You are on page 1of 10

CARBON 5 0 ( 2 0 1 2 ) 2 4 7 2 –2 4 8 1

Available at www.sciencedirect.com

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

TiO2 nanoparticles loaded on graphene/carbon composite


nanofibers by electrospinning for increased photocatalysis

Chang Hyo Kim a, Bo-Hye Kim b,*


, Kap Seung Yang b,c,**

a
Department of Advanced Chemicals & Engineering, Chonnam National University, 300 Yongbong-dong, Gwangju, Republic of Korea
b
Alan G. MacDiarmid Energy Research Institute, Chonnam National University, 300 Yongbong-dong, Gwangju, Republic of Korea
c
Department of Polymer & Fiber System Engineering, Chonnam National University, 300 Yongbong-dong, Gwangju, Republic of Korea

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

Article history: Graphene/carbon composite nanofibers (CCNFs) with attached TiO2 nanoparticles (TiO2–
Received 29 October 2011 CCNF) were prepared, and their photocatalytic degradation ability under visible light irra-
Accepted 26 January 2012 diation was assessed. They were characterized using scanning and transmission electron
Available online 4 February 2012 microscopy, X-ray diffraction, Raman spectroscopy, X-ray photoelectron spectroscopy,
and ultraviolet–visible diffuse spectroscopy. The results suggest that the presence of graph-
ene embedded in the composite fibers prevents TiO2 particle agglomeration and aids the
uniform dispersion of TiO2 on the fibers. In the photodegradation of methylene blue, a sig-
nificant increase in the reaction rate was observed with TiO2–CCNF materials under visible
light. This increase is due to the high migration efficiency of photoinduced electrons and
the inhibition of charge–carrier recombination due to the electronic interaction between
TiO2 and graphene. The TiO2–CCNF materials could be used for multiple degradation cycles
without a decrease in photocatalytic activity.
 2012 Elsevier Ltd. All rights reserved.

1. Introduction preparation and modification of TiO2 to narrow its band gap and
enhancethephotocatalyticactivityundervisiblelightradiation[10–30].
TiO2 is one of the most promising catalysts because of its supe- Among the carbon nanostructures (e.g., C60, carbon nano-
rior photocatalytic performance, easy availability, long-term tubes, and graphenes), graphenes offer new opportunities in
stability, and nontoxicity [1,2]. Typically, photoexcited elec- photovoltaic conversion and photocatalysis by the hybrid
tron–hole pairs can be generated by irradiation with light with structures with a variety of nanomaterials, due to their excel-
an energy greater than the band gap energy of TiO2 lent charge carrier mobility, a large specific surface area, and
(Ebg = 3.2 eV for anatase). However, several problems can arise good electrical conductivity [31–36]. For example, TiO2 com-
when TiO2 is applied as a photocatalyst: (1) photogenerated bined with graphene acts as an electron trap, which promotes
electron–hole pairs can recombine quickly, which affects the electron–hole separation and facilitates interfacial electron
photocatalytic efficiency [3–7], and (2) TiO2 can only be excited transfer [14,24–26,30]. Furthermore, graphene can help con-
with ultraviolet (UV) light, which is less than 5% of solar light, trol the morphology of TiO2 nanoparticles because it controls
because of its wide band gap [7–9]. These disadvantages of nucleation and growth of TiO2 nanoparticles and allows for
TiO2 result in a low photocatalytic activity in practical applica- optimal chemical interactions and bonding between nanopar-
tion. Therefore, a number of recent studies have focused on the ticles and graphene.

* Corresponding author at : Alan G. MacDiarmid Energy Research Institute, Chonnam National University, Yong-Bong dong 300, 500-757
Gwangju, Republic of Korea. Tel.: +82 62 530 0774; fax: +82 62 530 1779.
** Corresponding author at : Department of Polymer&Fiber System Engineering, Chonnam National University, Yong-Bong dong 300,
500-757 Gwangju, Republic of Korea. Tel.: +82 62 530 0774; fax: +82 62 530 1779.
E-mail addresses: bohye@chonnam.ac.kr (B.-H. Kim), ksyang@chonnam.ac.kr (K.S. Yang).
0008-6223/$ - see front matter  2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.carbon.2012.01.069
CARBON 5 0 ( 20 1 2 ) 2 4 7 2–24 8 1 2473

