You are on page 1of 14

Chemical Engineering Journal 284 (2016) 202–215

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Physicochemical characteristic of regenerated cellulose/N-doped TiO2


nanocomposite membrane fabricated from recycled newspaper with
photocatalytic activity under UV and visible light irradiation
Mohamad Azuwa Mohamed a,b, W.N. W. Salleh a,b,⇑, Juhana Jaafar a,b, A.F. Ismail a,b,
Muhazri Abd Mutalib a,b, N.A.A. Sani a,b, S.E.A. M. Asri c, C.S. Ong a
a
Advanced Membrane Technology Research Centre, Universiti Teknologi Malaysia, 81310 Skudai, Johor Bahru, Malaysia
b
Faculty of Petroleum and Renewable Energy Engineering, Universiti Teknologi Malaysia, 81310 Skudai, Johor Bahru, Malaysia
c
Department of Chemistry, Faculty of Science, Universiti Teknologi Malaysia, 81310 Skudai, Johor Bahru, Malaysia

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

 Fabrication of novel green


photocatalytic membrane. Reduction

e- e- e- e-
 The utilization of recycled newspaper Feed CB hv
e- e-
hv CB
as the cellulose source. 3d Ti3+
3d Ti3+

 Conversion of dense to porous hv OH OH OH OH


OH OH- hv hv OH-
Abs
OH + + + + Abs
membrane structure by the addition +
+ +
+ h+ h+ h+
N-doping level
h+
+ + + + Oxidation
N-doping level
Oxidation
TiO2 nanorods. RC/ TiO2 + + VB
VB

+ + +
Nanocomposite + + Rutile
 RC/TiO2-0.5 shows highest catalytic Membrane
+
+
+
+ + Anatase OH
+ + OH
+ + +
activity under UV and visible light + + +
+
+
hv = light irradiation
irradiation. Phenol molecules e- = electron at conduction band
h+ = positive hole at valence band

 Potential to be applied as a
VB = Valence band
H2O molecules Permeate CB = Conduction band
Abs = Photocatalyst absorb light
OH
photocatalytic membrane for +
Hydroxyl radical

Positive holes

wastewater treatment.

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

Article history: The use of recycled newspaper as sustainable cellulose resource for the fabrication of green organic/
Received 8 June 2015 inorganic hybrid photocatalytic membrane via phase inversion method was highlighted in this study.
Received in revised form 2 August 2015 The incorporation of N-doped TiO2 nanorods as a nanocomposite in regenerated cellulose membrane
Accepted 27 August 2015
matrix to great extent has altered its morphological and physicochemical properties, as revealed by
Available online 3 September 2015
FESEM, AFM, FTIR, XRD, XPS, and UV–visible spectroscopy analyses. FTIR analysis suggested that there
is a strong interaction between the hydroxyl groups of regenerated cellulose (RC) and the TiO2 nanorods
Keywords:
through hydrogen bonding interactions. The UV–visible spectroscopy and XPS analysis confirmed that
Inorganic–organic hybrid
Green photocatalytic membrane
the highly visible light absorption capability of the prepared RC/TiO2 nanocomposite membrane is due
Physicochemical characteristic to the existence of nitrogen as dopant in the TiO2 lattice structure. The resultant membranes showed a
UV and visible light active significant photocatalytic performance in the degradation of phenol in aqueous solution under UV and
Photocatalytic properties visible light irradiation, respectively. It was found that 0.5 wt% of TiO2 nanorods was the best loading
in the regenerated cellulose membrane (RCM) with desirable physicochemical and photocatalytic
properties. This study promotes the use of RC/TiO2 nanocomposite membrane as a new and green
portable photocatalyst in the field of wastewater treatment without leaving any photocatalyst in the
reaction system. It is crucial to emphasize that the use of a non-toxic solvent-based system in this study
provide a significant contribution towards the development of a green technology system.
Ó 2015 Elsevier B.V. All rights reserved.

⇑ Corresponding author at: Advanced Membrane Technology Research Centre,


Universiti Teknologi Malaysia, 81310 Skudai, Johor Bahru, Malaysia.
E-mail address: hayati@petroleum.utm.my (W.N. W. Salleh).

http://dx.doi.org/10.1016/j.cej.2015.08.128
1385-8947/Ó 2015 Elsevier B.V. All rights reserved.
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 203

Abbreviations Q volume of the permeate water per unit time (m3 s1)
PWF pure water flux Ww weight of the wet membrane (g)
RC regenerated cellulose Wd weight of the dry membrane (g)
RCM regenerated cellulose membrane qH density of water (0.998 g/cm3)
RC/TiO2 regenerated cellulose/titanium dioxide qc density of cellulose (1.5 g/cm3)
FESEM field emission scanning electron microscope rm membrane mean pore radius (nm)
EFTEM energy filter transmission electron microscopy l membrane thickness (m)
FT-IR fourier transform infrared DP load pressure (Pa)
XPS X-ray photoelectron spectroscopy Ra mean surface roughness (nm)
XRD X-ray diffraction spectroscopy C0 initial concentration at time t = 0
AFM atomic force microscopy Ct concentration at time interval

Symbols Greek symbols


Eg band gap energy e membrane porosity
J water flux (L m2 h1) h water contact angle
V volume of permeate (L) g water viscosity (8.9  104 Pa s)
Dt time (s)
A area (m2)

