edge at 380

nm, the pure g-C3N4 showing its absorption edge at 460 nm whereas the HNTs exhibits
hardly any optical properties. Compared with pure TiO2, the absorption edge of

g-C3N4/TiO2, CTH-0.5 composite showed increasing red shifts in the visible light

region, and the highest absorption peak of CTH-0.5 composite was significantly

increased. Additionally, the band structure of as-prepared samples could be estimated

by the transformed Kubelka-Munk equation and valence band XPS [46, 69]. As shown

in Fig .9(b), the band gaps of TiO2 and g-C3N4 were found to be 3.2 and 2.7 eV,

respectively. Compared with TiO2 and g-C3N4, the band gap of CTH-0.5 composite was

narrowed to 2.5 eV. It indicated that coupling TiO2, HNTs with g-C3N4 to form

composite could extend the optical absorption properties. which producing more

electron-hole pair [70]. For CTH-0.5 composite, the reduced band gap perhaps be

attributed to the interfacial synergistic interaction between HNTs and semiconductors,

and the Chemical bonding between TiO2 and g-C3N4 [36, 38, 71]. As illustrated in Fig.

10., the valence band (VB) of the pure TiO2 and g-C3N4 were determined to be 1.58

eV, 2.91 eV, respectively. Then the conduction bands (CB) of the pure TiO2 and

g-C3N4 were respectively calculated to be -1.12 eV and -0.29eV by considering their

band gap energies.

Fig. 9. (a) UV-vis diffuses reflectance spectra of TiO2, g-C3N4, CTH-0.5 composites;

(b) Kubelka-Munk function transformed differential spectra of TiO2, g-C3N4,

CTH-0.5 composite.

Fig. 10. XPS valence band spectra of the pure TiO2 and g-C3N4.
3.8 Photoelectrochemical measurements

The photocurrent was widely used to demonstrate the photoinduced electrons-holes

pair separation efficiency of the semiconductor materials [72]. And generally the

stronger the photocurrent is, the faster the photoinduced electrons-holes pairs

separation efficiency is, which could enhance the photocatalytic activity [73]. In order

to investigate the photoelectochemical (PEC) performance, a series of transient

photocurrent responses of pure TiO2, pure g-C3N4, TiO2/HNTs, g-C3N4/TiO2 and

CTH-0.5 composites with four light-off and light-on cycles of visible-light irradiation

in the photocurrent-time curves were shown in the photocurrent-time curves. As shown

in Fig. 11(a), it could be observed that TiO2/HNTs , g-C3N4/TiO2 and CTH-0.5

composite composites showed a stronger photocurrent density than TiO2 and g-C3N4

under the visible-light irradiation, which CTH-0.5 composite possessed the highest

photocurrent among the four samples. It demonstrated that CTH-0.5 composite

possesses a higher separation efficiency of photoinduced electrons-holes pairs, lower

rate of photogenerated charge carrier recombination and extends the lifetime of

photo-generated charge carriers, which might be due to the formation of

heterojunctions between g-C3N4 and TiO2 and the loaded of HNTs [66].

In addition, the electrochemical impedance spectroscopy (EIS) measurement was

used to further investigate the transfer of the charge and separation efficiency of

photoinduced electrons-holes pairs at semiconductor interface. The EIS Nyquist plot of

pure TiO2, pure g-C3N4, TiO2/HNTs, g-C3N4/TiO2and CTH-0.5 composites were

shown in Fig. 11(b). It was known that the semicircle in the high frequency region
correlates to the resistance during the charge transfer process at the photoelectrode

interface, and the smaller radius represents the smaller resistance [74]. The arc radius of

CTH-0.5 composite was smaller comparing with other samples, which meant the

smaller resistance of CTH-0.5 composite. It indicated that more effective separation of

photoinduced electrons-holes pairs. These results demonstrated that the formation of

heterojunctions between TiO2 and g-C3N4 could obviously enhance the transfer and

separation efficiency of photogenerated electron-hole pairs in CTH-0.5 composite,

thereby improving the photocatalytic activity. The result was consistent with

photoelectochemical (PEC) performance measurement.

Fig. 11. (a) Transient photocurrent responses of TiO2, g-C3N4, TiO2/HNTs,

g-C3N4/TiO2, and CTH-0.5 composites; (b) EIS spectra of TiO2, g-C3N4, TiO2/HNTs,

g-C3N4/TiO2, and CTH-0.5 composites.

3.9 PL spectral

As is well-known, the PL spectrum was widely applied to evidence the efficiency of

charge transfer and trapping, as well as the recombination rate of photoexcited

electrons and holes in semiconductor [49, 75]. Fig. 12(a) showed the PL spectra of the

TiO2, g-C3N4, CTH-0.5 composite. The samples were excited at 360 nm. The pure

g-C3N4 possessed a strongest emission peak at about 480 nm, which revealed that the

recombination rate of photoexcited electrons and holes is highest. The pure TiO2 had an

emission peak at about 440nm [76]. Compared with pure g-C3N4 and TiO2, the PL
intensity of CTH-0.5 composite was obviously reduced, implying a lower

recombination rate of photoexcited electrons and holes. The result of PL spectra

demonstrated that the heterojunctions between TiO2 and g-C3N4 reduced the

recombination rate of photoexcited electrons and holes, and improved the efficiency of

charge transfer, thereby enhanced the photocatalytic activity.