In this work, carbon nanofibers (CNFs) containing tance between the tip and collector was 17 cm, and the flow
micropores were used to support TiO2 because CNFs exhibit rate of the spinning solution was 3 ml/h. The electrospun fi-
a high adsorption capability and adsorption rate due to the bers were collected on aluminum foil wrapped on a metal
shallow and uniform pore structure. Shallow micropores drum rotating at approximately 300 rpm. The electrospun fi-
open directly to the surface, which results in a large adsorp- ber webs were stabilized in flowing air at 280 C with a heat-
tion capacity and fast adsorption/desorption [37,38]. In addi- ing rate of 1 C/min for 1 h. A 30 mL ethanol solution
tion, electrospinning has been widely used as a versatile containing 2.04 g of titanium n-butoxide (Ti(OnBu)4, STREM
technique to fabricate various hybrid nanofibers [39,40]. How- CHEMICALS) was diluted with 30 mL of toluene. The stabi-
ever, reports have seldom been published on the fabrication lized nanofibers (SNF) were dipped in the prepared sol solu-
and evaluation of TiO2 nanoparticles loaded on graphene/car- tion, washed sufficiently with ethanol and water, dried in
bon composite nanofibers (TiO2–CCNF) produced by electros- air, and carbonized at 800 C with a heating rate of 5 C/min
pinning techniques. in N2 atmosphere with a flow rate of 200 mL/min. This sample
In the present work, TiO2–CCNF catalysts were prepared by was identified as TiO2–CCNF. To compare the photocatalytic
electrospinning to determine if they exhibit improved photo- activity under visible light irradiation, the TiO2 loaded on
catalytic efficiency in the visible spectrum, as illustrated in PAN based carbon nanofiber (TiO2–CNF), Graphene/carbon
Fig. 1. By increasing the photocatalytic activity of the TiO2– composite nanofibers (CCNF), and PAN based carbon nanofi-
CCNF materials, the current study aims to (i) extend the light ber (CNF) samples were synthesized.
absorption spectrum into the visible region, (ii) reduce
electron/hole pair recombination, and (iii) enhance the 2.2. Characterization
adsorption capability and high adsorption/desorption rate of
organic pollutants, which increases the reaction efficiency The surface structures of the TiO2–CCNF materials were char-
because of the high specific surface area of the CNFs. acterized by a Hitachi S-4700 field-emission scanning electron
microscope (FE-SEM) and energy dispersive X-ray spectros-
2. Experimental copy (EDX). The transmission electron micrographs (TEM)
and the selected area electron diffraction (SAED) micrographs
2.1. Materials and methods were taken with a TECHNAI-F20 unit (KJ111, Phillips) operat-
ing at 200 kV. The crystalline structure was characterized with
The graphenes used in this study were xGNP-C750-grade an X-ray diffraction unit (XRD, D-Max-2400 diffractometer)
materials produced by XG Science, USA. The elementary anal- equipped with graphite monochromatized CuKa radiation
ysis of graphene was characterized as 88.68% carbon, 0.79% (k = 0.15418 nm). Backscattering Raman measurements were
hydrogen, 1.11% nitrogen, and 7.65% oxygen using the Mettler carried out with a Renishaw in via-Reflex using 633 nm at
method (Metler-Toledo AG, Switzerland). The graphene with room temperature. The chemical state of the surface was
functional group can be dispersed in organic solvent. The examined by X-ray photoelectron spectroscopy (XPS) using a
3 wt.% grpahene (3 wt.% relative to PAN) sample was im- VGScientific ESCALAB250 spectrometer equipped with a
mersed in dimethylformamide (DMF) and sonicated in a monochromatized AlK X-ray source (15 mA, 14 kV). The
bath-type sonicator. The PAN (Mw = 150,000, Aldrich Chemical surface functionalities of the copolymer were examined by
Co.) was dissolved in the homogeneous graphene-DMF solu- Fourier Transform Infrared spectroscopy (FT-IR, Nicolet 200
tion. This mixture was continuously stirred at 60 C until a instrument). All the samples were analyzed using the KBr pel-
homogeneous solution formed, and it was then cooled to let technique and scanned in the range from 4000 to 400 cm
1
room temperature. This solution was then spun into nanofi- . Specific surface areas were analyzed by the Brunauer–
ber webs using an electrospinning apparatus (NTPS-35K, Emmett–Teller (BET) method using a surface area analyzer
Ntsse Co., Korea) operating at 25 kV. Spinning solutions were (Micrometrics, ASPS2020, USA). To observe the optical proper-
fed through a capillary tip (diameter = 0.5 mm) using a syr- ties of the samples, the photoluminescence (PL) emission
inge (30 ml). The anode of the high voltage power supply spectra (Laser Raman and PL spectrometer, SPEX1403;
was clamped to a syringe needle tip and the cathode was con- 325 nm (55 mW) He–Cd Laser) were measured which give
nected to a metal collector. During electrospinning, the dis- information about the recombination rate of photogenerated

Fig. 1 – TiO2–CCNF hybrids and its response under visible irradiation.


