You are on page 1of 12

Journal of Colloid and Interface Science 527 (2018) 202–213

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Regular Article

Enhanced degradation of ciprofloxacin by graphitized mesoporous


carbon (GMC)-TiO2 nanocomposite: Strong synergy of adsorption-
photocatalysis and antibiotics degradation mechanism
Xuyang Zheng a, Shiping Xu a,⇑, Yao Wang b, Xiang Sun a, Yuan Gao c, Baoyu Gao a
a
Shandong Key Laboratory of Water Pollution Control and Resource Reuse, School of Environmental Science and Engineering, Shandong University, Jinan 250100, China
b
Key Laboratory of the Colloid and Interface Chemistry (Ministry of Education), School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, China
c
Key Laboratory of Industrial Ecology and Environmental Engineering (Ministry of Education), School of Environmental Science and Technology, Dalian University of
Technology, Dalian 116024, China

g r a p h i c a l a b s t r a c t

a r t i c l e i n f o a b s t r a c t

Article history: In order to achieve remarkable synergy between adsorption and photocatalysis for antibiotics elimination
Received 19 April 2018 from water, in this study, a graphitized mesoporous carbon (GMC)-TiO2 nanocomposite was successfully
Revised 15 May 2018 synthesized by an extended resorcinol-formaldehyde (R-F) method. In the composite, the lamellar GMC
Accepted 18 May 2018
nanosheets possessed large specific surface area and mesoporous structure, and could adsorb and enrich
Available online 19 May 2018
antibiotics effectively. This could not only reduce the antibiotic concentration in water shortly, but also
greatly increase the chances for antibiotics to contact with and be degraded by photocatalysts and active
Keywords:
species. Interestingly, GMC could also facilitate the transportation of photogenerated electrons to further
Synergy
Adsorption
improve the photocatalytic efficiency of TiO2, and 15 mg/L ciprofloxacin (CIP) could be totally mineral-
Photocatalysis ized in 1.5 h. Meanwhile, the biological inhibition of reaction solution on luminescence bacteria
Antibiotics decreased obviously with antibiotics degradation until non-toxicity, reinforcing the thorough elimination
Degradation pathway of antibiotics. Besides, from the viewpoint of organic chemistry, several plausible CIP degradation path-
ways were established using HPLC-MS technique, and an interesting intermediate with five-membered
ring structure was firstly proposed, which is helpful to deeply understand CIP degradation. Strong syn-
ergy between adsorption and photocatalysis, along with quick and efficient antibiotics elimination, dou-
ble confirm the great potential of GMC-TiO2 nanocomposite for practical antibiotic wastewater
purification.
Ó 2018 Elsevier Inc. All rights reserved.

⇑ Corresponding author.
E-mail address: shiping.xu@sdu.edu.cn (S. Xu).

https://doi.org/10.1016/j.jcis.2018.05.054
0021-9797/Ó 2018 Elsevier Inc. All rights reserved.
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 203

1. Introduction cm3/g [19]. On this basis, some researchers tried to introduce R-F
reaction into TiO2 synthesis to tailor properties of the products, like
Antibiotic residues in aquatic environment have become one of morphology and pore structures etc. Ding et al. used R-F polymer
the major concerns in environmental protection, and adsorption nanospheres as templates, and successfully synthesize a TiO2-Au-
technology has been proved to be effective to remove antibiotics TiO2 sandwich structured nanocomposite with increased visible
quickly from water [1]. However, as the most commonly employed light harvesting efficiency [20]. Liu et al. prepared hollow TiO2
adsorbent, traditional activated carbon with abundant micropores nanoshells with controllable crystallinity and phase by an R-F
is not ideal for macromolecular antibiotics adsorption due to the resin-protected calcination method, and the product showed sig-
size exclusion effect [2]. Recently, mesoporous carbon with tunable nificantly enhanced photocatalytic activity [21]. Although the pho-
porous structure, strong thermal stability and high surface area [3], tocatalytic performance of TiO2 has been improved, R-F resin only
attracts considerable attention as adsorbents [4], catalyst carriers serves as the template or protection layer, and would be removed
[5], and electrode materials [6] etc. For example, Ji et al. success- in the subsequent calcination process, resulting in no obvious
fully synthesized ordered mesoporous carbons with higher adsorp- adsorption ability reserved in the final product. Inspired by the
tion capacity for tetracycline and tylosin than commercial above elaboration, herein we try to improve the photocatalytic
microporous activated carbons [7]. Unfortunately, physical adsorp- material by retaining R-F during the preparation via employing
tion can only transfer antibiotics from the aqueous solutions to the low-temperature calcination process, where R-F can not only tailor
adsorbents, but cannot degrade the antibiotics completely [8], TiO2 morphology but also produce carbon materials with remark-
moreover, the contaminated adsorbents also possess some security able adsorption capability and fast charge transfer property, this
concerns and need further disposal. Therefore, sole adsorption can- further enhanced the photocatalytic activity of TiO2, and exhibited
not solve the antibiotic residue problems fundamentally. strong synergy between adsorption and photocatalysis. In this
Photocatalysis has been considered as one of the effective and study, a novel graphitized mesoporous carbon (GMC)-TiO2 com-
promising technology for the antibiotics residues elimination from posite was successfully synthesized by using TiCl3 as titanium pre-
water [9]. Of all the photocatalysts, titanium dioxide (TiO2) has cursor via an extended R-F method. TEM observation suggested the
attracted great interests due to its high photocatalytic activity, as prepared GMC-TiO2 composite preserved lamellar structured
chemical stability, non-toxicity, wide application and low cost carbon, where TiO2 nanoparticles were well dispersed. With
[10]. Sturini et al. reported that TiO2 was a valid remediation for Degussa P25 as comparison, ciprofloxacin (CIP), a common fluoro-
various fluoroquinolone antibacterials in surface water and the quinolone antibiotic, was employed as representative to evaluate
degradation rate was two to five times larger than direct photolysis the adsorption ability and photocatalytic activity of the GMC-
[11]. However, the practical application of TiO2 photocatalyst for TiO2 composite, and further explore the antibiotics adsorption
antibiotics removal is hindered by following aspects: (i) high behavior and photocatalytic degradation mechanism. At the same
recombination rate of photogenerated electron and hole pairs sup- time, HPLC-MS technology was used to analyze the molecular
presses the photocatalytic efficiency [12]; (ii) low concentration of structure of intermediates and possible degradation pathways of
antibiotics residues in water environment decreases their contact CIP. Additionally, bioluminescence inhibition experiments using
with photocatalysts [13]; (iii) the agglomeration of TiO2 during luminous bacteria were also conducted to evaluate the toxicity of
the preparation and practical usage restrains its exposure to light CIP and its intermediates during photocatalytic degradation.
irradiation and pollutants [14].
Aiming at the problems mentioned above, some researchers
propose to couple TiO2 with carbon materials, expecting that car-
bon materials with high adsorption capability would quickly 2. Experimental
adsorb and enrich antibiotics followed by mass transfer to TiO2,
and hence increase the chances for the antibiotics and their inter- 2.1. Materials
mediates to contact with TiO2; Meanwhile, carbon materials with
good electrical conductivity would suppress TiO2 agglomeration CIP (C17H18FN3O3, 98%, HPLC grade) with molecular structure
and recombination of charge carriers [15]. Liu et al. successfully showed in Fig. 1 was purchased from Aladdin (Shanghai, China).
synthesized carbon-TiO2 photocatalysts with excellent photocat- Resorcinol, formaldehyde (37 wt%) and TiCl3 (20 wt% TiCl3 dis-
alytic activities for degradation of methyl orange and phenol solved in 3 wt% HCl solution), Triblock copolymer Pluronic F127
[16], while, Hamandi et al. prepared reduced graphene oxide (PEO106PPO70PEO106, EO = ethylene oxide, PO = propylene oxide),
(rGO)/TiO2 nanotube composites with high separation of photo- sodium sulfate (Na2SO4), sodium nitrate (NaNO3) were all of ana-
generated charges [17]. Nevertheless, although photocatalytic lytical grade and purchased from Sinopharm Chemical Reagent
activities of these carbon-TiO2 composites were improved, adsorp- Co., Ltd, China. All the chemicals were used directly without fur-
tion capacities failed to be fulfilled simultaneously. Cheng and co-
workers compared carbon-TiO2-hybrids with different carbon con-
tents, and found that increasing carbon amount would improve the
adsorption capability, whereas, lower carbon content showed bet-
ter photocatalytic performance [18]. To date, there is no win–win
synergy between adsorption and photocatalysis being achieved
and reported in the open literature. It would be very interesting
to develop an ideal photocatalytic material combining high adsorp-
tion ability with excellent photocatalytic activity for contaminants
elimination.
In recent years, various methods based on the cost-effective
resorcinol-formaldehyde (R-F) reaction have been developed to
prepare mesoporous carbon materials [4]. Xu et al. reported a
mesoporous carbon prepared via R-F reaction, and the BET surface
area was up to 2660 m2/g and total pore volume increased to 2.01 Fig. 1. Structure of CIP (m/z = 332).
204 X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213