The charge carriers had also been explored by the transient fluorescence. The

fluorescence decay curves were obtained through multiple index fitting. As shown in

Fig. 12(b), the fluorescent lifetime of g-C3N4. TiO2 and CTH-0.5 composite were

6.17ns, 0.83ns and 0.63ns, respectively. The fluorescent lifetime of CTH-0.5

composite was much shorter than that of TiO2 and g-C3N4, suggesting that an

additional nonradiative decay channel maybe opened by the transfer of electrons from

TiO2 to g-C3N4, and suppressed the recombination of electron-hole pairs [77, 78].

Therefore, the heterojunctions between TiO2 and g-C3N4 and the introduction of HNTs

suppressed the recombination of photoexcited electrons and holes, thus improving the

photocatalytic activity.

Fig. 12. (a) PL spectrum and (b) fluorescence decay curves of TiO2, g-C3N4, CTH-0.5

composite.

3.10 Photocatalytic experiments

The photocatalysitic activities of all the samples were evaluated via the degradation

of ciprofloxacin (CIP) under visible light irradiation. The results of degradation are
exhibited in Fig. 13(a), the pure TiO2 and TiO2/HNTs display no photocatalysitic

activities under visible light irradiation after 60 min. The pure g-C3N4 had removed

45% of CIP, the photocatalysitic activity was enhanced by the formation of

heterojunction with carbon nitride, and 59% of CIP had been degraded until 60min.

Compared with pure TiO2, pure g-C3N4, TiO2/HNTs and g-C3N4/TiO2, the degradation

effect of the heterojunction photocatalysts with different content were significantly

enhanced. In particular, the ability of CTH-0.5 to degrade CIP was the best, which the

removal reaches about 87% after 60 min. These indicated that the introduction of the

g-C3N4 and HNTs surly enhanced the photocatalytic activity of TiO2. For the

photocatalysitic performance, the study of stability is indispensable. Therefore, we

evaluate the stability of CTH-0.5 by photocatalytic degradation of CIP under the same

conditions for four-run recycling experiments. As shown in the Fig. 13(b), the

photocatalytic efficiency of CTH-0.5 showed a slightly reduced, 80% of CIP could be

degraded by CTH-0.5 in the fourth cycle. This indicated that the CTH-0.5 composites

possess a good stability.

Fig. 13. (a) Photocatalytic degradation of CIP over TiO2, g-C3N4, TiO2/HNTs,

TiO2/g-C3N4 and different mass ratio of g-C3N4/TiO2/HNTs composites under visible

light irradiation; (b) Recycling degradation curve for the CTH-0.5 composite.

3.11 MS

To investigate the degradation pathways and the intermediates, HPLC-MS was


performed. The MS spectrums with different time were shown in Fig. 14. The

probable intermediates and degradation pathways were presented in Fig. 15. The CIP

was obtained m/z=318 via the fragmentation of quinolone ring and the loss of the =O

group. After that, the m/z=274 was produced by the leaving of carboxylate group.

After decarboxylation, the loss of -F group was further fragmented and formed

m/z=256[79]. Another pathway was photo-hole (h+) oxidation. The piperazine was

oxidized by photo-hole (h+), which caused it fracturing. The m/z=306 was produced

via loss of Ethyl group. The m/z=263 possessed the similar MS fragmentation patterns,

which was formed by the oxidation of m/z=306. Finally, the above intermediates could

be degraded into CO2, H2O and other molecules.

Fig. 14. m/z of degrading CIP over CTH-0.5. (a) the initial CIP solution; (b) after 30

min; (c) after 60 min).

Fig. 15. The probable intermediates and degradation pathways of photodegradation of

CIP under visible light irradiation.

3.12 Active species trapping and quantification experiments

In the photocatalytic process, in order to confirm the main active species, the

trapping experiment had been always carried out. In this study, 1mM triethanolamine

(TEOA) was added to scavenge the photoinduced hole (h+), 0.1 mM benzoquinone

(BQ) was used to quench the superoxide radical (•O2−), and the 1 mM tert-Butanol
(TBA) was adopted to remove the hydroxyl radical (•OH) [55,80]. As presented in Fig.

16, 87% of CIP could be degraded by the CTH-0.5 composite without adding any

quencher. Along with the addition of TBA, degradation efficiency of CIP showed a

slight decrease, and could be also degraded 85% of CIP, meaning that the •OH was not

the active species in the photocatalytic system. On the contrary, when the TEOA and

BQ were introduced, the degradation efficiency was significantly decreased, which

degraded 30% and 20% of CIP, respectively. The results indicated that the •O2− and the

h+ were the main active species in the photocatalytic process.

Fig. 16. Trapping experiment of active species during the photocatalytic degradation

of CIP over CTH-0.5 composite in the presence of different scavengers under visible

light irradiation.

3.13 Electron paramagnetic resonance spectra

The electron spin resonance (ESR) spin-trap technique with DMPO was generally

applied to detect the existence of •OH and •O2- radicals. As shown in Fig. 17, the six

characteristic signals of DMPO-•O2- and four characteristic peaks signals of

DMPO-•OH both could be detected under visible light illumination,While no signals

could be observed in dark. This indicated that •OH and •O2- are present in

g-C3N4/TiO2/HNTs reaction system. The six characteristic signals of the DMPO-•O2-

could be seen obviously under visible light irradiation that the •O2- can be produced

during light irradiation. In addition, the peaks of DMPO-•OH could be also observed,
but the intensity was weak, because when the light irradiation g-C3N4, the holes on the

VB of g-C3N4 (~1.58 eV) could not directly oxidize OH- (Eθ=1.99 eV) or H2O (Eθ =

2.40 eV) into •OH. The •OH was formed by H2O2. [46] Therefore, it can confirm that

the •O2- was the major active species of CTH-0.5 composite in the photocatalytic

degradation process, consistent with active species trapping experiments results.