2474 CARBON 5 0 ( 2 0 1 2 ) 2 4 7 2 –2 4 8 1

carriers. The porosity was investigated from the nitrogen act as reactive sites for the nucleation and growth of gold
adsorption isotherm at 77 K (ASAP 2020, Micromeritics, nanoparticles. They observed that the nucleation and growth
USA). The specific surface area, the mesopore size distribu- of these metal nanoparticles were dependent on the density
tion, and the micropore size distribution of the samples were of oxygen functional groups in the graphene surface sheets
evaluated using the Brunauer–Emmett–Teller (BET) method [42]. TiO2 nanoparticles directly grown on graphene appeared
and the Barrett–Joyner–Halenda (BJH) method. Photocatalytic to exhibit strong interactions, which should lead to advanced
efficiencies were evaluated by decolorizing an aqueous solu- hybrid materials for various applications, including
tion of methylene blue (MB) under irradiation by a fluorescent photocatalysis.
lamp. A 0.1 g sample of the photocatalyst powder was dis- TEM images, depicted in Fig. 3a, were collected to obtain
persed in 100 mL of 15 ppm MB (Aldrich) solution by stirring microstructural information for the TiO2–CCNF materials.
with a magnetic stirrer. A 13 W fluorescent lamp (FRX 13EX- The images indicated that TiO2 nanoparticles, with an aver-
D) with light output in the 400–800 nm spectrum range was age particle size of 10–30 nm, had been successfully grown
used as the visible light source. The MB concentration was on the surface of CNFs. SAED measurements (Fig. 3b) con-
measured every 30 min using a UV–Visible spectrophotome- firmed the nanocrystalline nature of the investigated TiO2
ter (Shimadzu 160-A). samples. Based on the SAED ring pattern shown in Fig. 3b,
the clear lattice fringes indicate that the nanoparticles have
3. Results and discussion some crystallinity. In addition, the SAED patterns indicate
that the nanoparticles, identified as TiO2 rutile structures, re-
Fig. 2 presents SEM images of TiO2–CCNF and TiO2–CNF hy- flected from the (1 1 0), (2 1 1), and (1 0 1) lattice planes as in-
brids prepared with and without graphene. The electrospun dexed in Fig. 3b. The corresponding elemental mapping of
PAN based- and PAN/graphene based nanofibers showed a an individual TiO2–CCNF materials provided information on
smooth surface with an average diameter of 180 and the distribution of Ti, O, and C atoms in the fiber (Fig. 3c).
270 nm, respectively, before coating of TiO2. It was observed Fig. 4 presents the XRD patterns of the TiO2–CCNF and
that the nanoparticles were uniformly distributed across the TiO2/CNF materials. The broad peak located between 20
surface of the TiO2–CCNF materials (Fig. 2a) without aggrega- and 30 was attributed to the (0 0 2) plane of the carbon struc-
tion, whereas the TiO2 on pure PAN/based fibers (TiO2–CNF) ture. The diffraction peaks of TiO2 were also clearly observed
consisted of large, aggregated nanoparticles with diameters in Fig. 4a and b. The peak locations for TiO2 are cited from the
in the range of 80–110 nm (Fig. 2b). The graphene, functional- Joint Committee on Powder Diffraction Standards (JCPDS)
ized with carboxylic acid groups, is capable of binding to me- database. The peaks located at 25 Correspond to the (1 0 1)
tal oxide particles, such as TiO2, without aggregation effects plane of the anatase phase (JCPDS 21-1272), and the peaks lo-
and yields small nanparticles [14]. Goncalves et al. [41]. have cated at 28o, 36o, 41o, and 54 Correspond to the (1 1 0), (1 0 1),
shown that oxygen functionalities at the graphene surface (1 1 1), and (2 1 1) planes of the rutile phase (JCPDS 21-1276),

Fig. 2 – SEM images of (a) electrospun PAN based nanofiber, (b) electrospun PAN/graphene based nanofiber, (c) TiO2–CNF, (d)
TiO2–CCNF.
CARBON 5 0 ( 20 1 2 ) 2 4 7 2–24 8 1 2475

Fig. 3 – (a) TEM image of the TiO2–CCNF hybrids. (b) SAED patterns of the TiO2 nanostructures. (c) Elemental mapping of the
individual TiO2–CCNF hybrids.

Fig. 4 – XRD patterns of the (a) TiO2–CCNF and (b) TiO2/CNF.

respectively [43]. All products were confirmed to be a mixture two structural types of TiO2 coexist, which is consistent with
of the anatase and rutile phases. The anatase phase content the reported XRD pattern. In addition to different TiO2 modes,
for all products was calculated from the XRD patterns using the broad D-band (defect-induced mode) at 1340–1360 cm 1
the following equation: Xa = [1 + 1.26(Ir/Ia)] 1, where Xa is the and G band (E2g2 graphite mode) at 1570–1590 cm 1 were
share of anatase in the mixture and Ir and Ia are the integrated observed. The intensity ratio of D band to G band (ID/IG) is pro-
intensities of the (1 0 1) reflection of anatase and the (1 1 0) posed to be an indication of disorder in TiO2–CCNF, TiO2–CNF,
reflection of rutile, respectively. TiO2–CCNF has a 1:1 ratio of and CNF, originating from defects associated with vacancies,
anatase to rutile, whereas the ratio for TiO2–CNF is 1:4. There- grain boundaries, and amorphous carbons. It is seen that
fore, this result indicates that the addition of graphene to TiO2 there is little change in the ID/IG of TiO2–CCNF (1.0547) com-
can effectively suppress the transformation of TiO2 from an pared with TiO2–CNF (1.0551) and CNF (1.0522). This result
anatase to a rutile-type phase and prevents the growth of seems to indicate that graphitic (sp2) nature of the carbon in
TiO2 particles. the photocatalysts remained the same after loading with TiO2.
The Raman spectrum (Fig. 5) indicates the presence of TiO2 XPS analysis can provide valuable insight into the surface
crystalline nanoparticles and showed three Raman bands: structure of the CCNF-supported TiO2 photocatalyst. The Ti2p
144(B1g) for anatase and 431(Eg)/612(A1g) for the rutile structure photoelectron peaks of TiO2–CCNF materials are presented in
ranging between 100 and 800 cm 1 [44]. Fig. 5 shows that the Fig. 6a. The peaks at 465.23 and 459.62 eV represent the Ti2p1/2
2476 CARBON 5 0 ( 2 0 1 2 ) 2 4 7 2 –2 4 8 1