1. Introduction chain and increase the energy to break down polymer chain [17].
The feasibility of regenerated cellulose (RC)/TiO2 nanocomposites
Currently, the advent of inorganic–organic nanocomposite membrane in water and wastewater treatment has been
membrane which combines the processability of polymers and extensively studied. For instance, Zeng et al. (2010) proposed
the superior properties of inorganic materials has captured the TiO2 immobilization in cellulose matrix for photocatalytic
attention of researchers, owing to the unique advantages of this degradation of phenol under weak UV light irradiation [13]. Zhu
novel membrane in comparison to the conventionally-made et al. (2012) developed a novel inorganic–polymer hybrid mem-
polymeric membrane. Previous works have shown that the brane by the incorporation of nano-TiO2 into RC with high perfor-
incorporation of inorganic materials such as in TiO2, ZnO2, and mance for dehydration of caprolactam by pervaporation [18].
Al2O3 in polymeric-based membranes could potentially improve Zhang and co-workers prepared bacterial cellulose/TiO2 composite
membrane stability and separation performance [1]. For example, membrane doped with rare earth elements, and evaluated its pho-
the utilization of TiO2 particles in a membrane modification has tocatalytic properties [8]. They found that, the resultant compos-
been shown that the TiO2 have a good stabilization, is a ites membrane has high strength, ultrafine nanoporosity, and
hydrophilicity agent, anti-bacterial, has anti-fouling character, water absorption characteristics, whereas the photocatalysis effi-
and photocatalytic properties [2–4]. ciency was significantly enhanced after the TiO2/Bacterial cellulose
Nowadays, photocatalytic technology has become one of membrane was doped with rare earth ions. Furthermore, the
efficient and green approaches for the elimination of hazardous obtained RC/TiO2 nanocomposites membrane also exhibited high
pollutants in wastewater treatment [5–7]. TiO2 has recently UV–vis light absorption [19]. Moreover, reusable photocatalytic
attracted substantial attention and has been proven to be the most titanium dioxide-cellulose based films showed the potential for
promising catalyst with strong oxidation activity. However, for degradation of organic molecules in natural water sources [20,21].
large scale applications, TiO2 as a photocatalyst has two main Previous studies have utilized various sources of cellulose in the
drawbacks; (1) its photocatalysis oxidation rates for many target fabrication of regenerated cellulose membrane (RCM) such as cot-
pollutants are too slow to be practically applied, and (2) it tends ton linter, soft wood pulp, and microcrystalline cellulose [22–24].
to agglomerate and difficult to separate in the purpose of catalyst To the best of our knowledge, there is still no recorded study on
recycling [8]. This issue can be solved by introducing nanomateri- cellulose from recycled newspapers in photocatalytic membrane
als in membrane matrix, in addition to offering great promises in application in the literature. The RCM made of recycled newspaper
the wastewater treatments [9–12]. has a great interest towards sustainable future and control the
Cellulose is one of the potential candidates for supporting TiO2 white pollution [25]. The main drawback of pure TiO2 is its wide
nanoparticles due to its superfine networks structure. The band gap (3–3.2 eV). Particularly, it absorbs only the UV part of
advantages of superfine networks are they do not only provide solar radiation that accounts for only 4% of the total solar radiation,
mechanical support but also help to disperse the inorganic leaving most of the visible light irradiation [26]. Therefore, it is
nanoparticles and to improve the particles stability, retain the important to develop photocatalyst that can be utilized under vis-
special morphology, and control the growth of nanoparticles by ible light. Among the reported photocatalysts with visible-light
providing a template surface for nucleate precipitation [13–15]. response, TiO2 doped with nitrogen has been extensively studied
The nanoscale of cellulose fibers is approximately 10–100 nm in because of its comparable atomic size with oxygen, small ioniza-
the form of a web-like network microstructure, which makes tion energy, eco-friendly, relatively high stability and simple syn-
cellulose one of the most highly porous materials [16]. In addition, thesis methods [27–29]. Therefore, it is important to study the
cellulose nanofibers can act as an attractive matrix material for the potential, feasibility and compatibility of both materials; cellulose
suspension of photocatalytic particles due to their desirable from recycled newspaper and N-doped TiO2 nanorods for the
mechanical and optical properties [17]. There is a good compatibil- preparation of photocatalytic membrane. The application of this
ity between the TiO2 nanoparticles and cellulose chain. The photocatalytic membrane can overcome the difficulty in recollect-
interaction is due to covalent bonds between TiO2 nanoparticles ing and removal of TiO2 suspension in water after photocatalytic
with cellulose chain that can improve the rigidity of the polymer treatment. Furthermore, this approach is a truly green process
204 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