Fig. 17. ESR spectra of CTH-0.5 composite in the dark and under visible light

irradiation. (a) DMPO-•O2- ; (b) DMPO-•OH.

3.14 Proposed mechanism in the photodegradation system

Based on these analyses above, a possible mechanism for photodegradation of CIP

using g-C3N4/TiO2/HNTs heterojunction composites under visible light irradiation

was constructed in Fig. 18. The CB and VB of TiO2 and g-C3N4 could be calculated

by an equation [81]:

ECB=X-Ee-0.5Eg

EVB=ECB+Eg

Where ECB was CB edge potential, EVB was VB edge potential, X was

electronegativity of the semiconductor, Ee represented the energy of free electrons in

the hydrogen scale with a constant of 4.5 eV, Eg represented the band gap energy of

the semiconductor. Based on this calculation, the CB and VB value of TiO2 and

g-C3N4 were 2.91 eV and -0.29 eV, 1.58 eV and -1.12 eV, respectively, which were

consistent with UV-DRS and VB XPS test results. When visible light irradiation,
g-C3N4 was excited and produced photogenerated electrons (e-) and holes (h+). Owing

to the CB of g-C3N4 (-1.12 eV) was more negative than TiO2, the e- stand in the CB of

g-C3N4 rapidly transfered to the CB of TiO2 (-0.29 eV) through the heterojunction

between g-C3N4, TiO2 and HNTs, resulting in a large number of e- stand in the CB of

TiO2, which was conducive to suppress the recombination of e- and h+ and to transfer

the e- and h+ quickly. Subsequently the e- were reacted with the O2 to formed •O2- and

a small amount of H2O2. And then pollutant was oxidized by •O2-. The H2O2 were

further reacted with e- to produce •OH. The •OH also oxidized the pollutant.

Simultaneously, the h+ of the VB of g-C3N4 were directly involved in the oxidation of

pollutants. The main reaction could be expressed by the following process [36, 46]:

g-C3N4/TiO2/HNTs + hv → g-C3N4 (h+) + TiO2 (e-)

O2 + e- → •O2-

•O2- + e- + 2H+ → H2O2

H2O2 + e- → 2 •OH

•O2- / •OH / h+ +CIP → CO2 + H2O + other molecules

Fig. 18. Schematic of possible mechanism for photodegradation of CIP over

g-C3N4/TiO2/HNTs heterojunction composites.

4.Conclusion

In summary, g-C3N4-TiO2/HNTs intercalated heterojunction composites were

successfully prepared by sol-gel method and calcination method. Compared with pure
TiO2, the heterojunction composite exhibited higher visible light photocatalytic activity

towards degradation of ciprofloxacin, 87% of CIP could be degraded by g-C3N4-TiO2/

HNTs heterojunction composites, and possessed a good stability. The photoelectric

performance also greatly improved, which may be ascribed to the heterojunction

structure between g-C3N4 and TiO2/HNTs, the introduction of HNTs, resulting in the

rapid transfer and separation of photoelectron-hole pairs. In addition, the main active

species of g-C3N4-TiO2/HNTs heterojunction composites were •O2− and the h+ in the

process of photocatalytic degradation. This work could show the important

application of this photocatalyst in the photocatalytic degradation antibiotics from the

wastewater.

Acknowledgments

We gratefully acknowledge the financial support of the National Natural Science

Foundation of China (grant numbers 21576125, 21407064, 21676127, U1662125,

U1510126); the National Science Foundation of Jiangsu Province (grant numbers

BK20151349,BK20150536); the China Postdoctoral Science Foundation Funded

Project (grant numbers 2015M571683, 2016M590418); and the Postdoctoral Science

Foundation funded Project of Jiangsu Province (1501102B).

References

[1] J.M.A. Blair, M.A. Webber, A.J. Baylay, D.O. Ogbolu, L.J.V. Piddock, Nat. Rev.

Microbiol. 13 (2015) 42.


[2] J. Xu, Y. Xu, H. Wang, C. Guo, H. Qiu, Y. He, Chemosphere 119 (2015) 1379.

[3] H.Q. Wang, J.Z. Li, M.J. Zhou, Q.F. Guan, Z.Y. Lu, P.W. Huo, Y.S. Yan, J. Ind.

Eng. Chem. 30 (2015) 64.

[4] Y. Gao, Y. Li, L. Zhang, H. Huang, J. Hu, S.M. Shah, J. Colloid Interf. Sci. 368

(2012) 540.

[5] F.C. Moreira, S. Garcia-Segura, A.R.B. Rui, E. Brillas, V.J.P. Vilar, Appl. Cat. B:

Environ. 160-161 (2014) 492.

[6] Z.Q. Ma, Y.L. Ma, L. Xie, Z.Y. Zhang, M. Wang, Environ. Sci. Technol. 35 (2012)

46.