centered at 285.0, 286.3, 287.4, and 289.1 eV, which can be as-
signed to graphitic carbon (C–C and CHn), alcohol or ether car-
bon (C–OH, C–O–C), carbonyl groups (C@O), and carboxyl or
ester carbon (O@C–O), respectively. The O1s core level spec-
trum (Fig. 7b) was deconvoluted into two peaks, indicating
the presence of C@O or Ti–O bond of TiO2 (530.5 eV) and
C–O in ether (O2 , 532.2 eV) [46]. Atomic ratios of titanium to
oxygen for TiO2–CCNF and TiO2–CNF can be calculated
through the rough stoichiometry using the XPS data. The
Ti:O ratios of TiO2–CCNF and TiO2–CNF were determined to
be 1:2.63 and 1:2.50, respectively. However, it has been sup-
posed that the titanium-oxygen ratio is less than 1:2. There
is an overall increase in the Ti:O ratio, which is consistent
with the increase in the proportion of oxygen in the photocat-
alysts. It means that there may be more oxygen species such
as carboxylic acid group of graphene in the TiO2–CCNF photo-
catalyst surface besides the crystal lattice oxygen [46,47].
Fig. 5 – Raman spectrum of TiO2–CCNF hybrids. The FT-IR spectrum of the TiO2–CCNF was reported in
Fig. 8. We can observe absorption at 3439.08 cm 1 (OH broad-
ened band of either alcoholic or hydroxyl groups), 1612.49/
and Ti2p3/2 of Ti4+, respectively, and the Ti3+ peaks at 1454.33 cm 1 (C@C and C–C stretching bands in the aromatic
461.90 eV (Ti2p1/2) and 458.58 eV (Ti2p3/2) were also observed. range, carbonyl), and 1000–1300 cm 1 (C–O stretching band of
However, in the case of TiO2–CNF materials (Fig. 6b), the peaks ether). In particular, the intensity of SNF is larger than that of
of Ti2p1/2 and Ti2p3/2 of Ti4+ were present, whereas the peaks TiO2–CCNF, which represent indirectly that TiO2 nanoparticles
of Ti3+ in Ti2O3 or TiO were barely observed. The Ti3+ species are grown on surface of CCNF with the aid of the oxygen func-
were generated for the TiO2/CCNFs, which might be due to tionalities as reactive sites at the graphene surface. Hence,
the carboxyl groups by graphene. Some of these groups com- the XPS and FT-IR spectra show the presence and decrease
bine with titanium through chelation, and then the positions of the oxygen functional groups, such as –C–O–, –C@O, and
of the oxygen in TiO2 were occupied by carboxyl groups. The –O–C@O, on the TiO2–CCNF catalyst after carbonization.
chelation effect reduce the valence of titanium, whereas in- The photoluminescence (PL) emission spectroscopy has
crease the content of Ti3+ [46]. Generally, the Ti3+ state on been widely used to study the transfer behavior of the photo-
the TiO2 surface plays a dominant role in the photocatalytic generated electrons and holes in semiconductor materials, so
activity because it can trap photogenerated electrons and that it can reflect the separation and recombination of photo-
leave behind unpaired charges to promote photoactivity. generated carries. [48,49]. The measured PL-emission spectra
Therefore, the increasing Ti3+ density promotes effective seg- of TiO2–CCNF, TiO2–CNF, and CNF in the range of 300–850 nm
regation of electron and cavity, interface charge transfer as are presented in Fig. 9. The spectra are broad (extending from
well as lowered the probability of compounding cavity, and 400 to 700 nm) and centered around 485 nm (2.56 eV) for TiO2–
then increases the photocatalysis performance. CCNF and 497 nm (2.50 eV) for TiO2–CNF. The intensity of the
Fig.7 shows the high-resolution XPS spectra of the C1s and PL-spectra decreased in the following order for the photocat-
O1s region on the surface of TiO2–CCNF. The deconvolution of alysts: CNF, TiO2–CNF, and TiO2–CCNF; this result indicates
the C1s XPS spectrum (Fig. 7a) revealed three Gaussian curves that the reduction of the PL-intensity shows the diminution

Fig. 6 – XPS spectra of the deconvoluted Ti(2p) peaks for the (a)TiO2–CCNF and (b) TiO2–CNF materials.
CARBON 5 0 ( 20 1 2 ) 2 4 7 2–24 8 1 2477