ther purification. All solutions were prepared using deionized 2.5. Photocatalytic tests
water produced by a Milli-Q system.
Photocatalytic degradation of CIP was carried out using a UV
lamp (14 W, 254 nm, Shanghai JG Special Lighting) as light source
2.2. Synthesis in a cylindrical Pyrex reactor. In each experiment, 70 mg photocat-
alyst was added into 200 mL CIP solution (15 mg/L) under mag-
In a typical synthesis, 1.65 g resorcinol and 2.5 g F127 were dis- netic stirring. Before irradiation, the suspension was maintained
solved in the mixture of 20 mL water and 20 mL ethanol, and stir- in the dark to reach the adsorption–desorption equilibrium. Then
red for 30 min at room temperature. Then, 2.32 mL formaldehyde mixture was exposed to the light. At given time intervals, concen-
was slowly added. After 1 h of stirring, 0.17 mL 37% hydrochloric tration of CIP remaining in solution was measured. In addition, to
acid was added dropwise into the mixed solution. The homoge- evaluate the mineralization degree of CIP during photocatalytic
neous yellowish solution A was obtained after stirring for 15 min. degradation, total organic carbon (TOC) of the solution was deter-
Solution B was prepared as follows: 20 mL TiCl3 was mixed with mined by a TOC analyzer (TOC-V CPH, Shimadzu) as well.
20 mL water, then, 0.7 g Na2SO4 and 0.5 g NaNO3 were added into
the above mixture under vigorous stirring until the solution turned
2.6. Analytical methods
from purple to colorless.
Later, solution A was added into B under continuous stirring to
Concentration of CIP in solution was measured by an UltiMate
form a salmon pink solution. After that, the above mixture was
3000 HPLC system (Thermo Scientific, USA Kyoto, Japan) with a
transferred into a Teflon-lined autoclave and heated at 60 °C for
2.1  150 mm Atlantis C18 column. Mobile phase was composed
2 d. After cooling to room temperature, the white polymeric mono-
of water (A) and methanol (B) containing 0.1% formic acid (v/v)
lith product was collected by centrifuging and washing with deion-
at a A:B ratio of 78:22 with the flow rate of 0.4 mL/min. The detec-
ized water and absolute ethanol thoroughly, and then dried at 80
tion wavelength was set at 275 nm.
°C overnight. Carbonization was then carried out in a tubular fur-
CIP photodegradation intermediates were further identified by
nace under N2 atmosphere at 350 °C for 3 h (2 °C/min), then 800
HPLC-MS technique equipped with a LCQ-Fleet mass spectrometer
°C for 2 h (5 °C/min). Subsequently, in order to improve the crys-
(Thermo Finnigan). Mobile phase composition was same with
tallinity of TiO2 and adjust the carbon content, the obtained sample
above, and the gradient (A:B, v/v) was 80:20 at 0 min, 70:30 at
was calcined at 450 °C for 2 h (5 °C/min) in a muffle oven, and the
2.5 min, 20:80 at 6 min, then returned to initial composition
final product was denoted as GMC-TiO2.
within 5 min and kept unchanged within 4 min. MS was scanned
In comparison, pure GMC and TiO2 were prepared by using the
by mass range from m/z 50 to 550 using ESI in the positive ion
same method except adding the solution B and A, respectively. P25
mode.
(from Degussa) was used as reference material.

2.7. Toxicity tests


2.3. Characterization
Toxicity analysis of intermediates was performed with biolumi-
The X-ray diffraction (XRD) analysis was carried out using a nescent bacteria Vibrio fischeri according to the standard ISO
Bruker D8 Advance X-ray diffractometer with monochromated 11348–3:2007. The inhibition ratio was calculated by the following
high-intensity Cu Ka radiation (k = 1.5418 Å). Raman scattering equation (L means luminescence), and higher inhibition ratio mir-
spectrum was measured on a LabRAM HR800 (HORIBA Jobin Yvon, rors bigger toxicity of the solution [22]. In addition, all toxicity
France) under 633 nm excitation. High resolution transmission experiments were repeated three times, and the average value
electron microscope (HRTEM) image was obtained by JEOL JEM- was adopted.
2100F. Thermogravimetry analysis (TGA) was conducted on an
SDT Q600 (TA Instruments, USA) from room temperature to 800
°C under atmospheric conditions. Brunauer-Emmett-Teller (BET) 3. Results and discussion
surface area and pore size distribution were determined by using
N2 adsorption–desorption isotherms at 77 K on an automated pore 3.1. Characterizations
size and surface area analyzer (JW-BK122W, Beijing JWGB). X-ray
photoelectron spectroscopy (XPS) analysis was carried out using 3.1.1. Crystal structure and morphology
a Thermo ESCALAB 250Xi equipped with standard monochromatic Phase structures of the as-prepared samples were characterized
Al Ka excitation source (hm = 1486.6 eV). Photoluminescence (PL) by XRD analysis. As shown in Fig. 2a, sharp diffraction peaks in
spectra was measured using LS-55 fluorescence spectrometer with GMC-TiO2 can be assigned to anatase TiO2 (JCPDS 21-1272) [23],
an excitation wavelength of 300 nm. The photocurrent measure- and crystallite size of TiO2 is around 12 nm calculated by the
ments were recorded with an electrochemical station (Chenhua Scherrer formula from the full width at half-maximum of (1 0 1)
Instruments, 760E). peak. In addition, compared to pure TiO2, a short but wide bump
around 23° could also be observed in the diffraction pattern of
GMC-TiO2, which holds the same position with one of the two
2.4. Adsorption tests GMC diffraction peaks, along with another weaker peak appeared
at around 43°. These broad and weak peaks at around 23° and
In a typical experiment, 35 mg photocatalyst (GMC-TiO2 or P25) 43° are usually ascribed to the characteristic peaks of graphite
was dispersed in 100 mL of CIP solution with certain initial concen- [24,25], implying that carbon in both GMC-TiO2 and GMC samples
trations. At given time intervals, samples were collected and then were graphitized partly with a certain degree of crystallinity [26].
filtered using a 0.22 lm membrane. The CIP concentration in the XRD analyses suggest that GMC-TiO2 is composed of anatase TiO2
filtrate was then measured with method as described later. In nanoparticles and partially graphitized carbon, which is further
the kinetic study, the initial concentration of CIP was 4–20 mg/L confirmed by Raman analysis and TEM observation.
and temperature was kept at 298 K. Adsorption isotherm tests Raman analysis is widely employed to evaluate the phase and
were performed in the same CIP concentration range with temper- composition of sample. As shown in Fig. 2b, six peaks could be
ature from 298 K to 318 K. clearly observed in the Raman spectrum of GMC-TiO2. Among
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 205

(a) GMC-TiO2 (b) D band


GMC anatase TiO2 G band
TiO2

Intensity (a.u.)