[7] C. Zhao, M. Pelaez, X.D. Duan, H.P. Deng, K. O’Shea, D. Fatta-Kassinos, D.D.

Dionysiou, Appl. Cat. B: Environ. 134-135 (2013) 83.

[8] J. Xue, S. Ma, Y. Zhou, Z. Zhang, M. He, Acs Appl. Mater. Interfaces 7 (2015)

9630.

[9] S.Z. Li, X.Y. Li, D.Z. Wang, Sep. Purif. Technol. 34 (2004) 109.

[10] E.S. Aazam, J. Ind. Eng. Chem. 20 (2014) 2870.

[11] M.Y. Chang, C.Y. Chang, Y.H. Hsieh, K.S. Yao, T.C. Cheng, C.T. Ho, Adv. Mater.

Res. 47-50 (2008) 471.

[12] Y. Hou, X. Li, X. Zou, X. Quan, G. Chen, Environ. Sci. Technol. 43 (2009) 858.

[13] Z.B. Xiang, Y. Wang, D. Zhang, P. Ju, J. Ind. Eng. Chem. 40 (2016) 83.

[14] Y.J. Chen, D.D. Dionysiou, Appl. Cat. B: Environ. 80 (2008) 147.

[15] I.Nitoi, P.Oancea, M.Raileanu, M.Crisan, L.Constantin, I.Cristea, J. Ind. Eng.

Chem. 21 (2015) 677.


[16] L. Zhou, L.Z. Wang, J.Y. Lei, Y.D. Liu, J.L. Zhang, Catal. Commun. 89 (2017) 125.

[17] L. Zhou, L.Z. Wang, J.L. Zhang, J.Y. Lei, Y.D. Liu, Res. Chem. Intermed. 43 (2017)

2081.

[18] H. Li, L. Zhou, L.Z. Wang, Y.D. Liu, J.Y. Lei, J.L. Zhang, Phys. Chem. Chem. Phys.

17 (2015) 17406.

[19] V. Subramanian, E. Wolf, P.V. Kamat, J. Phys. Chem. B 105 (2001) 11439.

[20] R.C. Pawar, S. Kang, S.H. Ahn, C.S. Lee, RSC Adv. 5 (2015) 24281.

[21] Y.H. Jiang, F. Li, Y. Liu, Y.Z. Hong, P.P. Liu, L. Ni, J. Ind. Eng. Chem. 41 (2016)

130.

[22] Z.S. Li, S.Y. Yang, J.M. Zhou, D.H. Li, X.F Zhou, C.Y. Ge, Y.P. Fang, Chem.

Eng. J. 241 (2014) 344.

[23] J. Di, J.X. Xia, S. Yin, H. Xu, M.Q. He, H.M. Li, L. Xu, Y.P. Jiang, Rsc Adv. 3

(2013) 19624.

[24] J.X. Sun, Y.P. Yuan, L.G. Qiu, X. Jiang, A.J. Xie, Y.H. Shen, J.F. Zhu, Dalton T.

41 (2012) 6756.

[25] H. Katsumata, Y. Tachi, T. Suzuki, S. Kaneco, Rsc Adv. 4 (2014) 21405.

[26] J.Y. Zhang, Y.S. Wang, J. Jin, J. Zhang, F. Huang, J.G. Yu, Acs Appl. Mat. Interf.

5 (2013) 10317.

[27] X.F. Chen, J. Wei, R.J. Hou, Y. Liang, Z.L. Xie, Y.G. Zhu, X.W. Zhang, H.T.

Wang, Appl. Cat. B: Environ. 188 (2016) 342.

[28] J. Li, M. Zhang, Q.Y. Li, J.J. Yang, Appl. Surf. Sci. 391 (2017) 184.

[29] N.A. Abdelwahab, F.M. Helaly. J. Ind. Eng. Chem, 50 (2017) 162.
[30] D. Jiang, Y. Xu, B. Hou, D. Wu, Y.H. Sun, J. Solid State Chem. 180 (2007) 1787.

[31] C.P. Li, J. Wang, H. Guo, S.J. Ding, J. Colloid Interf. Sci. 458 (2015) 1.

[32] R.A. Rather, S. Singh, B. Pal, J Ind. Eng. Chem. 37(2016) 288.

[33] E. Pajootan, M. Rahimdokht, M. Arami, J Ind. Eng. Chem. 55(2017) 149.

[34] J.H. Zhang, L.L. Zhang, S.Y. Zhou, H. Zhong, Y.J. Zhao, X. Wang, J Ind. Eng.

Chem. 20 (2014) 3884.

[35] R.J. Wang, G.H. Jiang, Y.W. Ding, Y. Wang, X.K. Sun, X.H. Wang, W.X. Chen,

Acs Appl. Mat. Interf. 3 (2011) 4154.

[36] J.Z. Li, M.J. Zhou, Z.F. Ye, H.Q. Wang, C.C. Ma, P.W. Huo, Y.S. Yan, Rsc Adv. 5

(2015) 91177.

[37] P. Zheng, Y. Du, P.R. Chang, X. Ma, Appl. Surf. Sci. 329 (2015) 256.

[38] Z.F. Ye, J.Z. Li, M.J. Zhou, H.Q. Wang, Y. Ma, P.W. Huo, L.B. Yu, Y.S. Yan,

Chem. Eng. J. 304 (2016) 917.