Fig. 7 – Deconvolution of (a) C1s, (b) O1s core levels of TiO2–CCNF according to XPS spectra.

photocatalysts [50]. Therefore, the low PL intensity for TiO2–


CCNF indicates that the photocatalytic activities of TiO2 may
be improved due to the interactions between the excited elec-
tron of TiO2 particles and the graphene.
The photocatalytic activities of the TiO2–CCNF, TiO2–CNF,
CCNF, and CNFs were evaluated by measuring the decolor-
ation rate of MB under visible light irradiation (Fig. 10) at
18 C. The experimental results showed that the MB decom-
position rate in the samples occurred in the following order:
TiO2–CCNF > TiO2–CNF  CCNF > CNF. The concentration re-
duced rapidly over the first 30 min, and then the decomposi-
tion rate slowed down or stopped after the amount of time for
all of the samples. As seen in Fig. 10a, the TiO2–CCNF showed
the best photocatalytic activity under visible light irradiation.
The use of the TiO2–CCNF catalyst yielded nearly a 100%
reduction in the first 30 min of irradiation. The MB decompo-
Fig. 8 – FT-IR spectra of the TiO2–CCNF and SNF. sition reaction followed pseudo-first-order kinetics, and the
kinetic constants (k, min 1) were calculated and are pre-
sented in the inset in Fig. 10a. The superior photocatalytic
of the electron/hole pairs recombination process. Generally, activity of TiO2–CCNF materials was also demonstrated in
the lower PL intensity suggests the lower the recombination the degradation of MB. The average rate constant for the
rate of photogenerated electron–hole pairs, which leads to TiO2–CCNF (k = 0.16 min 1) is much larger than those for
the high the photocatalytic activity of semiconductor TiO2–CNF (k = 0.012 min 1), CCNF (k = 0.011 min 1), and CNFs

Fig. 9 – Photoluminescence spectra of (a) TiO2–CCNF, (b) TiO2–CNF, (c) CNF.


2478 CARBON 5 0 ( 2 0 1 2 ) 2 4 7 2 –2 4 8 1

Fig. 10 – (a) Photocatalytic degradation of MB monitored as the concentration change versus irradiation time (inset: average
reaction rate constant (min 1) for the photodegradation of MB with free TiO2). (b) Photocatalytic activity of the TiO2–CCNF
hybrids for MB degradation over five degradation cycles.

(k = 0.003 min 1). Additionally, the stability of the TiO2–CCNF P/P0 = 0.5 was observed, which was typical type II behavior
materials was examined by following the degradation of MB of mesoporous adsorption and micropore filling were ob-
during a five-cycle experiment. After each run, the TiO2–CCNF served at a low relative pressure of P/P0 < 0.1, which was
photocatalyst was evacuated for 30 min and was reused in the typical type I behavior [51]. The pore size distribution in
next run. As shown in Fig. 10b, the photocatalytic degradation TiO2–CCNF, TiO2–CNF, and CNF samples based on the BJH
of MB with TiO2–CCNF phtocatalysts under visible light irradi- method show a broad distribution of mesopores, with sizes
ation was consistently effective over five degradation cycles. ranging from 2–50 nm (Fig. 11b). In particular, TiO2–CCNF
In general, the efficiency of the photodegradation by TiO2 sample has a much larger pore volume around 3.8 nm than
depends on the illuminated catalyst surface area in contact the other samples, suggesting that the presence of mesopores
with the solution and on the mass transfer to the catalyst sur- can lead to adsorption of MB molecular easily. Because the
face [37,38]. Nitrogen adsorption isotherms and pore size dis- molecular size of MB is 1.36 · 0.47 · 0.24 nm [20], and it can
tributions for TiO2–CCNF, TiO2–CNF, and CNF are shown in access pores with diameters larger than 1.5 nm [52]. Further-
Fig. 11. The adsorption isotherms (Fig. 11a) of TiO2–CNF and more, as shown in Table 1, the pore characteristics of the
CNF show typical type I behavior representing the micropo- photocatalysts decreased in the following order: TiO2–CCNF,
rous adsorption, and the adsorption of nitrogen was nearly CNF, and TiO2–CNF; this result indicates that the exposed
complete at a low relative pressure (P/P0 < 0.1). The adsorption CNFs surface with high surface area will be functioned as
isotherms of TiO2–CCNF exhibit combined type I and type II centers of condensing substrates with a physical adsorption
characteristics. Hysteresis at a relative pressure higher than process. Therefore, rapid decomposition of MB in the