Intensity (a.u.)
10 20 30 40 50 60 70 80 500 1000 1500 2000

Fig. 2. XRD patterns of GMC-TiO2, GMC and TiO2 (a); Raman spectra (b); TEM images of GMC-TiO2 under different magnifications (c).

them, four peaks below 1000 cm1 is indexed to anatase TiO2 [27], phases is strong, which could stabilize the anatase polymorph in
and the other two peaks at 1328.10 cm1, 1598.47 cm1 could be GMC-TiO2 even at high temperature. Thus, it is reasonable to state
assigned to the disordered SP3 carbon atoms (D band) and graphi- that, the as-prepared GMC-TiO2 is composed of well-dispersed
tic SP2 hybrid carbon atoms (G band) [28,29], respectively. Inter- TiO2 nanoparticles and lamellar graphitized carbon nanosheets,
estingly, peak of G band is sharp and symmetric, which confirms both of which are firmly combined together.
the existence of graphite in the sample [30,31]. It is generally Generally, the resorcinol-formaldehyde (R-F) reaction is often
accepted that graphite is a good conductor of electricity, thus it used to prepare carbon aerogel or xerogel [36]. In this study, by
is supposed that the graphitized carbon in GMC-TiO2 composite introducing titanium precursor into the R-F mixture, the obtained
is beneficial for the photo-excited electrons transfer and further sample exhibits lamellar structure containing partially graphitized
photocatalytic activity of the composite. carbon, as deduced from XRD, Raman and TEM analyses. This inter-
Morphology of the as prepared GMC-TiO2 composite was esting lamellar graphitic carbon structure, on one hand, can pro-
observed via TEM and HRTEM. As clearly shown in Fig. 2c, some mote the fast transfer and utilization of photo-generated
nanoparticles with average size about 15 nm were dispersed electrons, on the other hand, would provide TiO2 with more and
homogeneously on the lamellar nanosheets, and lattice fringes of stronger binding sites to have a better dispersion, both of which
the nanoparticles could be clearly observed with lattice spacing signify the good pollution remediation capability of GMC-TiO2
of 0.35 nm ascribing to planes of anatase TiO2 (JCPDS 21-1272). sample, mirroring great potential of the extended R-F approach.
This agrees well with the XRD results. Lattice fringe of the In addition, for TiO2 preparation, several titanium precursors
nanosheet structure is not well-defined, but combing with the (e.g. isopropyl titanate, tetrabutyl titanate, titanium tetrachloride
XRD and Raman results, it is reasonable to infer that the lamellar etc.) are usually employed. However, for the R-F aqueous solution
structure corresponds to the partially graphitized carbon. system, the common titanium precursors are highly reactive and
It is generally known that nanoscale TiO2 is inclined to aggre- sensitive to moisture, resulting in uncontrollable morphology and
gate together during the synthesis due to the huge surface energy properties of the obtained TiO2 [37,38]. While aqueous TiCl3 in
[32], just as shown in the TEM image of pure TiO2 (Fig. S1). Inter- HCl solution could keep stable after being introduced into R-F solu-
estingly, for the GMC-TiO2 composite obtained in this study, TiO2 tion, and more importantly, the synchronously added HCl could
nanoparticles are well dispersed on the carbon matrix, which serve as catalyst to accelerate the R-F reaction [4]. Nevertheless,
mainly benefits from the lamellar nanosheet structure of GMC TiCl3 precursor usually produces rutile phase TiO2 with relatively
and firm combination between the two phases, where lamellar low photocatalytic activity, since the octahedral [TiO6] units prefer
structure of GMC can provide more anchor points for TiO2 to grow along [0 0 1] direction under acid condition to lower the
nanoparticle; meanwhile, strong interaction between the two combination barrier [39]. In this study, by employing Na2SO4 as
phases can further impede the sintering of TiO2, and inhibit the crystal regulator, SO24 ion would alter the manner and location
agglomeration of TiO2 nanoparticles [33,34]. of [TiO6] units combining with the original nucleus through chelat-
Strong interaction between GMC and TiO2 is also evidenced by ing effect, resulting in the change of [TiO6] growth direction from
the sole and pure anatase phase of TiO2 in GMC-TiO2 sample. Usu- initial single direction to oblique growth, and crystal phase from
ally, for pure TiO2, the metastable anatase polymorph would trans- rutile to anatase [40,41].
form to rutile phase at temperature above 500 °C [35]. In this
study, GMC-TiO2 sample was carbonized at 800 °C, but only ana- 3.1.2. TGA and pore structure analyses
tase phase is observed in XRD and HRTEM analyses, indicating that TGA was used to measure the content of carbon in the as-
TiO2 phase is well dispersed on the carbon matrix [33], as clearly prepared GMC-TiO2 composite. As shown in Fig. 3a, a substantial
observed in TEM images; Moreover, the interaction between both weight loss from 400 to 600 °C is clearly observed owing to the
206 X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213

100 Fig. 4a, peaks of C, O and Ti could be obviously found in the survey
(a) spectrum, revealing that the sample consists of C, O and Ti
elements.
In the high resolution XPS spectrum of Ti 2p (Fig. 4b), two
Weight loss (wt %)

strong peaks located at 459.8 and 465.5 eV could be ascribed to


90 Ti 2p3/2 and Ti 2p1/2 of Ti4+; meanwhile, the peaks of Ti3+ at around
457 and 461 eV could not be found [27], proving that Ti3+ was
totally oxidized to Ti4+ during the synthesis process. In addition,
two shoulder peaks found at binding energy 461.6 and 467.8 eV
are interrelated to surface structural defect [45], which could effec-
80 tively capture the photogenerated charges and prevent charge
recombination by forming an additional localized state in the band
gap of TiO2 [46], and further improve the activity of the photocat-
0 200 400 600 800 alyst. Interestingly, the peaks of Ti in GMC-TiO2 shift to a higher
Temperature (ºC) binding energy compared to pure TiO2. Binding energy is related
300 to the chemical and electrostatic environment of the neighbouring
0.04
(b) atom, therefore, this shift suggests the decreased electron cloud
dV/dD (cm /nm g)