[39] Y. Du, P. Zheng, Korean J. Chem. Eng. 31 (2014) 2051.

[40] S.C. Yan, Z.S. Li, Z.G. Zou, Langmuir 25 (2009) 10397.

[41] F.H. Liu, J. Yu, G.Y. Tu, L. Qu, J.C. Xiao, Y.D. Liu, L.Z. Wang, J.Y. Lei, J.L.

Zhang, Appl. Cat. B: Environ. 201 (2017) 1.

[42] X. Song, Y. Hu, M.M. Zheng, C.H. Wei, Appl. Cat. B: Environ. 182 (2016) 587.

[43] I. Papailias, N. Todorova, T. Giannakopoulou, J.G. Yu, D. Dimotikali, C. Trapalis,

Catal. Today 280 ( 2017) 37.

[44] Z. Tong, D. Yang, T. Xiao, Y. Tian, Z. Jiang, Chem. Eng. J. 260 (2015) 117.

[45] C.Q. Li, Z.M. Sun, Y.L. Xue, G.Y. Yao, S.L. Zheng, Adv. Powder Technol. 27
(2016) 330.

[46] Z. Lu, L. Zeng, W.L. Song, Z.Y. Qin, D.W. Zeng, Ch.S. Xie, Appl. Cat. B:

Environ. 202 (2017) 489.

[47] Z.F. Jiang, K. Qian, C.Z. Zhu, H.L. Sun, W.M. Wan, J.M. Xie, H.M. Li, P.K.

Wong, S.Q. Yuan, Appl. Cat. B: Environ. 210 (2017) 194.

[48] H.G. Yang, G. Liu, S.Z. Qiao, C.H. Sun, Y.G. Jin, S.C. Smith, J. Zhou, H.M.

Cheng, G.Q. Lu, J. Am. Chem. Soc. 131 (2009) 4078.

[49] X. Xu, Z.H. Gao, Z.D. Cui, Y.Q. Liang, Z.Y. Li, S.L. Zhu, X.J. Yang, J.M. Ma,

Acs Appl. Mat. Interf. 8 (2005) 91.

[50] X.Y. Pan, X.X. Chen, Z.G. Yi, Defective, Acs Appl. Mat. Interf. 8 (2016) 10104.

[51] Y.C. Yang, J.W. Wen, J.H. Wei, R. Xiong, J. Shi, C.X. Pan, Acs Appl. Mat. Interf.

5 (2013) 6201.

[52] A. Thomas, A. Fischer, F. Goettmann, M. Antonietti, J.O. Mueller, R. Schloegl, J.

M. Carlsson, Cheminform, 18 (2008) 4893.

[53] Z.F. Jiang, D.L. Jiang, Z.X. Yan, D. Liu, K. Qian, J.M. Xie, Appl. Cat. B:

Environ. 170-171 (2015) 195.

[54] M. Reli, P.W. Huo, M. Šihor, N. Ambrožová, I. Troppová, L. Matějová, J. Lang,

L. Svoboda, P. Kuśtrowski, M. Ritz, P. Praus, K. Kočí, J. Phys. Chem. A 120 (2016)

8564.

[55] C. Xue, T.X. Zhang, S.J. Ding, J.J. Wei, G.D. Yang, Acs Appl. Mat. Interf. 9

(2017) 16091.

[56] P. Sun, G.M. Liu, D. Lv, X. Dong, J.S. Wu, D.J. Wang, Rsc Adv. 5 (2015) 52916
[57] S.Y. Yang, Z.M. Liu, Y.Q. Jiao, Y.P. Liu, C.W. Ji, Y.F. Zhang, J. Mater.

Sci. 49 (2014) 4270.

[58] Q.J. Xiang, J.G. Yu, M. Jaroniec, J. Phys. Chem. C 115 (2011) 7355

[59] J. Li, B. Shen, Z. Hong, B. Lin, B. Gao, Y. Chen, Chem. Commun. 48 (2012)

12017.

[60] H.X. Yu, Y.T. Zhang, X.B. Sun, J.D. Liu, H.Q. Zhang, Chem. Eng. J. 237 (2014)

322.

[61] Y. Zhang, L.J. Fu, H.M. Yang, Colloid Surface A 414 (2012) 115.

[62] D.Z. Lu, P.F. Fang, X.Z. Liu, S.B. Zhai, C.H. Li, X.N. Zhao, J.Q. Ding, R.Y.

Xiong, Appl. Catal. B: Environ. 179 (2015) 558.

[63] Y.G. Li, X.L. Wei, H.J. Li, R.R. Wang, J. Feng, H. Yun, A.N. Zhou, RSC Adv. 5

(2015) 14074.

[64] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso, J.

Rouquerol, K. S.W. Sing, Pure Appl. Chem. 38 (2011) 25.

[65] C.Q. Li, Z.m. Sun, Y.l. Xue, G.Y. Yao, S.L. Zheng, Adv. Powder Technol. 27

(2016) 330.

[66] L.W. Zhang, H.B. Fu, Y.F. Zhu, Adv. Funct. Mat. 18 (2010) 2180.

[67] K.H. Leong, S.L. Liu, C.S. Lan, P. Saravanan, M. Jang, S. Ibrahim, Appl. Surf.

Sci. 358 (2015) 370.

[68] J.Z. Jiang, L. Ou-yang, L.H. Zhu, A.M. Zheng, J. Zou, X.F. Yi, H.Q. Tang,

Carbon 80 (2014) 213.