Fig. 11 – (a) Nitrogen adsorption isotherms at 77 K, (b) Differential pore volume of various photocatalysts as a function of the
pore diameter.
CARBON 5 0 ( 20 1 2 ) 2 4 7 2–24 8 1 2479

color could be due to consequential processes of adsorption


Table 1 – Pore characteristicsa of the photocatalysts used in
this study. into CNF layer and then decomposition on the surface of
TiO2 particle.
Samples BET surface Total pore Average pore As CNFs and graphene of photocatalysts have some
area volume diameter
adsorption capacity for MB, as shown in Fig. 11 and Table 1,
(m2 g 1) (cm3 g 1) (nm)
a comparison between adsorption and photocatalytic
TiO2–CCNF 434 0.23 2.17 degradation of MB was carried out experimentally with and
TiO2–CNF 361 0.15 1.60 without visible light irradiation to evaluate the actual photo-
CCNF 447 0.17 1.58
catalytic activity. Under the same conditions, adsorption of
CNF 405 0.18 1.81
TiO2–CCNF and TiO2–CNF to MB in the dark was also tested
a
Pore parameters were obtained from N2 adsorption. and the results are shown in Fig. 12. When the catalyst is illu-
minated by visible light irradiation, the MB concentration is
much higher decreased than that without visible light irradi-
ation, indicating that MB is rapidly decomposed by the
TiO2–CCNF catalysts photocatalytically with visible light
irradiation.
The results indicate that TiO2–CCNF has a higher efficiency
for the decomposition of MB than the TiO2–CNF, CCNF, and
CNFs. It was confirmed that the introduction of graphene
and CNFs to the TiO2 photocatalyst could increase the decom-
position rate of some organic compounds using a photocata-
lytic process. The reasons for these improvements are
attributed to the following functions, as shown in Fig. 13.

(i) The graphene with carboxylic acid groups is capable of


binding TiO2 without aggregation effects, resulting in
small nanoparticles. As the particle size of TiO2
becomes small, the photogenerated electrons and holes
can easily migrate to reaction sites on the surface,
thereby reducing the recombination probability [13,45].
Fig. 12 – Adsorption behavior of MB on TiO2–CCNF and TiO2– (ii) In the TiO2–CCNF materials with a mixture of anatase
CNF. and rutile structures, the graphene may act as an elec-
tron acceptor (Fig. 13a) for rutile (Ebg = 3.0 eV) and as a
beginning is reasonably supposed to be due to further adsorp- photosensitizer (Fig. 13b) for anatase (Ebg = 3.2 eV),
tion of MB into CNF layer, and the following gradual change in which prevents the recombination of electron–hole

Fig. 13 – Schematic illustration of photocatalysis with TiO2–CCNF hybrids.