0.03
250 density, and further indicates the transmission of electron. For
3

0.02
Volume of N2 (cm /g)

GMC-TiO2, electrons in TiO2 are most likely to be transferred to


3

0.01
200 the closely connected carbon, which could be further verified by
0.00
1 10 100 the following electrochemical analysis. Efficient charge transfer is
Pore Width (nm)
150 beneficial to prevent the photo-induced electron-hole recombina-
tion, and further enhance the activity of the product [47].
100
Besides, the characteristic peaks in the O 1s spectrum of GMC-
TiO2 (Fig. 4c), located at 531.0 and 532.7 eV, could be assigned to
lattice oxygen bound to Ti4+ (TiAOATi) and surface hydroxyl
50
groups (OAH bond) [48], respectively, which further evidences
the formation of TiO2 in GMC-TiO2 sample.
0
0.0 0.2 0.4 0.6 0.8 1.0 The high resolution spectrum of C 1s shows three peaks with
P/P0 the binding energy at 284.8, 286.0, and 289.5 eV, respectively.
Since the as-prepared GMC-TiO2 contains carbon component, the
Fig. 3. TGA analysis (a) and N2 adsorption/desorption isotherm of GMC-TiO2 (inset: strong peak at 284.8 eV could be assigned to SP2 and SP3 hybrid
pore size distribution curve) (b). CAC bond of the carbon matrix [49], along with the inevitable
adventitious elemental carbon. Other two peaks at 286.0 and
289.5 eV suggest the existence of CAO and C@O bonds, respec-
burning and elimination of carbon. Thus, carbon content of GMC-
tively, indicating the possible formation of CAOATi between
TiO2 composite is calculated to be around 14.5%, which is higher
GMC and TiO2 [24,28,48], which favors the charge transfer upon
than most of the carbon doped TiO2 materials (< 5%) [6,42,43].
light excitation [50]. In addition, characteristic peak of TiAC bond
Generally, carbon materials possess rich pore structures, which
located at around 282 eV does not appear, showing that carbon
is the premise for excellent adsorption capability. In order to inves-
did not replace the lattice oxygen of TiO2 [28].
tigate the pore structure of the abundant carbon in GMC-TiO2 com-
posite, N2 adsorption–desorption behavior was recorded, and the
3.1.4. PL and photocurrent response analyses
adsorption–desorption isotherm and pore size distribution plot of
PL analysis is frequently employed to reveal the separation,
GMC-TiO2 composite are labeled in Fig. 3b. It can be seen that
migration and recombination of photo-generated electrons and
the isotherm exhibits a typical type IV curve, revealing the meso-
holes upon light irradiation, and weaker PL intensity usually indi-
porous structure of the sample, meanwhile, pore size distribution
cates lower recombination rate of electron-hole pairs and better
curve (inset of Fig. 3b) also confirms the presence of mesopores.
photocatalytic performance [51]. Fig. 5a shows the PL spectra of
In addition, N2 adsorption–desorption isotherm represents a type
P25 and GMC-TiO2 composite. It can be found that, compared to
H3 IUPAC (International Union of Pure and Applied Chemistry)
P25, emission peak intensity of GMC-TiO2 is much lower, mirroring
hysteresis loop, indicating that the mesopores mainly root from
the less recombination of electrons-holes pairs in GMC-TiO2 sam-
the slit-shaped pores between the graphite layer [44]. Generally,
ple. Efficient migration of photo-induced carriers in GMC-TiO2
traditional activated carbon with micropores is not ideal for the
could be attributed to the existence of carbon component, espe-
adsorption of large antibiotics molecules due to the size-
cially the graphitized carbon, which can quickly accept and trans-
exclusion effect [2], while the existence of abundant mesopores
fer electrons coming from conduction band (CB) of TiO2 via Ti-O-C
in GMC-TiO2 composite could effectively promote the enrichment
bond [47,52,53] and, thereby, efficiently suppress the recombina-
of CIP into sample. Additionally, BET specific surface area of
tion of photo-generated electrons and holes. This result is consis-
GMC-TiO2 composite is 286.0 m2/g, which is nearly six times larger
tent with the XPS analysis.
than that of commercial P25 owing to the large amount of carbon
In order to further investigate the charge separation and trans-
presenting in GMC-TiO2 sample according to TGA analysis. Both big
fer properties, the transient photocurrent measurements of GMC-
specific surface area and mesoporous structure could enhance the
TiO2 composite and P25 were conducted under UV irradiation. As
adsorption capacity of antibiotics, and further improve photocat-
displayed in Fig. 5b, a fast and stable photocurrent response was
alytic performance of the GMC-TiO2 sample.
obtained in each light on–off cycle for GMC-TiO2 sample, and its
current densities are three times higher than that of P25, indicating
3.1.3. Surface chemistry analysis the more efficient separation of photo-generated electrons and
XPS analysis was carried out to explore the surface chemical holes in GMC-TiO2. Both PL analysis and transient photocurrent
composition and electronic structure of GMC-TiO2. As shown in response fully indicate that synergy between the well-dispersed
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 207

(a) C 1s (b) 459.8 Ti 2p


O 1s

GMC-TiO2 Ti 3p
465.5
Intensity (a.u.)

Intensity (a.u.)
467.8 461.6
Ti 3p GMC-TiO2

TiO2

TiO2

1400 1200 1000 800 600 400 200 0 470 465 460 455 450
Binding energy (eV) Binding energy (eV)

(c) 531.0 O 1s C1s


(d) 284.8

532.7
Intensity (a.u.)

Intensity (a.u.)
GMC-TiO2

286.0
289.5
TiO2

540 536 532 528 297 294 291 288 285 282
Binding energy (eV) Bingding energy (eV)

Fig. 4. XPS survey spectrum (a) and high-resolution spectra of Ti 2p (b), O 1s (c) of GMC-TiO2 and TiO2, C 1s (d) of GMC-TiO2.

TiO2 nanoparticles and graphitized carbon in GMC-TiO2 sample can q2e k2 t


qt ¼ ð2Þ
significantly suppress the carriers’ recombination, enhance the 1 þ qe k2 t
effective charge transportation, and then further promote photo-
where qt (mg g1) and qe (mg g1) represent the adsorption
catalytic capability.
capacity of CIP on GMC-TiO2 composite at time t (min) and equilib-
rium, respectively; k1 (min1) and k2 (g mg1 min1) are the rate
3.2. Adsorption capability of GMC-TiO2
constants of pseudo-first-order and pseudo-second-order model,
respectively.
Concentrations of antibiotic residues in water environment are
The fitting curves of kinetic data using pseudo-first-order and
extremely low, which limits the contact with photocatalysts and
pseudo-second-order model are shown as Fig. 6(a), and the calcu-
further depresses the photocatalytic degradation, therefore, effi-
lated parameters are presented in Table 1. Compared to pseudo-
cient enrichment of antibiotics onto photocatalyst is beneficial to
first-order model, correlation coefficients of pseudo-second-order
promote the fast removal of antibiotic residues. Fig. 6(a) represents
model are much higher (R2 > 0.99) and the calculated qe values
CIP adsorptions on GMC-TiO2 sample with different initial CIP con-
match better with the experimental data, all of which indicate that
centrations and contact time. It could be clearly seen that the
the pseudo-second-order model is more appropriate to describe
adsorption rate was fast in the first 10 min with adsorption capac-
the adsorption behavior of CIP on GMC-TiO2 composite. This agrees
ity increasing sharply, and then decelerated gradually until reach-
well with the previous studies about CIP adsorption [26,55], and
ing the adsorption equilibrium. Meanwhile, with the increase of
implies adsorption process is controlled by chemisorption involv-
initial CIP concentration, equilibrium adsorption capacity
ing valence forces through exchanging or sharing electrons
increased and the equilibrium time extended from 20 to 90 min
between CIP and GMC-TiO2 [56].
due to the adsorption active sites competition [54]. In order to fur-
ther explore and analyze the CIP adsorption behavior and mecha-
nism onto GMC-TiO2 sample, adsorption kinetics and isotherm 3.2.2. Adsorption isotherm study
were studied. In this study, the adsorption isotherms of CIP onto GMC-TiO2
composite were analyzed by fitting the experimental data with
3.2.1. Adsorption kinetics study Langmuir and Freundlich isotherm models, respectively. The Lang-
The nonlinear curve-fittings of the pseudo-first-order model muir isotherm model (Eq. (3)) presumes monolayer adsorption
(Eq. (1)) and pseudo-second-order model (Eq. (2)) were used to occurring on a homogeneous surface with equivalent adsorption
analyze the kinetic adsorption behavior of GMC-TiO2 composite. sites; whereas the Freundlich model (Eq. (4)) assumes multilayer
qt ¼ qe ð1  ek1 t Þ ð1Þ adsorption over a heterogeneous surface with non-uniform
adsorption sites.
208 X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213

1400
GMC-TiO2 (a) 25 (a)
1200 P25
20
1000
Intensity (a.u.)

15

qt(mg/g)
800

600 100
10

400 50 5 4 mg/L 15 mg/L


6 mg/L 20 mg/L
350 400 450 500 550
200 0
8 mg/L Pseudo-first-order
10 mg/L Pseudo-second-order
0
0 20 40 60 80 100 120
350 400 450 500 550
Wavelength (nm)
Time (min)

GMC-TiO2 (b) 28 (b)


P25 26
24
Relative intensity (a.u.)