and cost-effective in terms of its development, preparation and 2.4. Photocatalytic nanocomposite membrane preparation
application.
In this study, the novel regenerated cellulose/N-doped TiO2 Fig. 1 shows the preparation of RC/TiO2 nanocomposite
nanocomposites membranes were prepared by phase inversion membranes by phase inversion technique. Specific amount of
method with the utilization of recycled newspaper as the cellulose TiO2 was added into a mixture of NaOH and urea in distilled water
source in NaOH/urea aqueous solvent system. The physicochemical (7:12:81 by wt%) and was stirred for 2 h and sonicated for 30 min.
characteristic of the nanocomposite membrane was studied using Subsequently, the desired amount of cellulose was immediately
field emission scanning electron microscopy (FESEM), atomic force dispersed into the aqueous solution under continuous vigorous
microscopy (AFM), X-ray diffraction (XRD), X-ray photoelectron mechanical stirring for 10 min, and the slurry was pre-cooled to
spectroscopy, and UV–visible spectroscopy analyses. The photocat- 15 °C for 24 h. The frozen sample was thawed and stirred exten-
alytic performance of the resultants photocatalytic membrane was sively at room temperature to obtain an almost transparent 4 wt%
further investigated by using phenol as a model of pollutants. On of cellulose/TiO2 dope solution. The resulting dope solution was
the other hand, for the photocatalytic experiment, photocatalytic centrifuged to eliminate air bubbles and discard the remaining
membrane reactor was used in this study, involving a flat-sheet undissolved portion at 4000 rpm for 1 h. The cellulose/TiO2 dope
dead-end filtration mode. The results suggested that the obtained solution was casted on a glass plate and immediately immersed
nanocomposites membrane may provide a new and green photo- in 5 wt% of H2SO4 coagulant bath for 10 min at 25 °C for coagula-
catalytic composite membrane to be applied in wastewater tion and regeneration. The membrane was detached and washed
treatment. using distilled water to remove excess H2SO4 on the surface of
resultant membrane. Finally, the wet membrane was transferred
on a glass plate and fixed with adhesive tapes to prevent shrinkage,
2. Experimental and later was dried at ambient temperature to obtain RC/TiO2
membrane. The same procedures were repeated by varying the
2.1. Materials loading of TiO2 from 0 to 0.7 wt%. The samples were denoted as
RC/T-0, RC/T-0.3, RC/T-0.5, and RC/T-0.7, respectively.
Non-printed area of recycled newspaper was used as cellulose
source. Titanium-n-butoxide Ti(OBu)4 from Sigma–Aldrich was 2.5. Membrane characterization
used as the titanium precursor. NaOH, urea, H2SO4, Nitric acid
(HNO3 65%), and isopropanol (C3H7OH) were purchased from QReC The membrane surface and cross-sections of the samples were
chemicals. The entire chemicals used were of analytical grade. Dis- examined using field-emission scanning electron microscope
tilled water was used throughout this experiment. (FESEM) (JEOL JSM-7600F, Japan). The dry membrane samples
were immersed in liquid nitrogen and fractured, followed by
sputter-coating with platinum using a sputtering device. The
2.2. Synthesis of N-doped TiO2 nanorods
membrane surface roughness was conducted using atomic force
microscopy (AFM) technique. The AFM measurement was per-
The N-doped TiO2 nanorods were synthesized according to our
formed using XE-100 Park System with SSS-NCHR non-contact
previous study [30]. Titanium precursor (Ti(OBu)4) was added
probes at 1 lm/s of scan speed. The particles size analysis of the
dropwise into the isopropanol solution under magnetic stirring
prepared TiO2 nanorods was conducted by using EFTEM LIBRA-
to obtain a homogeneous mixture. The mixture was then added
120, Carl Zeiss AG Company (Oberkochen, Germany). Fourier-
dropwise into distilled water to stimulate hydrolysis of titanium
transform infrared (FTIR) spectra of the membrane samples were
precursor. 30 min of hydrolysis-polymerization process was
performed using attenuated total reflectance Fourier transform
allowed under vigorous magnetic stirring. Then, HNO3 was added
infrared spectroscope (ATR-FTIR) (Nicolet 5700, Thermo Electron
dropwise into the mixture and stirred for 30 min. The resultant
Scientific Instruments Corporation). These spectra were recorded
mixture followed a volume ratio of 25:8:200:3 for Ti(OBu)4:
at resolution between 650 and 4000 cm1. The crystallinity of the
C3H7OH:H2O:HNO3. The prepared mixture was then aged in tight
samples was determined by using Siemens X-ray diffractometer
air for several days until the formation of yellowish sol–gel was
(XRD) D5000 with CuKa radiation of wavelength 0.15406 nm at
observed. The gel was then dried at 75 °C for 74 h in a vacuum oven
40 kV and 40 mA. The diffracted intensity was measured by scan-
until yellowish powder was obtained and subsequently calcined at
ning range of 2h = 5–60° with a step speed of 2°/min. The optical
400 °C for 2 h at heating rate of 5 °C min1. The resultants powder
property of the prepared sample was evaluated using UV–vis-NIR
were washed using distilled water and isopropanol for several
spectrophotometer (UV-3101PC Shidmadzu) between 200 and
times and dried at 60 °C for 24 h.
600 nm. The X-ray photoelectron spectroscopy (XPS) spectra of
the prepared samples were attained by means of Kratos Analytical
2.3. Extraction of cellulose Axis Ultra DLD photoelectron spectrometer using Al Ka radiation
monochromatic source. The porosity of prepared films was deter-
Cellulose was extracted according to our previous study [31]. mined by using gravimetric method as expressed in the following
The recycled newspaper was ground until its fibrous strand was equation [32]
obtained and then dried in oven at 50 °C for 24 h. The ground recy- ðW w W d Þ
cled newspaper was treated with 5 wt% of NaOH at 100 °C for 4 h. e ð%Þ ¼ ðW w WqdHÞ ð1Þ
The ratio between ground recycled newspaper and NaOH solution qH þ Wq d
c
was 1:40 (g/mL). Consequently, bleaching treatment was carried
out using NaClO2 solutions. A few drops of 60 wt% HNO3 were where Ww and Wd are the weight of wet and dry membrane (g),
added dropwise into 800 mL 2 w/v% of NaClO2 solution at 100 °C respectively. qH is the density of water (0.998 g/cm3) and qc is the
under vigorous mechanical stirring. After that, the sample was density of cellulose (1.5 g/cm3). The mean pore radius rm, was calcu-
immediately added into NaClO2 solution and boiled at 100 °C for lated accordingly to the Guerout–Elford–Ferry equation (Eq. (2))
4 h under continuous vigorous mechanical stirring. The pH of the rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð2:9  1:7eÞ  8glQ
sample was neutralized and then was dried in the oven at 50 °C rm ¼ ð2Þ
for 24 h.
e  A  DP
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 205

Fig. 1. RC/TiO2 nanocomposite membrane formation process.