2480 CARBON 5 0 ( 2 0 1 2 ) 2 4 7 2 –2 4 8 1

pairs and promotes photodegradation [13]. This phe- dyes in aqueous dispersions under visible light illumination. J
nomenon may be responsible for extending photocata- Phys Chem B 1999;103:4862–7.
lytic activity into the visible light range. [8] Zhao W, Chen C, Li X, Zhao J, Hidaka H, Serpone N.
Photodegradation of Sulorhodamine-B Dye in Platinum as a
(iii) A large surface area is necessary for the photocatalytic
Functional Co-Catalyst. J Phys Chem B 2002;106:5022–8.
degradation of organic compounds because adsorption
[9] Zhang Z, Shao C, Zhang L, Li X, Yichun L. Electrospun
of the organic compound is an important process. The nanofibers of V-doped TiO2 with high photocatalytic activity. J
exposed fibers with high-surface-area TiO2–CCNF Colloid Interface Sci 2010;351:57–62.
materials will function as centers of substrate conden- [10] Chen X, Mao SS. Titanium dioxide nanomaterials: synthesis,
sation in the physical adsorption process. There the properties modifications and applications. Chem Rev
peculiar properties of CNFs are expected to elevate 2007;107:2891–5.
[11] Liu G, Wang L, Yang HG, Cheng HM, Lu GQ. Titania-based
the photocatalytic activities of TiO2.
photocatalysts-crystal growth, doping and heterostructuring.
J Mater Chem 2010;20:831–43.
[12] Wang W, Serp P, Kalck P, Faria JL. Visible light
4. Conclusion photodegradation of phenol on MWNT-TiO2 composite
catalysts prepared by a modified sol–gel method. J Mol Catal
A: Chem 2005;235:194–9.
Small TiO2 nanoparticles were successfully deposited on elec-
[13] Leary R, Westwood A. Carbonaceous nanomaterials for the
trospun CCNFs by the sol–gel method, and these composites
enhancement of TiO2 photocatalysis. Carbon 2011;49:741–72.
were very active photocatalysts in the photodegradation of [14] Liang Y, Wang H, Casalongue HS, Chen Z, Dai H. TiO2
MB under visible light irradiation. Furthermore, TiO2–CCNF nanocrystals grown on graphene as advanced photocatalytic
materials could be recycled without a decrease in the photo- hybrid materials. Nano Res 2010;3:701–5.
catalytic activity. The results suggest that graphene acts as an [15] Li Y, Li X, Li J, Yin J. TiO2-coated active carbon composites
electron acceptor and a photosensitizer, which causes an in- with increased photocatalytic activity prepared by a properly
controlled sol–gel method. Mater Lett 2005;59:2659–63.
crease in the photodegradation rate and reduces electron–
[16] Kominami H, Kohno M, Matsunaga Y, Kera Y. Thermal
hole pair recombination. In addition, the CNFs’ high surface
decomposition of titanium alkoxide and silicate ester in
area synergistically improved the photocatalytic activity of organic solvent: a new method for synthesizing large-
TiO2 by enhancing the physical adsorption of the substrate. surface-area, silica-modified titanium(iv) oxide of high
Therefore, TiO2–CCNF materials are expected to be a promis- thermal stability. J Am Ceram Soc 2001;84:1178–80.
ing candidate for photocatalytic processes using sunlight [17] Chen H, Shi J, Zhang W, Ruan M, Yan D. Incorporation of
irradiation. titanium into the inorganic wall of ordered porous zirconium
oxide via direct synthesis. Chem Mater 2001;13:1035–40.
[18] Chun H, Yizhong W, Hongxiao T. Influence of adsorption on
Acknowledgements the photodegradation of various dyes using surface bond-
conjugated TiO2/SiO2 photocatalyst. Appl Catal B Environ
This research was supported by the Basic Science Research 2001;35:95–105.
Program through the National Research Foundation of Korea [19] Takeda N, Torimoto T, Sampath S, Kuwabata S, Yoneyama H.
Effect of inert supports for titanium dioxide loading on
(NRF) funded by the Ministry of Education, Science and
enhancement of photodecomposition rate of gaseous
Technology (No. 2011-0005890) and the Ministry of Education, propionaldehyde. J Phys Chem 1995;99:9986–91.
Science and Technology (MEST) (K20903002024-11E0100- [20] Ding Z, Lu GQ, Greenfield PF. A kinetic study on
04010). photocatalytic oxidation of phenol in water by silica-
dispersed titania nanoparticles. J Colloid Interface Sci
2000;232:1–9.
[21] Hirano M, Nakahara C, Ota K, Inagaki M. Direct formation of
R E F E R E N C E S
zirconia-doped titania with stable anatase-type structure by
thermal hydrolysis. J Am Ceram Soc 2002;85:1333–5.
[22] Kim B-H, Nataraj SK, Yang KS, Woo H-G. Synthesis,
[1] Zhang J, Xu Q, Feng Z, Li M, Li C. Importance of the characterization, and photocatalytic activity of TiO2/SiO2
Relationship between Surface Phases and Photocatalytic nanoparticles loaded on carbon nanofiber web. J Nanosci
Activity of TiO2. Angew Chem In Ed 2008;47:1766–9. Nanotechnol 2010;10:3331–5.
[2] Ni M, Leung KH, Leung DYC, Sumathy K. A review and recent [23] Luo Y, Heng Y, Dai X, Chen W, Li J. Preparation and
developments in photocatalytic water-splitting using TiO2 for photocatalytic ability of highly defective carbon nanotubes. J
hydrogen production. Renew Sust Energy Rev 2007;11:401–25. Solid State Chem 2009;182:2521–5.
[3] Fox MA, Dulay MT. Heterogeneous photocatalysis. Chem Rev [24] Williams G, Seger B, Kamat PV. TiO2-graphene
1993;93:341–57. nanocomposites. UV-assisted photocatalytic reduction of
[4] Heller A. Chemistry and applications of photocatalytic graphene oxide. ACS Nano 2008;2:1487–91.
oxidation of thin organic films. Acc Chem Rev 1995;28:503–8. [25] Zhang H, Lv X, Li Y, Wang Y, Li J. P25-graphene composite as a
[5] Blake DM, Maness PC, Huang Z, Wolfrum EJ, Huang J, Jacoby high performance photocatalyst. ACS Nano 2010;4:380–6.
WA. Application of the photocatalytic chemistry of titanium [26] Jiang G, Lin Z, Chen C, Zhu L, Chang Q, Wang N, et al. TiO2
dioxide to disinfection and the killing of cancer cells. Sep nanoparticles assembled on graphene oxide nanosheets with
Purif Methods 1999;28:1–50. high photocatalytic activity for removal of pollutants. Carbon
[6] Kamat PV. Photochemistry on nonreactive and reactive 2011;49:2693–701.
(semiconductor) surfaces. Chem Rev 1993;93:267–300. [27] Tsumura T, Kojitani N, Umemura H, Toyoda M, Inagaki M.
[7] Wu T, Liu G, Zhao J, Hidaka H, Serpone N. Evidence for H2O2 Composites between photoactive anatase-type TiO2 and
generation during the TiO2-assisted photodegradation of adsorptive carbon. Appl Surf Sci 2002;196:429–36.
CARBON 5 0 ( 20 1 2 ) 2 4 7 2–24 8 1 2481