[69] N. Serpone, D. Lawless, R. Khairutdinov, J. Phys. Chem. 99 (1995) 16646.


[70] K. Li, S.M. Gao, Q.Y. Wang, H. Xu, Z.Y. Wang, B.B. Huang, Y. Dai, J. Lu, Acs

Appl. Mat. Interf. 7 (2015) 9023.

[71] G.Y. Li, X. Nie, J.Y. Chen, Q. Jiang, T.C. An, P. K. Wong, H.M. Zhang, H.J. Zhao,

H. Yamashita, Water Res. 86 (2015) 17.

[72] M. Kong, Y.Z. Li, X. Chen, T.T. Tian, P.F. Fang, F. Zheng, X.J. Zhao, J. Am.

Chem. Soc. 133 (2011) 16414.

[73] Y.M. He, J. Cai, T.T. Li, Y. Wu, Y.M. Yi, M.F. Luo, L.H. Zhao, Ind. Eng. Chem.

Res. 51 (2011) 14729.

[74] G.R. Li, F. Wang, Q.W. Jiang, X.P. Gao, P.W. Shen, Angew. Chem. 49 (2010)

3653.

[75] J.X. Li, J.H. Xu, W.L. Dai, K.N. Fan, J. Phys. Chem. C 113 (2009) 8343.

[76] X.F. Chen, J. Wei, R.J. Hou, Y. Liang, Z.L. Xie, Y.G. Zhu, X.W. Zhang, H.T.

Wang, Appl. Cat. B: Environ. 188 (2016) 342.

[77] W.G. Tu, Y. Zhou, Q. Liu, S.C. Yan, S.S. Bao, X.Y. Wang, M. Xiao, Z.G. Zou,

Adv. Funct. Mater. 23 (2013) 1743.

[78] C.M. Li, Y.H. Du, D.P. Wang, S.M. Yin, W.G. Tu, Z. Chen, M. Kraft, G. Chen, R.

Xu, Adv. Funct. Mater. 27 (2016) 1604328.

[79] Z.Y. Lu, Z. Zhu, D.D. Wang, Z.F. Ma, W.D. Shi, Y.S. Yan, X.X. Zhao, H.J. Dong,

L. Yang, Z.H. Fa, Cat. Sci. Technol. 6 (2016) 1367.

[80] J.Z. Li, Y. Ma, Z.F. Ye, M.J. Zhou, H.Q. Wang, C.C. Ma, D.D. Wang, P.W. Huo,

Y.S. Yan, Appl. Cat. B: Environ. 204 (2017) 224.

[81] M.J. Zhou, D.L. Han, X.L. Liu, C.C. Ma, H.Q. Wang, Y.F. Tang, P.W. Huo, W.D.
Shi, Y.S. Yan, J.H. Yang, Appl. Cat. B: Environ. s 172-173 (2015) 174.
Figure captions
Fig. 1. XRD pattern of as-prepared (a) TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2,
g-C3N4/TiO2 /HNTs composites;(b) Different mass ratio of g-C3N4/TiO2 /HNTs
composites.
Fig. 2. SEM images of (a) TiO2, (b) g-C3N4, (c) HNTs, (d) TiO2/HNTs, (e) g-C3N4/TiO2,
(f) CTH-0.5 and the insets of (f) is the EDS of CTH-0.5 composite.
Fig. 3. TEM images of (a) TiO2, (b) g-C3N4, (c) TiO2/HNTs, (d) g-C3N4/TiO2, (e)
CTH-0.5; HR-TEM image of (f) CTH-0.5 and the insets of (f) is the SEAD of CTH-0.5
composite.
Fig. 4. XPS whole-range spectra of the CTH-0.5 composite.
Fig. 5. XPS spectra of (a) C 1s, (b) N 1s, (c) O 1s, (d) Ti 2p (e) Si 2p and (f) Al 2p of
CTH-0.5 composite.
Fig. 6. FT-IR spectra of TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2, CTH-0.5
composites.
Fig. 7. N2 adsorption-desorption isotherms of TiO2, g-C3N4, CTH-0.5 composite and
the inset is corresponding pore-size distribution curves.
Fig. 8. Raman spectra of TiO2, g-C3N4, CTH-0.5 composite.
Fig. 9. (a) UV-vis diffuse reflectance spectra of TiO2, g-C3N4, CTH-0.5 composite; (b)
Kubelka-Munk function transformed differential spectra of TiO2, g-C3N4, CTH-0.5
composite.
Fig. 10. XPS valence band spectra of the pure TiO2 and g-C3N4.
Fig. 11. (a) Transient photocurrent responses of TiO2, g-C3N4, TiO2/HNTs,
g-C3N4/TiO2, CTH-0.5 composite; (b) EIS spectra of TiO2, g-C3N4, TiO2/HNTs,
g-C3N4/TiO2, CTH-0.5 composite.
Fig. 12. (a) PL spectrum and (b) fluorescence decay curves of TiO2, g-C3N4, CTH-0.5
composite.
Fig. 13. (a) Photocatalytic degradation of CIP over TiO2, g-C3N4, TiO2/HNTs,
TiO2/g-C3N4 and different mass ratio of g-C3N4/TiO2/HNTs composites under visible
light irradiation; (b) Recycling degradation curve for the CTH-0.5 composite.
Fig. 14. m/z of degrading CIP over CTH-0.5. (a) the initial CIP solution; (b) after 30
min; (c) after 60 min).
Fig. 15. The probable intermediates and degradation pathways of photodegradation of
CIP under visible light irradiation.
Fig. 16. Trapping experiment of active species during the photocatalytic degradation
of CIP over CTH-0.5 composite in the presence of different scavengers under visible
light irradiation.
Fig. 17. ESR spectra of CTH-0.5 composite in the dark and under visible light
irradiation. (a) DMPO-•O2- ; (b) DMPO-•OH.
Fig. 18. Schematic of possible mechanism for photodegradation of CIP over
g-C3N4/TiO2/HNTs heterojunction composites.
♦ TiO2 ♣ g-C 3N 4 ♥ HNTs
(a) ♦♥♣
(b) ♦
♥♣
♦ TiO2 ♣ g-C3N4 ♥ HNTs
♦ ♦
g-C3N4-TiO 2/HNTs
♦♦ ♦ ♦ ♦ ♦♦ ♦ CTH-1.5
♦♥
♦ ♦ ♦♦ TiO 2/HNTs