22

qe (mg/g)
20
18 298 K
16 308 K
318 K
14
Langmuir
12 Freundlich
10
50 100 150 200 250 300 2 4 6 8 10 12 14 16 18 20 22
Time (s) Ce (mg/L)
Fig. 5. PL spectra of GMC-TiO2 and P25 (inset: the larger version of GMC-TiO2) (a);
transient photocurrent-time curves of GMC-TiO2 and P25 (b). Fig. 6. Adsorption kinetics (a) and isotherms (b) of CIP adsorption on GMC-TiO2.

qm K L C e where T (K) is the reaction adsorption temperature, R (8.314 J


qe ¼ ð3Þ
1 þ K LCe mol1 K1) is the ideal gas constant, and K (L mol1) is the Lang-
muir constant.
qe ¼ K F C e1=n ð4Þ The maximum adsorption capacity of CIP over GMC-TiO2 com-
1 posite decreases as temperature rising, suggesting lower tempera-
where qm (mg g ) is the saturated absorption capacity; Ce (mg
ture is more conducive to the CIP removal. This could be attributed
L1) is the CIP concentration at equilibrium; KL (L mg1) and KF
to the exothermic nature of the adsorption, evidenced by the neg-
((mg g1(1/mg1/n)1) are the Langmuir and Freundlich constants,
ative DH0 value as listed in Table 3. Moreover, the calculated DG0
respectively; n is the adsorption intensity indicator in the Fre-
values are negative under all the three tested temperature, indicat-
undlich model.
ing that the adsorption could occur spontaneously [55].
The nonlinear fitting curves for adsorption isotherms at differ-
ent temperatures are shown as Fig. 6(b), with isotherm constants
and correlation coefficients listed in Table 2. Compared to Fre- 3.3. Photocatalytic activity of GMC-TiO2
undlich model, Langmuir model with higher correlation coeffi-
cients (R2 > 0.9) is more suitable to represent the isotherms, Based on the above results of adsorption, the photocatalytic
suggesting that the adsorption of CIP on GMC-TiO2 composite is performance of GMC-TiO2 composite in degrading CIP under UV
monolayer adsorption and takes place on homogeneous adsorption irradiation was studied. The results as shown in Fig. 7(a) demon-
sites. Moreover, it is worthy to note that, benefited from the large strate that, during the dark period, GMC-TiO2 composite can
specific surface area and mesoporous structure of GMC-TiO2 com- remove about 50% CIP from water, by contrast, P25 showed almost
posite, the maximum saturated adsorption capacity of CIP calcu- no CIP removal. This great difference resulted from the significant
lated through the Langmuir model is more than 27 mg g1. Fast disparity of their adsorption capabilities as discussed above. In
adsorption rate and large adsorption capability ensure the poten- addition, CIP concentration presented a negligible change in the
tial of GMC-TiO2 for the rapid and effective removal of CIP from absence of photocatalyst, indicating CIP itself is stable and barely
water. degraded under UV illumination.
In addition, thermodynamic parameters including Gibbs free Under UV irradiation, especially in the initial 20 min, both
energy (DG0), enthalpy (DH0) and entropy (DS0) all can be calcu- GMC-TiO2 and P25 can degrade CIP rapidly with CIP concentration
lated by using the following equations: drop sharply, and the degradation rate constants (k) of both sam-
ples, calculated by fitting with the first-order kinetic equation,
DG0 ¼ RTlnK ð5Þ
were 0.102 and 0.107, respectively. The almost equal values indi-
cate the good photocatalytic activity of GMC-TiO2 as high as P25.
DG0 ¼ DH0  T DS0 ð6Þ It is worthy to note that, as calculated from TGA data, TiO2 content
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 209

Table 1
Kinetic parameters for the CIP adsorption onto GMC-TiO2.

C0 (mg L1) Pesudo-first-order Pesudo-second-order qe (exp.) (mg g1)


1 1 2 1 1 1 2
k1 (min ) qe (cal.) (mg g ) R k2 (g mg min ) qe (cal.) (mg g ) R
4.00 0.039 4.27 0.921 0.128 11.26 1.000 11.19
6.00 0.036 7.94 0.975 0.049 15.63 1.000 15.42
8.00 0.036 16.06 0.905 0.022 19.92 0.999 19.64
10.00 0.053 33.72 0.831 0.019 23.31 0.999 22.89
15.00 0.050 26.44 0.941 0.020 25.32 0.999 24.82
20.00 0.045 29.14 0.910 0.015 27.40 0.999 26.82

Table 2
Isotherm parameters for the CIP adsorption onto GMC-TiO2.

T (K) Langmuir Freundlich


qm (mg g1) KL (L mg1) R2 KF n R2
298 27.70 2.23 0.9991 17.31 4.535 0.9259
308 26.53 1.35 0.9954 15.10 4.403 0.9256
318 25.91 1.27 0.9824 15.63 5.682 0.9805

Table 3 in GMC-TiO2 composite is much lower than that of P25. Combining


Thermodynamic parameters for the CIP adsorption onto GMC-TiO2. with the characterization results, it is reasonable to infer that good
T (k) DG0 (kJ mol1) DH0 (kJ mol1) DS0 (kJ mol1 K1) photocatalytic activity of GMC-TiO2 with lower TiO2 content
298 33.48 22.00 0.0379 mainly benefits from its excellent adsorption capability and charge
308 33.32 transfer property: efficient adsorption can provide a continuous
318 34.24 CIP supply to photocatalyst surface, resulting in an obvious syner-
gistic effect between adsorption and photocatalysis; moreover,
good charge transfer property induced by the carbon component,
especially the graphitic carbon, can greatly suppress the recombi-
nation of photo-generated electrons and holes, both of which are
16 (a) Dark UV light irradiation beneficial to promote the photocatalytic activity of product.
Besides, the big specific surface area and good dispersion of TiO2
CIP concentration (mg/L)

12 particles in GMC-TiO2 composite also favor the photocatalytic


activity.
GMC-TiO2
It is worthy to note that, for GMC-TiO2, CIP concentration in
P25
8 solution was always lower than that of P25 during the whole pho-
Only irradiation
tocatalytic process, and the total time required to completely
remove CIP was shorter as well. Moreover, according to TOC results
4 shown in Fig. 7(b), the removed CIP was completely mineralized
into inorganic products during the photodegradation process. All
of these are very interesting and meaningful to achieve the pur-
0
pose of water purification in a fast and effective manner.

-90 0 50 100 3.4. Pathways of CIP photocatalytic degradation


Time (min)
Degradation of antibiotics is very complex, which involves
(b) many intermediates with bigger or smaller environmental impacts
1.0 compared to the original pollutant. In this study, in order to iden-
GMC-TiO2
tify the main intermediates and understand the possible pathways
P25
0.8 of CIP degradation, MS and HPLC-MS analyses were employed,
with profiles of the identified intermediates shown in Fig. 8. It
can be found that, at least six intermediates (denoted as A(A0 ) to
TOC/TOC0

0.6 E, listed in Table 4) were generated during the CIP photodegrada-


tion, with the protonated molecular ions at m/z = 348, 318, 274,
0.4 306, 279 (Fig. S2 and S3), respectively.
For GMC-TiO2 system, combining the profile of CIP (Fig. 7(a))
with its intermediates (Fig. 8a), it could be inferred that A should
0.2
be the first intermediate generated during CIP degradation, and
the other four intermediates (B-E) then gradually appeared along
0.0 with A degradation. However, although the four intermediates B
0 20 40 60 80 100 120 to E exhibited similar variation trend: firstly arise to the peak
Time (min) and then drop, time consumed to reach the peak value distin-
Fig. 7. Concentration profile of CIP during the dark period and UV irradiation period
guished intermediates B, C from D, E, indicating that products B,
(a); the change of solution TOC during CIP degradation by GMC-TiO2 and P25, C and D, E might stem from different degradation pathways. More-
respectively (b). over, comparing with D, E, variation of B, C proceeded much faster
210 X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213

Table 4
0.09 (a) A (m/z = 348) Intermediate products determined by MS analysis.
0.08 B (m/z = 318)
C (m/z = 274) Intermediate Molecular [M + H]+ Structural formula
0.07 products formula m/z
Relative peak area (%)

D (m/z = 306)
0.06 E (m/z = 279) CIP C17H18FN3O3 332

0.05
0.04
0.03 A C17H18FN3O4 348
0.02
0.01
0.00 A0 C17H18FN3O4 348
0 20 40 60 80 100 120
Time (min)

B C16H16FN3O3 318
0.18 (b) A (m/z = 348)
0.15 B (m/z = 318)
C (m/z = 274)
0.12
Relative peak area (%)

D (m/z = 306) C C15H16FN3O 274


0.09 E (m/z = 279)

0.04
0.03 D C15H16FN3O3 306

0.02
0.01
E C13H11FN2O4 279
0.00
0 20 40 60 80 100 120
Time (min)