[28] Modestov AD, Lev O. Photocatalytic oxidation of 2,4- nanosheets with gold nanoparticles: the role of oxygen
dichlorophenoxyacetic acid with titania photocatalyst. moieties at graphene surface on gold nucleation and growth.
Comparison of supported and suspended TiO2. J Photochem Chem Mater 2009;21:4796–802.
Photobiol A Chem 1998;112:261–70. [42] Kamat PV. Graphene-Based Nanoarchitectures. Anchoring
[29] Kim CH, Kim B-H, Yang KS. Visible light-induced Semiconductor and Metal Nanoparticles on a Two-
photocatalytic activity of Ag-containing TiO2/carbon Dimensional Carbon Support. J Phys Chem Lett. 2010; 1: 520–
nanofibers composites. Synthetic Metals 2011;161:1068–72. 527.
[30] Wenqing F, Qinghua L, Qinghong Z, Ye W. [43] Dai S, Wu Y, Sakai T, Du Z, Sakai H, Abe M. Preparation of
Nanocomposites of TiO2 and reduced graphene oxide as highly crystalline TiO2 nanostructures by acid-assisted
efficient photocatalysts for hydrogen evolution. J Phys hydrodrothermal treatment of hexagonal-structured
Chem C 2011;115:10694–701. nanocrystalline titania/cetyltrimethyammonium bromide
[31] Robel I, Bunker BA, Kamat PV. SWCNT-CdS nanocomposite as nanoskeleton. Nanoskeleton Nanoscale Res Lett
light harvesting assembly. Photoinduced charge transfer 2010;5:1829–35.
interactions. Adv Mater 2005;17:2458–63. [44] Zhou Q-f, Wu S-h, Shang Q-q, Zhang J-x. Structure and
[32] Oh WC, Zhang FJ, Chen ML. Synthesis and characterization of Raman Spectroscopy of Nanocrystalline TiO2 Powder Derived
V-C60/TiO2 photocatalysts designed for degradation of by Sol-Gel Process. Chin Phys Lett. 1997; 14: 306-369.
methylene blue. J Ind Eng Chem 2010;16:299–304. [45] Kudo A, Miseki Y. Heterogeneous photocatalyst materials for
[33] Geim AK, Novoselov KS. The rise of graphene. Nat Mater water splitting. Chem Soc Rev 2009;38:253–78.
2007;6:183–91. [46] Deqing M, DaiQi Y. Surface study of composite photocatalyst
[34] Bolotin KI, Sikes KJ, Jiang Z, Klima M, Fudenberg G, Hone J, based on plasma modified activated carbon fibers with TiO2.
et al. Ultrahigh electron mobility in suspended graphene. Surf Coat Tech 2009;203:1154–60.
Solid State Commun 2008;146:351–5. [47] Steven FD, Nilson CC, Elidiane CR, Mario ABM. Plasma
[35] Wang HL, Robinson JT, Diankov G, Dai HJ. Nanocrystal growth enhanced chemical vapor deposition of titanium (IV)
on graphene with various degrees of oxidation. J Am Chem ethoxide-oxygen-helium mixtures. Thin Solid Films
Soc 2010;132:3270–1. 2008;516:4940–5.
[36] Si YC, Samulski ET. Exfoliated graphene separated by [48] Tang H, Berger H, Schmid PE, Levy F, Burri G.
platinum nanoparticles. Chem Mater 2008;20:6792–7. Photoluminescence in TiO2 anatase single crystals. Solid
[37] Amjad HES, Alan PN, Hafid A-D, Suki P, Neil C, Steven Y. State Commun 1993;87:847–50.
Deposition of anatase on the surface of activated carbon. Surf [49] Jin J, Marshal D, Shin J-H, Kim J-C, Getoff N. Surface
Coat Tech 2004;187:284–92. properties and photoactivity of TiO2 treated with electron
[38] Liu J-H, Yang R, Li S-M. Preparation and application of beam. Radiat Phys Chem 2006;75:583–9.
efficient TiO2/ACFs photocatalyst. J Environ Sci [50] Xu L, Zhongqing L, Jian Z, Xin Y, Dandan L, Si C, et al.
2006;18:979–81. Characteristics of N-doped TiO2 nanotube arrays by N2-
[39] Kim SK, Lee SK, Hwang SH, Lim H. Photocatalytic deposition plasma for visible light-driven photocatalysis. J Alloys Compd
of silver nanoparticles onto organic/inorganic composite 2011;509:9970–6.
nanofibers. Macromol Mater Eng 2006;291:1265–70. [51] Wang D-W, Li F, Gao ML, Lu Q, Cheng H-M. 3D aperiodic
[40] Drew C, Liu X, Ziegler D, Wang XY, Bruno FF, Whitten J, et al. hierarchical porous graphitic carbon material for high-rate
Folding transition of large DNA completely inhibits the action electrochemical capacitive energy storage. Angew Chem Int
of a restriction endonuclease as revealed by single-chain Ed 2008;47:373–6.
observation. Nano Lett 2003;3:143–6. [52] Yuan R, Guan R, Shen W, Zheng J. Photocatalytic degradation
[41] Goncalves G, Marques PAAP, Granadeiro CM, Nogueira HIS, of methylene blue by a combination of TiO2 and activated
Singh MK, Gracio J. Surface modification of graphene carbon fibers. J Colloid Interface Sci 2005;282:87–91.

You might also like