where g is the water viscosity (8.9  104 Pa s), l the membrane intensity 96 mW/cm2). The spectra region of white vis-LED emis-
thickness (m), Q is the volume of the permeate water per unit time sion is showed in Supplementary information. The experiments
(m3 s1), and DP the load pressure (Pa). The nanocomposite mem- were conducted in a stainless steel box having dimension of
brane water flux (J) was determined according to (Eq. (3)) [33] (60  35  42) cm. The schematic diagram of the flat-sheet dead-
end filtration set up used is illustrated in Fig. 2.
V
J¼ ð3Þ In order to evaluate the photocatalytic activity of the prepared
A Dt RC/TiO2 nanocomposite membrane with different TiO2 loadings
where J is the water flux (L/m2h), Q is the quantity of permeate (L), A (0–0.7 wt%), 40 ppm of 500 mL phenol was used as a water
is the effective membrane area (m2), and t is time (h) to obtain the pollutant model. Prior to each experiment, phenol solution was
quantity of Q. The contact angle of membranes was determined by oxygenated in the dark for 1 h to achieve adsorption/desorption
the sessile drop technique using a contact angle goniometer equilibrium by recirculating the permeate solution back into the
(Model: OCA 15EC, Dataphysics) with deionized water as the liquid. reactor with constant flow rate of 1.5 mL/min. Then, 3 mL of
At least 10 locations were arbitrarily chosen on the membrane sur- permeate was collected as a blank sample, prior to irradiation
face in order to yield an average value. effect, and was treated as the initial concentration of phenol (C0).
Subsequently, the solution was irradiated under UV with constant
2.6. Evaluation of photocatalytic nanocomposite membrane recirculation flow rate of 1.5 mL/min for 30 min. Then, the light
performance was turned off, and 3 mL of treated permeate was collected from
the reactor. The remaining permeate was recirculated back to
The photocatalytic membrane reactor (PMR) used in this study make sure the homogenous condition in the tank and the light
involved a flat-sheet dead-end filtration mode, in which the feed was turned on again to continue the degradation process. The same
was passed through the membrane to become filtrate (permeate). procedure was repeated to collect treated permeates for the next
The PMR device comprised of the cube filtration membrane cell 30 min interval within 6 h of the experimental period. The degra-
unit integrated with stainless steel filter, irradiation sources, an dation of phenol was monitored by using a Perkin Elmer UV–vis
air diffuser and a pump mounted at the permeate side of the mem- spectroscopy to measure the change of phenol concentration
brane cell. The cube filtration membrane cell unit was made from throughout the experiment at 296.35 nm. The performance of the
acrylic with an inner dimension of 15  15 cm and a depth of prepared nanocomposite membrane was evaluated by the concen-
16 cm. The lamps were installed 10 cm above the filtration cell. tration of phenol for each collected sample. The photocatalytic
The UV irradiation is obtained from an ultraviolet (UV) lamp (Vil- activity was expressed in the percentages of phenol degradation
ber Laurmat, France, VL-115-M, k = 312 nm, 30 W, light intensity according to the following equation (Eq. (4)):
1.45 mW/cm2). On the other hand, the visible light irradiation is C0  Ct
obtained from white light-emitting diode (LED) flood light (Wuhan Degradation of phenol ¼  100% ð4Þ
Ct
Co-Shine Technology Co., Ltd, China, CS-FL, k > 420 nm, 30 W, light
206 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

UV/Visible Lamp

Stainless steel
reactor body frame
Membrane
filtration cell

Peristaltic pump

Permeate
collector

Tubing

Fig. 2. Illustration of PMR with dead-end membrane filtration cell.

where C0 is the initial concentration at time t = 0, and Ct is the con- in obvious change on the nanoscopic morphologies, as shown in
centration at time interval. The same experiment was repeated by the cross sectional images in Fig. 3(e, f, h and i). By the addition
using visible light as the irradiation source. A blank experiment of TiO2 nanorods, substantial effect on the structural morphology
was also carried out using direct photolysis of phenol, irradiated of the RC membrane was observed, which in a way produced addi-
under UV and visible light, respectively and without the presence tional voids at the polymer chain. Similarities were also observed
of TiO2 in the membrane. After the experiment, the membrane cell elsewhere [37,38]. Thus, the addition of TiO2 nanoparticles had a
unit was washed using distilled water to remove all the remaining large effect on the membrane structure, which had converted the
pollutants. dense RC membrane into nanoporous membrane structure. The
porous structure is beneficial for pollutants adsorption capability
within the membrane pores and enhanced the contact area
3. Results and discussion
between the photocatalyst and the pollutants in the membrane
which will improve the photodegradation process. Furthermore,
3.1. Morphology analysis
this transformation might offer spaces for water molecules to per-
meate through the membranes due to the existing extra free vol-
The morphological structure of the nanocomposite membrane
umes within the polymer chains. It is interesting to note that, the
prepared by phase-inversion technique was observed by using
pore size of RC/TiO2-0.7 (Fig. 3(i)) was smaller than RC/TiO2-0.5
FESEM, as shown in Fig. 3. As can be seen in Fig. 3(a), the free sur-
(Fig. 3(f)). The high TiO2 loading in RC/TiO2-0.7 led to the aggrega-
face FESEM image of parent RC presented smooth surface with no
tion of TiO2 nanoparticles, which might have blocked the pores of
defects. The cross-sectional FESEM images (Fig. 3(b and c))
the membrane and reduce the porosity. The values of porosity and
revealed that the RC exhibited homogeneous dense symmetric
mean pore size are tabulated in Table 1. The prepared membrane
membrane structure consisted of skin layer. The skin layer was
could be considered in the range of ultrafiltration (1 nm–0.1 lm).
formed by the increase in polymer concentration promoted by
The modification of RC membrane with the addition of TiO2
the swift solvent depletion from the top layers of the membrane
nanorods revealed an increase in the surface roughness as
solution during regeneration in the coagulation bath [33]. The
compared to unmodified RC membrane. Fig. 4 shows the three
average thickness of the RC and its skin layer were determined
dimensional view of AFM images (10 lm  10 lm) of RC/TiO2
to be 42 lm and 737 nm, respectively. Fig. 3(d and g) show the
nanocomposite membranes with different TiO2 loading (wt%).
FESEM images of the RC nanocomposite membranes incorporated
The surface roughness of the prepared membranes was obtained
of TiO2 nanorods, namely RC/TiO2-0.5 and RC/TiO2-0.7, respec-
from AFM analysis software. As can be seen in Fig. 4, the surface
tively. The average size of the as-synthesized TiO2 nanorods was
roughness of prepared membrane increased with the addition of
determined by using TEM (see Fig. 3(j)), having length of 70 nm
TiO2 nanorods. RC/TiO2-0.7 exhibited the highest surface rough-
and width of 16 nm. Generally, it was clearly seen that highly
ness, followed by RC/TiO2-0.5, and RC with Ra values of 79.11 nm,
uniform TiO2 nanorods were well dispersed on the surface of the
55.93 nm, and 29.53 nm, respectively. Coarser membrane was
RC/TiO2 nanocomposite membranes as compared to previous
required in the preparation of RC/TiO2 nanocomposite membrane
studies which suffered TiO2 agglomeration [18]. This indicates that
with high photocatalytic activity properties. High surface rough-
there was a remarkable compatibility existing between cellulose
ness is a way to improve the hydrophilic properties. It is well
and TiO2 nanorods anchored onto the surface of the RC/TiO2
known that the higher surface roughness could contribute to
nanocomposite membrane. However, there was a slight
higher fouling in wastewater treatment. However, in the case of
agglomeration of TiO2 nanorods observed on the surface of RC/
incorporation of TiO2 in the polymeric membrane, the fouling phe-
TiO2-0.7 due to the nano-size effect of TiO2 nanoparticles that
nomena could be covered by the degradation activity of the TiO2
was easy to aggregate [34].
itself.
By the incorporation of TiO2 prior to RCM fabrication, an isotro-
pic mixed matrix membrane was formed, without the formation of
skin layer. The absence of the skin layer of RC/TiO2 nanocomposite 3.2. Fourier transform infra-red (FTIR) analysis
membrane was due to the presence the TiO2 nanorods which
might reduce the concentration of polymers on the top and thus, Fig. 5 shows the FTIR spectra of RC/TiO2 nanocomposite mem-
reducing the tendency of the skin layer formation. Similar brane with different TiO2 loading (0–0.7 wt%) and inset is the
observation also can be found elsewhere [35,36]. Furthermore, chemical structure of cellulose. As shown in Fig. 5, the FTIR analysis
the addition of TiO2 nanorod in the RC membrane matrix resulted revealed that the produced RC/TiO2 nanocomposite membrane
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 207