Intensity (a.u.)
♥ ♦

Intensity (a.u.)
♥ CTH-1.0
♥ ♥ ♥ ♥ HNTs

CTH-0.8
g-C 3N4/TiO 2
♦♣ ♦ ♦ ♦♦ ♦

CTH-0.5
♦ ♦ ♦♦ TiO 2
♣ ♦

g-C 3N4 CTH-0.2

10 20 30 40 50 60 70 80 10 20 30 40 50 60 70 80
2 Theta(degree) 2 Theta(degree)

Fig. 1. XRD pattern of as-prepared (a) TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2,
g-C3N4/TiO2 /HNTs composites;(b) Different mass ratio of g-C3N4/TiO2 /HNTs
composites.

Fig. 2. SEM images of (a) TiO2, (b) g-C3N4, (c) HNTs, (d) TiO2/HNTs, (e) g-C3N4/TiO2,
(f) CTH-0.5 and the insets of (f) is the EDS of CTH-0.5 composite.

Fig. 3. TEM images of (a) TiO2, (b) g-C3N4, (c) TiO2/HNTs, (d) g-C3N4/TiO2, (e)
CTH-0.5; HR-TEM image of (f) CTH-0.5 and the insets of (f) is the SEAD of CTH-0.5
composite.

O 1s
Intensity (a.u.)

Ti 2p
Ti 2s

C 1s
N 1s

Al 2s
Si 2p
Si 2s

Al 2p
800 700 600 500 400 300 200 100 0
Binding energy (eV)

Fig. 4. XPS whole-range spectra of the CTH-0.5 composite.

(a) 288.2 eV C 1s (b) N 1s

398.7 eV
Intensity (a.u.)

Intensity (a.u.)

399.5 eV

288.9 eV 401.1 eV
286.4 eV 404.4 eV
284.8 eV

291 290 289 288 287 286 285 284 283 406 404 402 400 398 396
Binding energy (eV) Binding energy (eV)

(c) O 1s (d) Ti 2p
532.1 eV
458.7 eV
Intensity (a.u.)
Intensity (a.u.)

530.0 eV
464.4 eV

538 536 534 532 530 528 526 468 466 464 462 460 458 456
Binding energy (eV) Binding energy (eV)

(e) Si 2p (f) Al 2p
103.1 eV
75.1 eV
Intensity (a.u.)
Intensity (a.u.)

110 108 106 104 102 100 98 96 82 80 78 76 74 72 70 68 66


Binding energy (eV) Binding energy (eV)

Fig. 5. XPS spectra of (a) C 1s, (b) N 1s, (c) O 1s, (d) Ti 2p, (e) Si 2p and (f) Al 2p of
CTH-0.5 composite.
TiO 2

HNTs

Transmittance (a.u.)
TiO 2 /HNTs

g-C 3N 4

g-C 3N 4/TiO 2

CTH-0.5

4000 3500 3000 2500 2000 1500 1000 500


-1
Wavenumber ( cm )

Fig. 6. FT-IR spectra of TiO2, g-C3N4, HNTs, TiO2/HNTs, g-C3N4/TiO2, CTH-0.5


composites.

140
7 TiO2
g-C3N4
6
Absorbed volume (cm /g.STP)

120 HNTs
CTH-0.5
3
Pore volume (cm /g)× 10

4
3

100 3
3

80 1

0
0 10 20 30 40 50
Pore Diameter (nm)
60
TiO2
40 g-C3N4
HNTs
20 CTH-0.5

0
0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (P/P0)

Fig. 7. N2 adsorption-desorption isotherms of TiO2, g-C3N4, HNTs, CTH-0.5


composite and the inset is corresponding pore-size distribution curves.
Intensity (a.u.)

100 120 140 160 180 200


-1
Raman Shift (cm )

g-C 3N 4

CTH-0.5

TiO 2

0 200 400 600 800 1000-1 1200 1400 1600


Raman Shift (cm )

Fig. 8. Raman spectra of TiO2, g-C3N4, CTH-0.5 composite.