Fig.8. Relative intensity variations of intermediates with GMC-TiO2 (a) and P25 (b)
as photocatalyst, respectively.
In the proposed structure of intermediate A, the piperazine ring
containing hydroxyl group formed a hemiacetal structure, which is
and in a much broader range, suggesting that degradation pathway unstable and would open loop easily, resulting in the formation of
of B and C should be preferential for CIP degradation over GMC- an aldehyde group in C12 (P1). The formed aldehyde group could
TiO2. quickly lose a hydrogen atom and further form an acyl radical, then,
For P25 system (Fig. 8b), the intermediate A was also generated oxidation occurred by losing one electron to form an oxonium ion
at first just like GMC-TiO2 system. However, for intermediates B-E, (P2). Subsequently, an intramolecular SN2 (bimoleculer nuceleophi-
it could be found that, no matter the variation ranges or the time lic substitution) reaction took place, followed by releasing one mole-
required to reach peak was quite different from that of GMC-TiO2 cule of CO. At last, C13 bonded with N11 and formed a five-
system; especially, amount of intermediate E produced was excep- membered ring, namely the formation of intermediate B. It is worthy
tionally large. All of these indicated that the preferential CIP degra- to note that, formation of an intermediate with a five-membered
dation pathways of both systems were definitely different. ring structure in piperazine ring moiety during CIP degradation
Based on the above results, the plausible CIP degradation path- was first proposed, and this new structure is very meaningful for a
ways involved in this study were proposed and illustrated in Fig. 9. deeper understanding of CIP degradation in aqueous solution.
Path I: CIP ? A (m/z = 348) ? B (m/z = 318) ? C (m/z = 274). Comparing with product B, the molecular weight of intermedi-
The specific transition process could be described as follows: ate C decreased by 44, just corresponding to one molecule of CO2.
molecular weight of intermediate A increased by 16 comparing Since decarboxylation is pretty common in MS, the intermediate C
with CIP, suggesting the addition of an oxygen atom or the substi- should be directly formed via decarboxylation of product B. Similar
tution of a hydrogen atom by a hydroxyl group. Since hydroxyla- degradation process was also reported by Deng et al. [57], Jiang
tion is commonly reported in CIP photodegradation [57], here, et al. [60] etc.
intermediate A was more likely formed by hydroxyl group substi- Path II: CIP ? A (m/z = 348) ? D (m/z = 306) ? E (m/z = 279). In
tution. In the open literature, there were many possible positions this degradation process, the cleavage of the piperazine ring
for hydroxyl substitution, for example, C5 or N14 in the piperazine mainly took place. After forming product A like in Path I, C13 in
ring as marked in Fig. 1 [58,59]. However, from the viewpoint of piperazine ring of A could also be attacked by a hydroxyl radical
organic chemistry, the most possible position should be C12 in to obtain two molecules of hemiacetal (P3). Intermediate D was
piperazine ring, since N11 in the piperazine ring was a tertiary then generated via hydrolysis of hemiacetal, which is ready to hap-
amine with strong electron-withdrawing property, whose a- pen in the solution. After that, the piperazine ring moiety of inter-
carbon (C12) was unstable and susceptible to be attacked by active mediate D was fully dealkylated, remaining only amino group;
species, like hydroxyl radical in photocatalytic system. Thus, it is meanwhile, C8 of CIP was attacked by hydroxyl radical, resulting
speculated that intermediate A was formed through hydroxyl sub- in hydroxyl substitution. After these two reactions, intermediate
stitution with C12 being attacked by a hydroxyl radical. E was formed.
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 211

Fig. 9. The main degradation pathways of CIP over GMC-TiO2 and P25.

Considering the profiles of all the identified intermediates, Path theless, since amount of intermediate E increased dramatically at
I might be the preferential pathway for CIP degradation over GMC- the very beginning of CIP degradation over P25, it is supposed that,
TiO2; whereas, Path II might be preferential for P25 system. Never- there is still another important possible pathway for P25 system
212 X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213

adsorb and degrade CIP rapidly, and finally achieved the complete
GMC-TiO2
80 mineralization. In addition, HPLC-MS analysis showed that the
Luminescence inhibition (%)

P25 degradation of CIP mainly followed three plausible pathways,


involved hydroxylation reaction, cleavage of piperazine ring and
60 decarboxylation. Toxicity study suggested that biological inhibi-
tion of intermediates on luminescent bacteria Vibrio fischeri
decreased significantly along with CIP degradation, which is ideal
40 for the treatment of complex antibiotic wastewater.

20 Acknowledgements

This work was supported by National Natural Science Founda-


0 tion of China (21607095), the Research Foundation for the Out-
0 20 40 60 80 100 120 standing Young and Middle-aged Scientists of Shandong Province
Time (min) of China (BS2014HZ005), the Taishan Scholar Program
(ts201511003), and Postdoctoral Innovation Project of Shandong
Fig. 10. The luminescence inhibition changes of CIP solution degraded by GMC-
Province of China (201302028).
TiO2 and P25, respectively.