(a) (b) (c)

Skin layer
1μm 30 μm 10 μm

(d) (e) (f)

Pores
TiO2 nanorods
TiO2 nanorods

1μm 30 μm 100 nm

(g) (h) (i)


TiO2 nanorods TiO2 nanorods

Agglomeration Pores

1μm 30 μm 100 nm

(j)

TiO2 nanorods
200 nm

Fig. 3. (a–c) FESEM images of RC, (d–f) FESEM images of RC/TiO2-0.5, (g–i) RC/TiO2-0.7 and (j) TEM image of TiO2 nanorods. (a, d and g) are surface images, (b, c, e, f, h and i)
are cross-sectional images.

which was shifted from 1428 cm1 (cellulose I) in t-CMF with


Table 1 decrease in absorption peak. As compared to the unmodified RC,
The membrane characteristic with different TiO2 nanorods wt% loading. it was observed that O–H stretching absorption peaks were slightly
TiO2 Water Surface Porosity Mean pore Pure water shifted from high to low absorption peak, and the intensity
loading contact roughness e (%) size rm flux J increased as TiO2 concentration increased in all prepared RC/TiO2
(wt%) angle (h) Ra (nm) (nm) (L h1 m2) nanocomposite membranes. This was attributed by the interaction
0 55.68 29.53 41.03 12.38 0.69 (±0.02) between the O–H groups of RC and Ti–O bond of TiO2. In addition,
(±3.83) (±2.37) (±0.41) previous study suggested that this might be due to a strong inter-
0.3 39.78 39.90 52.64 24.66 1.69 (±0.22)
action between the hydroxyl groups of RC and the TiO2 particles
(±4.12) (±3.21) (±1.28)
0.5 31.47 55.93 64.14 30.22 4.12 (±0.17)
through hydrogen bonding interactions [43]. Zhang and co-
(±2.46) (±0.47) (±3.21) worker also suggested that it might be due to the partial C–OH
0.7 24.54 79.11 60.87 29.57 2.36 (±0.15) occupied by a Ti–O bond [8]. Therefore, it promoted the strong
(±2.98) (±0.89) (±2.25) interfacial interaction between RC chain and TiO2 nanorods, as
schematically shown in Fig. 6. The summary of main absorbance
showed the characteristic peaks of cellulose II with evident main peaks of interest of this study is presented in Table 2.
absorption peak shift [39]. The band at 891 cm1 (cellulose II)
was assigned to C–O–C stretching at b-linked glucose of cellulose,
which was shifted from 897 cm1 (cellulose I) in t-CMF and 3.3. X-ray diffraction (XRD) analysis
RNP-pulp [31,40,41]. Similar observation was also found in
previous study [42]. In addition, the transformation from cellulose The XRD patterns of t-CMF and RC/TiO2 nanocomposite mem-
I to cellulose II was also observed at shoulder band of 1419 cm1 brane with different TiO2 wt% loading are shown in Fig. 7. It was
(cellulose II). This peak was assigned to CH2 symmetry bending, seen that t-CMF displayed a typical crystal lattice of cellulose I with
208 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

Fig. 4. 3D AFM images of (a) RC, (b) RC/TiO2-0.5, and (c) RC/TiO2-0.7.
1163

(a) intensity peak at 2h = 19.8–21.3° corresponding to the (1 1 0 crystal


897
1645
2918

1428

plane) and (2 0 0 crystal plane), respectively. These peaks were


1104
1315

(b)
3336

attributed to the typical crystal lattice of cellulose II. Similar XRD


1030

(c)
2905

pattern was also observed in previous studies [8,23]. The results


(d) revealed that the crystal lattice of cellulose in RC produced from
Transmittance (a.u)