1.8 3.5
TiO2
(a) 1.6 TiO2
g-C3N4
(b)3.0 g-C3N4
1.4 HNTs CTH-0.5
g-C3N4/TiO2 2.5
1.2

-2
Absorbance(a.u.)
CTH-0.5

(αhν)1/2 /eV2nm
1.0 2.0

0.8 1.5
0.6
1.0
0.4
0.2 0.5

0.0 0.0
200 300 400 500 600 700 800 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Wavelength(nm) hν (eV)

Fig. 9. (a) UV-vis diffuse reflectance spectra of TiO2, g-C3N4, CTH-0.5 composite; (b)
Kubelka-Munk function transformed differential spectra of TiO2, g-C3N4, CTH-0.5
composite.
Intensity (a.u.)

TiO2

g-C3N4

-2 0 2 4 6 8 10
Binding Energy (eV)

Fig. 10. XPS valence band spectra of the pure TiO2 and g-C3N4.

5 4k
(a) TiO2
g-C3N4
(b)
4
)

TiO2/HNTs
2
Photocurrent µ A/cm

3k
g-C3N4/TiO2
-Z''/ohm

CTH-0.5
3
(

2k
2
TiO2
g-C3N4
1k
1 TiO2/HNTs
g-C3N4/TiO2
CTH-0.5
0 0
20 40 60 80 100 120 140 160 180 200 0.0 2.0k 4.0k 6.0k 8.0k 10.0k
Time (s) Z'/ohm

Fig. 11. (a) Transient photocurrent responses of TiO2, g-C3N4, TiO2/HNTs,


g-C3N4/TiO2, CTH-0.5 composite; (b) EIS spectra of TiO2, g-C3N4, TiO2/HNTs,
g-C3N4/TiO2, CTH-0.5 composite.
12
(a) TiO2 (b) TiO2
10 g-C3N4
g-C3N4
CTH-0.5

PL Intensity 10 CPS
CTH-0.5
8

Intensity (a.u.)
τ1 τ2 τ
6
0.83 0.83 0.83
6 6.17 6.17 6.17
0.63 0.63 0.63
4

0
400 450 500 550 600 650 700 126 128 130 132 134 136 138 140
Wavelength/nm Time (ns)

Fig. 12. (a) PL spectrum and (b) fluorescence decay curves of TiO2, g-C3N4, CTH-0.5
composite.

1.0
(a) 1.0 (b)
0.8
0.8 1st 2nd 3rd 4th

0.6
C/C0

0.6
Ct/C

0.4
0.4 TiO2
TiO2/HNTs
0.2 g-C3N4 g-C3N4-TiO2 0.2
CTH-0.2 CTH-0.5
CTH-0.8 CTH-1.0 CTH-1.5
0.0 0.0
0 10 20 30 40 50 60 0 60 120 180 240
Irridition time (min) Time (min)

Fig. 13. (a) Photocatalytic degradation of CIP over TiO2, g-C3N4, TiO2/HNTs,
TiO2/g-C3N4 and different mass ratio of g-C3N4/TiO2/HNTs composites under visible
light irradiation; (b) Recycling degradation curve for the CTH-0.5 composite.

(a)100 (b) 100 (c)100


332
332
80 80 80 274
Relative Abundance
Relative Abundance

Relative Abundance

60 60 60
318

40 40 40 318
274

256 306 256


20 20 20 332
263 306

0 0 0
200 240 280 320 360 400 200 240 280 320 360 400 200 240 280 320 360 400
m/z m/z m/z

Fig. 14. m/z of degrading CIP over CTH-0.5. (a) the initial CIP solution; (b) after 30
min; (c) after 60 min).
Fig. 15. The probable intermediates and degradation pathways of photodegradation of
CIP under visible light irradiation.

100

80

60
η%

40

20

0
None quencher TBA BQ TEOA
Scavengers

Fig. 16. Trapping experiment of active species during the photocatalytic degradation
of CIP over CTH-0.5 composite in the presence of different scavengers under visible
light irradiation.

-
DMPO - •O2 DMPO - •OH
(a) (b)

♦ ♦

Light on
Intensity (a.u.)
Intensity (a.u.)

♠ ♠
♠ ♠ Light on

Dark Dark

3460 3480 3500 3520 3540 3560 3460 3480 3500 3520 3540 3560
Magnetic Field (G) Magnetic Field (G)

Fig. 17. ESR spectra of CTH-0.5 composite in the dark and under visible light
irradiation. (a) DMPO-•O2- ; (b) DMPO-•OH.
Fig. 18. Schematic of possible mechanism for photodegradation of CIP over
g-C3N4/TiO2/HNTs heterojunction composites.
Table
Table 1 The mass and atomic ratios of CTH-0.5 composite.
Table 2 BET surface area, pore size and pore volume of pure TiO2, pure g-C3N4 and
g-C3N4-TiO2/HNTs samples.

Table 1
The mass and atomic ratios of CTH-0.5 composite.
Weight%
C N O Al Si Ti Au

3.68 2.12 23.99 9.65 10.96 20.23 29.37

Atom%
C N O Al Si Ti Au

9.35 4.62 45.76 10.92 11.91 12.89 4.55

Table 2
BET surface area, pore size and pore volume of pure TiO2, pure g-C3N4 and
g-C3N4-TiO2/HNTs samples

Sample SBET Pore size Pore volume


(m2/g) (nm) (cm3/g)
TiO2 44.862 5.226 0.091
g-C3N4 9.772 23.313 0.083
HNTs 58.45 3.779 0.156
CTH-0.5 39.648 10.432 0.145
Graphical Abstract

You might also like