(Path III): CIP ? A0 (m/z = 348) ? E (m/z = 279). In brief, C8 in the Appendix A. Supplementary material
quinolone moiety of CIP was first hydroxylated to form A0 . Subse-
quently, hydroxylation and hydrolysis of C12 and C16 happened Supplementary data associated with this article can be found, in
in succession to form intermediate E. the online version, at https://doi.org/10.1016/j.jcis.2018.05.054.
It is worth noting that, for GMC-TiO2 system, after 120 min’
irradiation, nearly all the intermediates were degraded, and their
References
amounts were obviously lower than that of P25 system; moreover,
during the whole degradation process, amount changes of all the [1] M.B. Ahmed, J.L. Zhou, H.H. Ngo, W. Guo, Adsorptive removal of antibiotics
intermediates were gentle and no intermediate with abnormally from water and wastewater: progress and challenges, Sci. Total Environ. 532
(2015) 112–126.
large amount just like intermediate E in P25 system were gener-
[2] H. Liu, J. Zhang, H.H. Ngo, W.S. Guo, H.M. Wu, Z.Z. Guo, C. Cheng, C.L. Zhang,
ated. In addition, the bioluminescence inhibition experiments with Effect on physical and chemical characteristics of activated carbon on
results shown in Fig. 10 indicated that, for GMC-TiO2 system, tox- adsorption of trimethoprim: mechanisms study, RSC Adv. 5 (2015) 85187–
icity of the reaction solution was reduced significantly along with 85195.
[3] W. Gao, W. Li, Z. Xue, M. Pal, Y. Liu, C. Wang, J. Wang, S. Wang, X. Wan, Y. Liu, D.
CIP degradation, and luminescence inhibition of the solution was Zhao, Preparation of mesoporous TiO2–C composites as an advanced Ni
much lower than that of P25 during the whole degradation pro- catalyst support for reduction of 4-nitrophenol, New J. Chem. 40 (2016) 4200–
cess. This is in consistent with the profiles of CIP and its interme- 4205.
[4] L. Liu, Q.F. Deng, T.Y. Ma, X.Z. Lin, X.X. Hou, Y.P. Liu, Z.Y. Yuan, Ordered
diates in both systems. All of these evidence that, GMC-TiO2 mesoporous carbons: citric acid-catalyzed synthesis, nitrogen doping and CO2
could greatly reduce the secondary pollution risk caused by the capture, J. Mater. Chem. 21 (2011) 16001–16009.
intermediate products during pollutant elimination, which is of [5] Z. Zhao, X. Wang, Supported phosphotungstic acid catalyst on mesoporous
carbon with bimodal pores: a superior catalyst for Friedel-Crafts alkenylation
great significance for practical water purification applications. of aromatics with phenylacetylene, Appl. Catal. A-Gen. 526 (2016) 139–146.
Superiority of GMC-TiO2 should be mainly attributed to its good [6] L. Zeng, C. Zheng, L. Xia, Y. Wang, M. Wei, Ordered mesoporous TiO2–C
adsorption capability, which could provide more opportunities nanocomposite as an anode material for long-term performance lithium-ion
batteries, J. Mater. Chem. A. 1 (2013) 4293–4299.
for CIP and its intermediates to contact with photocatalysts and [7] L. Ji, F. Liu, Z. Xu, S. Zheng, D. Zhu, Adsorption of pharmaceutical antibiotics on
active species, and enhance the local concentration of contami- template-synthesized ordered micro- and mesoporous carbons, Environ. Sci.
nants and increase the driving force of photocatalytic reaction. Technol. 44 (2010) 3116–3122.
[8] J. Chen, F. Qiu, W. Xu, S. Cao, H. Zhu, Recent progress in enhancing
photocatalytic efficiency of TiO2-based materials, Appl. Catal. A-Gen. 495
4. Conclusion (2015) 131–140.
[9] S.P. Xu, X. Sun, Y. Zhao, Y. Gao, Y. Wang, Q.Y. Yue, B.Y. Gao, Carbon-doped
golden wattle-like TiO2 microspheres with excellent visible light
An interesting graphitized mesoporous carbon-TiO2 (GMC-TiO2) photocatalytic activity: simultaneous in-situ carbon doping and single-
nanocomposite was successfully synthesized by an extended R-F crystal nanorod self-assembly, Appl. Surf. Sci. 448 (2018) 78–87.
method under low temperature hydrothermal condition in this [10] Y. Zhao, S.P. Xu, X. Sun, X. Xu, B.Y. Gao, Unique bar-like sulfur-doped C3N4/TiO2
nanocomposite: excellent visible light driven photocatalytic activity and
study. The carbon component in the composite with lamellar mechanism study, Appl. Surf. Sci. 436 (2018) 873–881.
structure provided abundant anchor points for TiO2 to inhibit [11] M. Sturini, A. Speltini, F. Maraschi, A. Profumo, L. Pretali, E.A. Irastorza, E.
agglomeration, resulting in high dispersion of TiO2 nanoparticles Fasani, A. Albini, Photolytic and photocatalytic degradation of
fluoroquinolones in untreated river water under natural sunlight, Appl.
on the carbon matrix. Moreover, strong interaction between the Catal. B-Environ. 119 (2012) 32–39.
two phases stabilized the anatase polymorph in GMC-TiO2 even [12] Y. Zhou, C.H. Chen, N.N. Wang, Y.Y. Li, H.M. Ding, Stable Ti3+ self-doped
at high temperature. Besides, the carbon was graphitized partially, anatase-rutile mixed TiO2 with enhanced visible light utilization and
durability, J. Phys. Chem. C. 120 (2016) 6116–6124.
which was beneficial to the fast transfer of photogenerated elec- [13] R. Leary, A. Westwood, Carbonaceous nanomaterials for the enhancement of
trons and greatly inhibited the electron-hole recombination. All TiO2 photocatalysis, Carbon 49 (2011) 741–772.
of the above remarkably improve the photocatalytic performance [14] B. Wang, G. Zhang, X. Leng, Z. Sun, S. Zheng, Characterization and improved
solar light activity of vanadium doped TiO2/diatomite hybrid catalysts, J.
of GMC-TiO2 in CIP degradation. Meanwhile, the as prepared
Hazard. Mater. 285 (2015) 212–220.
GMC-TiO2 possessed a large specific surface area and mesoporous [15] B. Gao, P.S. Yap, T.M. Lim, T.-T. Lim, Adsorption-photocatalytic degradation of
structure, leading to an excellent CIP adsorption capability, and the Acid Red 88 by supported TiO2: effect of activated carbon support and aqueous
adsorption behavior could be well fitted by pseudo-second-order anions, Chem. Eng. J. 171 (2011) 1098–1107.
[16] J. Liu, Q. Zhang, J. Yang, H. Ma, M.O. Tade, S. Wang, J. Liu, Facile synthesis of
kinetic model and Langmuir isotherm. Benefitted from the strong carbon-doped mesoporous anatase TiO2 for the enhanced visible-light driven
synergy between adsorption and photocatalysis, GMC-TiO2 could photocatalysis, Chem. Commun. 50 (2014) 13971–13974.
X. Zheng et al. / Journal of Colloid and Interface Science 527 (2018) 202–213 213