NaOH/urea aqueous solution transformed from cellulose I into cel-


3336

(e)
lulose II. Furthermore, the transformation of cellulose I into cellu-
1419
1650

lose II was in good agreement with the FTIR analysis. It was seen
1312
1154
1370

891

that the peak intensity of RC/TiO2 membrane was lower than t-


2895

CMF, which indicated lower crystallinity as a result of rearrange-


3274

1015

ment of the cellulose macromolecules during dissolution and


regeneration [46]. In addition, it was seen that the main diffraction
peaks of anatase (1 0 1) and rutile (1 1 0) were observed at 2h = 25°
O-H stretching shifted to
lower absorption peak and 2h = 27.5° in all RC/TiO2 nanocomposite membrane. The pres-
ence of these peaks indicated incorporation of anatase/rutile mixed
Cellulose chain structure
phase TiO2 nanorods in the RC polymeric chain. It has been sug-
4000 3500 3000 2500 2000 1500 1000 650 gested that, anatase/rutile mixed phase TiO2 nanorods will exhibit
Wavenumber (cm-1) high photocatalytic activity as compared to single constituents of
TiO2 due to synergistic effect [47]. Therefore, the abundance of
Fig. 5. FTIR spectra of (a) t-CMF, (b) RC, (c) RC/TiO2-0.3, (d) RC/TiO2-0.5, and (e) RC/ anatase/rutile mixed phase TiO2 nanorods in the RC polymeric
TiO2-0.7. Inset: chemical structure of the cellulose.
chain could promote the production of high photocatalytic activity
main diffraction at 2h = 15.9 °C (1 1 0 crystal plane), 22.8° (2 0 0 RC/TiO2 nanocomposite membrane. Moreover, the peak intensity
crystal plane) and 34.6° (0 0 4 crystal plane) peak pattern [44,45]. of TiO2 in the membrane nanocomposite was lower when
Moreover, there were two peaks observed at 2h = 12.7° and low compared to pure TiO2, suggesting that it might be attributed to
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 209

22.8°

(110)
A – Anatase

(004)
(200)
15.9° 34.6°
R - Rutile
t-CMF
12.7° 19.8° - 21.3°

(200)
(110)
Cellulose chain structure RC

12.7°
O-H O-H O-H O-H O-H O-H O-H

Intensity (a.u)

(110)
A R RC/TiO2-0.3
12.7°

RC/TiO2 nanocomposite membrane

(110)
A R RC/TiO2-0.5

O-H O-H O O-H n 12.7°

partial C-OH Ti A

(110)
R RC/TiO2-0.7
was occupied by
a Ti-O bond Ti O
A R

(110)
(101)
Ti O
TiO2 nanorods
Hydrogen
bond
interaction H-O Ti O 10 15 20 25 30 35 40 45 50 55
TiO2 nanorods 2θ (degree)

Fig. 7. XRD pattern of RC/TiO2 nanocomposite membrane with different TiO2


nanorods loading.
Fig. 6. Schematic illustration of the interaction between the hydroxyl groups of RC
and the TiO2 particles in RC/TiO2 nanocomposite membrane.

100
UV region Visible region
Table 2 90
FT-IR absorption band for functional group of t-CMF and RC/TiO2 nanocomposite
membrane. 80
RC
Peak assignment Peak frequency (cm1) 70 RC/TiO2-0.3
Transmittance (%)

RC/TiO2-0.5
t- RC RC/TiO2- RC/TiO2- RC/TiO2- 60
CMF 0.3 0.5 0.7
RC/TiO2-0.7
O–H stretching 3336 3334 3287 3283 3281
50
C–H stretching 2918 2905 2895 2895 2895
40
O–H bending 1635 1650 1650 1650 1650
CH2 symmetric bending 1428 1419 1419 1419 1419 30
C–H bending 1375 1369 1369 1366 1372 RC
CH2 wagging 1315 1313 1315 1315 1312 20
C–O anti-symmetric 1163 1154 1154 1151 1154
stretching 10
C–O and C–C stretching 1104 1105 1105 1105 1105 RC/TiO2-0.5
C–O–C stretching 1053 1053 1053 1053 1054 0
C–O and C–C stretching 1030 1015 1020 1014 1015 200 300 400 500 600 700 800
C–H 897 891 891 891 891 Wavelength (nm)

Fig. 8. Optical transmittance spectra of the RC/TiO2 nanocomposite membrane with


different TiO2 nanorods loading. Inset is the photograph of the RC and RC/TiO2-0.5
the domination of RC matrix. It is interesting to note that the char- membrane.
acteristic peak of RC was unchanged after blending with TiO2.
Henceforth, the incorporation of TiO2 nanoparticles into cellulose
did not alter the crystalline structure of the RC. transmittance value about 75%, 72%, and 60%, respectively. These
results showed that all the prepared RC/TiO2 nanocomposite mem-
3.4. Optical properties branes exhibited excellent transparency, which indicated remark-
able homogeneity of the composite membrane [46]. It was
The optical transmittance was conducted to study the trans- observed that the cellulose microfiber was dissolved in NaOH/urea
parency of the prepared membranes. Fig. 8 shows the optical trans- aqueous system, which resulted in high optical transparencies. The
mittance spectra of the RC/TiO2 nanocomposite membrane with transparency decreased with the increment of TiO2 loading from 0
different TiO2 nanorods loading. The membranes with relatively to 0.7 wt%. This was due to the fact that the tendency of high TiO2
low TiO2 nanorods content had a high transmittance in the visible nanorods content in the RC matrix might lead to the increase of
region, and decreased towards the ultraviolet region, where the membrane opacity, consequently low light would diffuse through
band gap of TiO2 was located. The transmittance values of mem- the RC/TiO2 nanocomposite membrane. Similar observations can
brane samples were compared at 550 nm, since the human eye also be found elsewhere [21,49]. In addition, previous study sug-
has the highest sensitivity at this wavelength [48]. The membranes gested that the decrease in optical transmittance was ascribed to
were transparent at lower TiO2 content but tended to become the uniform dispersion of nanofiller in the RC matrix as a result
milky as the concentration of nanoparticles increased, as shown of the interfacial interactions between TiO2 and RC. Therefore,
in the inset photograph in Fig. 8. It was clear that the parent RC the optical transmittance results obtained were in good agreement
membrane possessed high transmittance up to nearly 90%, with the FESEM analysis. In the case of photocatalytic application,
followed by RC/TiO2-0.3, RC/TiO2-0.5, and RC/TiO2-0.7, with the the optical properties of transparent nanocomposite membrane
210 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

(a) the presence of nitrogen atoms in TiO2 lattice structure which


UV region Visible region
induced the local states above the valence band edge, while the
oxygen vacancies gave rise to the local states containing of the
3d states of Ti3+ below the conduction band edge [52]. Therefore,
the results suggested a considerably narrower band gap suitable
for the photocatalytic activity under visible light irradiations, as
compared to the P25 photocatalyst.

3.5. X-ray photoelectron spectroscopy (XPS) analysis


RC/P25
RC
The X-ray photoelectron spectroscopy (XPS) was carried out to
RC/TiO2-0.3
investigate the surface composition and chemical state of the
RC/TiO2-0.5
RC/TiO2-0.7 nanocomposite membrane. Fig. 10 shows the XPS survey spectra
of RC/TiO2-0.5 membrane and as-synthesized N-doped TiO2
nanorod as reference. All the C, O, Ti, and N elements are presented
in Fig. 10(a). The inset of Fig. 10(a) shows the high resolution XPS
200 300 400 500 600 700 800
spectra of N 1s peak with the range of the binding energy from 404
Wavelength (nm)
to 394 eV. The existence of the binding energy at 396.8, 397.5,
398.7, and 399.8 eV confirmed the presence of the N-doping in
(b) the TiO2 lattice structure responsible for visible light absorption
capability in RC/TiO2 nanocomposite membrane, as discussed in
Section 3.3 [53,27,51,54–58]. As shown in Fig. 10(b), it is notewor-
thy that XPS results clearly ruled out the existence of C as a dopant
in the N–TiO2 photocatalysts because the doped carbon in TiO2
exhibited a very low C 1s binding energy at 281–282 eV [30]. The
(αhv)1/2

peaks of the binding energies corresponding to C 1s were observed


at 283.1, 284.4, 286.8, and 288.2 eV, assigned to C, C–C or C–H, C–
O, and C@O [28,59]. These peaks were assigned to adventitious car-
bon contamination species from XPS measurement [60,61]. Fig. 10
RC/P25 (c) shows the high resolution of C 1s peak in the range of binding
RC/TiO2-0.3 energy from 292 to 280 eV for sample RC/TiO2-0.5. The binding
RC/TiO2-0.5 energy detected at 284.6, 285.8, 287.1, and 289.8 eV assigned to
RC/TiO2-0.7 (C–C–OH bonding), (O–C–O or C@O bonding), (O–C–O bonding),
and (O–C–O or C@O bonding), respectively, revealing that the car-
2.0 2.5 3.0 3.5 4.0 bon atoms came from cellulose molecules [62,63].
Eg (eV)
3.6. Physical characteristic analysis
Fig. 9. (a) UV–vis absorption spectra for RC/TiO2 nanocomposite membrane with
different TiO2 nanorods loading. (b) (ahm)1/2 versus vs the energy of absorbed light Based on the contact angle analysis, the RC exhibited the high-
afford the band gaps of the different samples.
est contact angle, followed by RC/TiO2-0.5, and RC/TiO2-0.7 with
the equilibrium contact angle found to be 55.68°, 31.47° and
24.54° respectively. Low value in the surface contact angle indi-
are imperative for the maximization of UV or visible light irradia- cated higher hydrophilicity. The improvement in the hydrophilic-
tion throughout the membrane matrix. These optical properties of ity was due to the presence of TiO2 nanorods polarity [64]. The
RC/TiO2 nanocomposite membrane can lead to the enhancement of polar group of TiO2 could interact with water molecules through
electron distribution and transfer to the surface of TiO2. As a result, van der Waal’s force and hydrogen bond. As the hydrophilicity
this will increase the photocatalytic activity of TiO2 in treating increased, it led to the increase of chemisorbed –OH content on
pollutants. the surface of TiO2 nanorods and led to the formation of superficial
The UV–vis absorption spectra of prepared RC/TiO2 nanocom- hydroxyl group [65]. It is well-known that the TiO2 superficial
posite membranes are illustrated in Fig. 9, referred from P25. As hydroxyl group plays an important role in the photocatalytic activ-
portrayed in Fig. 9(a), it was observed that there was no significant ity [66].
absorption capability for RC membrane in both UV and visible The high hydrophilicity properties promotes high water fluxes
regions. As the N-doped TiO2 nanorods were incorporated in RC and can avoid membrane fouling. Compared to the modified RC
matrix, all the RC/TiO2 nanocomposite membrane samples exhib- membrane, the unmodified RC membrane displays a lower value
ited excellent absorption capability in UV and visible region. In of pure water flux (PWF). Unmodified RC membrane exhibits dense
addition, the UV and visible light absorption capability increased structure with the lowest porosity, mean pore size, and
with the increment of TiO2 content from 0.3 to 0.7 wt% in the RC hydrophilicity, which resulted to the lowest PWF. The incorpora-
matrix. This observation obviously indicated that the addition of tion of TiO2 nanorods in the RC membrane improved the PWF as
higher N-doped TiO2 content could effectively absorb higher frac- shown in Table 1. The prepared RC/TiO2 nanocomposite mem-
tion of photons in UV and visible region. Kubelka–Munk function branes exhibit the maximum pure water flux (4.12 ± 0.17 L h1 -
was used to estimate the band gap energy of the prepared samples m2) with the addition of TiO2 nanorods up to 0.5 wt%. Shi and
[50,51]. As TiO2 is an indirect transition semiconductor, plots of the co-worker have postulated that this is mainly due to the morphol-
(ahm)1/2 versus the energy of absorbed light afford the band gaps of ogy and structural changes of the membrane arising from the crys-
the different samples as shown in Fig. 9(b). The obtained Eg for RC/ tallization changes with different TiO2 nanorods loading [67]. As
TiO2-0.3, RC/TiO2-0.5, RC/TiO2-0.7 and RC/P25 were found to be mentioned before, the addition of TiO2 nanorods in the RCM matrix
3.07, 2.94, 2.83, and 3.24 eV. The band gap narrowing arose from leads to the formation of the free spaces in the polymer chain. The
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 211

(a) O 1s N 1s 397.5 eV
398.7 eV

Intensity (cps)
399.8 eV 396.8 eV
Ti 2P

Intensity (cps) 404 402 400 398 396 394

Ti 2s Binding energy (eV)