[17] M. Hamandi, G. Berhault, C. Guillard, H. Kochkar, Reduced graphene oxide/TiO2 [39] X. Sun, S.P. Xu, Y. Gao, M. Yue, Q.Y. Yue, B.Y. Gao, 3D hierarchical golden
nanotube composites for formic acid photodegradation, Appl. Catal. B-Environ. wattle-like TiO2 microspheres: polar acetone-based solvothermal synthesis
209 (2017) 203–213. and enhanced water purification performance, CrystEngComm 19 (2017)
[18] G. Cheng, F. Xu, J. Xiong, F. Tian, J. Ding, F.J. Stadler, R. Chen, Enhanced 2187–2194.
adsorption and photocatalysis capability of generally synthesized TiO2-carbon [40] E. Hosono, S. Fujihara, H. Lmai, I. Honma, I. Masaki, H. Zhou, One-step synthesis
materials hybrids, Adv. Powder Technol. 27 (2016) 1949–1962. of nano-micro chestnut TiO2 with rutile nanopins on the microanatase
[19] J. Xu, A. Wang, T. Zhang, A two-step synthesis of ordered mesoporous octahedron, Acs Nano 1 (2007) 273–278.
resorcinol-formaldehyde polymer and carbon, Carbon. 50 (2012) 1807–1816. [41] Y.Z. Li, N.H. Lee, D.S. Hwang, J.S. Song, E.G. Lee, S.J. Kim, Synthesis and
[20] Q. Ding, Y. Zhang, G. Wang, H. Zhou, H. Zhang, Enhanced photocatalytic characterization of nano titania powder with high photoactivity for gas-phase
activity of a hollow TiO2–Au–TiO2 sandwich structured nanocomposite, RSC photo-oxidation of benzene from TiOCl2 aqueous solution at low
Adv. 6 (2016) 18958–18964. temperatures, Langmuir 20 (2004) 10838–10844.
[21] H. Liu, J.B. Joo, M. Dahl, L. Fu, Z. Zeng, Y. Yin, Crystallinity control of TiO2 hollow [42] S. Goriparti, E. Miele, M. Prato, A. Scarpellini, S. Marras, S. Monaco, A. Toma, G.
shells through resin-protected calcination for enhanced photocatalytic C. Messina, A. Alabastri, F. De Angelis, L. Manna, C. Capiglia, R.P. Zaccaria,
activity, Energ. Environ. Sci. 8 (2015) 286–296. Direct synthesis of carbon-doped TiO2-bronze nanowires as anode materials
[22] X.D. Zhu, Y.J. Wang, R.J. Sun, D.M. Zhou, Photocatalytic degradation of for high performance lithium-ion batteries, ACS Appl. Mater. Interfaces 7
tetracycline in aqueous solution by nanosized TiO2, Chemosphere. 92 (2013) (2015) 25139–25146.
925–932. [43] Y.J. Yang, D.W. Ni, Y. Yao, Y.T. Zhong, Y. Ma, J.N. Yao, High photocatalytic
[23] W.K. Wang, J.J. Chen, X. Zhang, Y.X. Huang, W.W. Li, H.Q. Yu, Self-induced activity of carbon doped TiO2 prepared by fast combustion of organic capping
synthesis of phase-junction TiO2 with a tailored rutile to anatase ratio below ligands, Rsc Adv. 5 (2015) 93635–93643.
phase transition temperature, Sci. Rep. 6 (2016) 20491. [44] K.A. Cychosz, R. Guillet-Nicolas, J. Garcia-Martinez, M. Thommes, Recent
[24] W. Zheng, Z. Yan, Y. Dai, N. Du, X. Jiang, H. Dai, X. Li, G. He, Interpenetrated advances in the textural characterization of hierarchically structured
networks between graphitic carbon infilling and ultrafine TiO2 nanocrystals nanoporous materials, Chem. Soc. Rev. 46 (2017) 389–414.
with patterned macroporous structure for high-performance lithium ion [45] V.K. Mahajan, M. Misra, K.S. Raja, S.K. Mohapatra, Self-organized TiO2
batteries, ACS Appl. Mater. Interfaces. 9 (2017) 20491–20500. nanotubular arrays for photoelectrochemical hydrogen generation: effect of
[25] D. Liu, L.J. Xia, D.Y. Qu, J.H. Lei, Y. Li, B.L. Su, Synthesis of hierarchical fiberlike crystallization and defect structures, J. Phys. D: Appl. Phys. 41 (2008) 125307.
ordered mesoporous carbons with excellent electrochemical capacitance [46] X. Zhang, Z. Zhao, W. Zhang, G. Zhang, D. Qu, X. Miao, S. Sun, Z. Sun, Surface
performance by a strongly acidic aqueous cooperative assembly route, J. defects enhanced visible light photocatalytic H2 production for Zn-Cd-S solid
Mater. Chem. A. 1 (2013) 15447–15458. solution, Small 12 (2016) 793–801.
[26] X. Peng, F. Hu, J. Huang, Y. Wang, H. Dai, Z. Liu, Preparation of a graphitic [47] M. Xing, F. Shen, B. Qiu, J. Zhang, Highly-dispersed boron-doped graphene
ordered mesoporous carbon and its application in sorption of ciprofloxacin: nanosheets loaded with TiO2 nanoparticles for enhancing CO2 photoreduction,
kinetics, isotherm, adsorption mechanisms studies, Micropor. Mesopor. Mater. Sci. Rep. 4 (2014) 6341.
228 (2016) 196–206. [48] J. Shao, W. Sheng, M. Wang, S. Li, J. Chen, Y. Zhang, S. Cao, In situ synthesis of
[27] Z.Y. Jiang, Y.Y. Liu, T. Jing, B.B. Huang, Z.Y. Wang, X.Y. Zhang, X.Y. Qin, Y. Dai, carbon-doped TiO2 single-crystal nanorods with a remarkably photocatalytic
Enhancing visible light photocatalytic activity of TiO2 using a colorless efficiency, Appl. Catal. B-Environ. 209 (2017) 311–319.
molecule (2-methoxyethanol) due to hydrogen bond effect, Appl. Catal. B- [49] Z. Wang, J. Li, F. Tang, J. Lin, Z. Jin, Polydopamine nanotubes-templated
Environ. 200 (2017) 230–236. synthesis of TiO2 and its photocatalytic performance under visible light, RSC
[28] W. Wang, D. Xu, B. Cheng, J. Yu, C. Jiang, Hybrid carbon@TiO2 hollow spheres Adv. 7 (2017) 23535–23542.
with enhanced photocatalytic CO2 reduction activity, J. Mater. Chem. A. 5 [50] B. Li, Z. Zhao, F. Gao, X. Wang, J. Qiu, Mesoporous microspheres composed of
(2017) 5020–5029. carbon-coated TiO2 nanocrystals with exposed 001 facets for improved visible
[29] M. Nawaz, W. Miran, J. Jang, D.S. Lee, One-step hydrothermal synthesis of light photocatalytic activity, Appl. Catal. B-Environ. 147 (2014) 958–964.
porous 3D reduced graphene oxide/TiO2 aerogel for carbamazepine [51] L. Shi, L. Liang, J. Ma, F. Wang, J. Sun, Enhanced photocatalytic activity over the
photodegradation in aqueous solution, Appl. Catal. B-Environ. 203 (2017) Ag2O–g-C3N4 composite under visible light, Catal. Sci. Technol. 4 (2014) 758–
85–95. 765.
[30] H. Sohn, D. Kim, J. Lee, S. Yoon, Facile synthesis of a mesostructured TiO2– [52] M. Xing, X. Li, J. Zhang, Synergistic effect on the visible light activity of Ti3+
graphitized carbon (TiO2–gC) composite through the hydrothermal process doped TiO2 nanorods/boron doped graphene composite, Sci. Rep. 4 (2014)
and its application as the anode of lithium ion batteries, RSC Adv. 6 (2016) 5493.
39484–39491. [53] V. Sharavath, S. Sarkar, D. Gandla, S. Ghosh, Low temperature synthesis of
[31] Z.P. Yao, Y.Q. Meng, Y.J. Zhao, G.J. Liu, Q.X. Xia, J.K. Wang, Z.H. Jiang, A one-step TiO2-b-cyclodextrin–graphene nanocomposite for energy storage and
preparation and enhanced electrochemical properties of C-TiO2 composite photocatalytic applications, Electrochim. Acta 210 (2016) 385–394.
films, Electrochim. Acta 254 (2017) 320–327. [54] S.P. Xu, Y. Gao, X. Sun, M. Yue, Q.Y. Yue, B.Y. Gao, Facile one-pot synthesis of
[32] J. Wan, W. Chen, C. Jia, L. Zheng, J. Dong, X. Zheng, Y. Wang, W. Yan, C. Chen, Q. carbon incorporated three-dimensional hierarchical TiO2 nanostructure for
Peng, D. Wang, Y. Li, Defect effects on TiO2 nanosheets: stabilizing single highly efficient pollutant removal, RSC Adv. 6 (2016) 101198–101207.
atomic site Au and promoting catalytic properties, Adv. Mater. 30 (2018) [55] Y.J. Kan, Q.Y. Yue, B.Y. Gao, Q. Li, Comparative study of dry-mixing and wet-
1705369. mixing activated carbons prepared from waste printed circuit boards by NaOH
[33] E. Bailón-García, A. Elmouwahidi, M.A. Álvarez, F. Carrasco-Marín, A.F. Pérez- activation, RSC Adv. 5 (2015) 105943–105951.
Cadenas, F.J. Maldonado-Hódar, New carbon xerogel-TiO2 composites with [56] W. Wang, X. Ding, M. He, J. Wang, X. Lou, Kinetic adsorption profile and
high performance as visible-light photocatalysts for dye mineralization, Appl. conformation evolution at the DNA-gold nanoparticle interface probed by
Catal. B-Environ. 201 (2017) 29–40. dynamic light scattering, Anal. Chem. 86 (2014) 10186–10192.
[34] G. Li, L. Lv, H. Fan, J. Ma, Y. Li, Y. Wan, X.S. Zhao, Effect of the agglomeration of [57] J. Deng, Y.J. Ge, C.Q. Tan, H.Y. Wang, Q.S. Li, S.Q. Zhou, K.J. Zhang, Degradation
TiO2 nanoparticles on their photocatalytic performance in the aqueous phase, of ciprofloxacin using alpha-MnO2 activated peroxymonosulfate process:
J. Colloid Interface Sci. 348 (2010) 342–347. effect of water constituents, degradation intermediates and toxicity
[35] L. Zhang, Z. Xing, H. Zhang, Z. Li, X. Wu, X. Zhang, Y. Zhang, W. Zhou, High evaluation, Chem. Eng. J. 330 (2017) 1390–1400.
thermostable ordered mesoporous SiO2–TiO2 coated circulating-bed biofilm [58] T.C. An, H. Yang, G.Y. Li, W.H. Song, W.J. Cooper, X.P. Nie, Kinetics and
reactor for unpredictable photocatalytic and biocatalytic performance, Appl. mechanism of advanced oxidation processes (AOPs) in degradation of
Catal. B-Environ. 180 (2016) 521–529. ciprofloxacin in water, Appl. Catal. B-Environ. 94 (2010) 288–294.
[36] I.D. Alonso-Buenaposada, N. Rey-Raap, E.G. Calvo, J. Angel Menéndez, A. [59] Z. Guo, S. Zhu, Y. Zhao, H. Cao, F. Liu, Radiolytic decomposition of ciprofloxacin
Arenillas, Effect of methanol content in commercial formaldehyde solutions on using gamma irradiation in aqueous solution, Environ. Sci. Pollut. Res. Int. 22
the porosity of RF carbon xerogels, J. Non-Cryst. Solids. 426 (2015) 13–18. (2015) 15772–15780.
[37] F. Héroguel, L. Silvioli, Y.P. Du, J.S. Luterbacher, Controlled deposition of [60] C. Jiang, Y. Ji, Y. Shi, J. Chen, T. Cai, Sulfate radical-based oxidation of
titanium oxide overcoats by non-hydrolytic sol gel for improved catalyst fluoroquinolone antibiotics: Kinetics, mechanisms and effects of natural water
selectivity and stability, J. Catal. 358 (2018) 50–61. matrices, Water Res. 106 (2016) 507–517.
[38] S. Okunaka, H. Tokudome, Y. Hitomi, R. Abe, Facile preparation of stable
aqueous titania sols for fabrication of highly active TiO2 photocatalyst films, J.
Mater. Chem. A 3 (2015) 1688–1695.

You might also like