N -TiO2 nanorod

C 1s
Ti 3P
Ti 3s O 2s
N 1s
RC/TiO2 – 0.5

600 500 400 300 200 100 0


Binding energy (eV)

(b) C 1s (c) C 1s
283.1 eV
285.8 eV
Intensity (cps)

Intensity (cps)

284.4 eV
287.1 eV 286.2 eV

286.8 eV 289.8 eV
284.6 eV
288.2 eV

292 290 288 286 284 282 280 292 290 288 286 284 282 280
Binding Energy (eV) Binding Energy (eV)

Fig. 10. XPS spectra of N-TiO2 nanorod and RC/TiO2-0.5 membrane. (a) The survey XPS spectra for N-TiO2 nanorod and RC/TiO2-0.5 membrane. Inset is high resolution XPS
spectra of N 1s peak. (b and c) High resolution XPS spectra of C 1s of N-TiO2 nanorod and RC/TiO2-0.5 membrane respectively.

formation of the free spaces could increase PWF of the nanocom-


posite membrane. However, as the TiO2 nanorods reached 0.7 wt 35
%, the PWF value started to decrease. This might be due to the
blockage of the pores of the membrane which resulted from the 30
pore-plugging of the nanocomposite membrane at higher TiO2
nanorods loading. Moreover, Rahimpour and co-worker have also 25
Adsorption (%)

suggested that the PWF decreases with the increasing amount of


TiO2 nanorods deposition. It is attributed to pore-plugging which 20
decreases its PVDF/sulfonated PES membrane flux [68]. Similar
finding can also be found elsewhere [64]. 15
RC/TiO2-0.5
10 RC/TiO2-0.7
3.7. Photocatalytic activity properties of RC/TiO2 nanocomposite
RC/TiO2-0.3
membrane
5 RC

0
3.7.1. Adsorption of phenol in the dark 0 20 40 60 80 100 120 140
The preliminary studies of phenol adsorption at 25 °C were con- Time (min)
ducted on all prepared RC/TiO2 nanocomposite membranes for
120 min. As shown in Fig. 11, it can be seen that, most of Fig. 11. Kinetics of phenol adsorption in the dark.
adsorption occurred within 50 min in all samples. After 120 min
adsorption of phenol in the dark, RC/TiO2-0.5 exhibited the highest
phenol adsorption, followed by RC/TiO2-0.7, RC/TiO2-0.3, and RC explained by the facts that all samples possessed different porosity
with the adsorption value of 30%, 26%, 24%, and 19%, respectively. and mean pore size. It has been reported that, as the porosity and
The differences in adsorption of phenol on all samples could be mean pore size increases, high adsorption sites of the catalyst can
212 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

4.0
(a) RC/TiO2-0.5 (c) 4.0 RC/TiO2-0.5
3.5 RC/TiO2-0.7 kapp= 9.8 x 10-3 min-1 kapp= 9.8 x 10-3 min-1
RC/P25
RC/TiO2-0.3
3.0 RC/AA
RC 3.0
Photolysis kapp= 7.5 x 10-3 min-1
2.5
ln (C0/C)

ln (C0/C)
2.0 kapp= 4.9 x 10-3 min-1
2.0

1.5 kapp= 3.8 x 10-3 min-1

1.0 kapp= 2.9 x 10-3 min-1


1.0

0.5
kapp= 3.0 x 10-4 min-1
0.0 kapp= 4.0 x 10-5 min-1 0.0
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Irradiation Time (min) Irradiation Time (min)

2.0
(b) 2.0 (d) RC/TiO2-0.5
RC/TiO2-0.5
kapp= 4.8 x 10-3 min-1 kapp= 4.8 x 10-3 min-1
RC/TiO2-0.7 RC/P25
RC/TiO2-0.3 RC/AA
1.5 1.5
RC
ln (C0/C)

ln (C0/C)
Photolysis
kapp= 2.8 x 10-3 min-1 1.0
1.0

kapp= 1.4 x 10-3 min-1 0.5


0.5
kapp= 6.0 x 10-4 min-1
kapp= 3.0 x 10-4 min-1
kapp= 5.0 x 10-4 min-1
0.0 kapp= 4.0 x 10-5 min-1 0.0
0 60 120 180 240 300 360 0 60 120 180 240 300 360

Irradiation Time (min) Irradiation Time (min)

Fig. 12. Kinetic of disappearance of phenol by RC/TiO2 nanocomposite membrane with different TiO2 loading (0–0.7 wt%) (a and b) and comparison study consist of RC/TiO2-
0.5, RC/AA, and RC/P25 (c and d) under UV and visible light irradiation respectively.

Reduction

e- e- e- e-
e- e-
Feed CB hv
3d Ti3+ hv CB
3d Ti3+

hv OH OH OH OH hv OH-
OH OH- hv Abs
OH + + + + Abs
+ + h+ h+ h+ h+
+ + + + N-doping level N-doping level
+ + Oxidation Oxidation
RC/ TiO2 + + VB
VB

+ + +
Nanocomposite + + Rutile
+ + + Anatase
Membrane + + + OH OH
+
+ + +
+ +
+ + +
hv = light irradiation
Phenol molecules e- = electron at conduction band
h+ = positive hole at valence band
VB = Valence band
H2O molecules Permeate CB = Conduction band
Abs = Photocatalyst absorb light
OH Hydroxyl radical

+ Positive holes

Fig. 13. Illustration of degradation process of phenol with RC/TiO2 nanocomposite membrane. Enlarge is the photocatalytic mechanism over the N-TiO2 anatase/rutile mixed
phase.

be obtained [69,70]. Therefore, RC/TiO2-0.5 was expected to have 3.7.2. Photocatalytic activity
high adsorption capability as compared to others due to the fact The photocatalytic degradation kinetics of the phenol aqueous
that it had the highest porosity and mean pore size, as a result from solution with different TiO2 nanorods contents under UV and visi-
TiO2 nanorods incorporation, as shown in Table 1. It is well known ble irradiation are shown in Fig. 12(a and b), respectively. In gen-
that, for a heterogeneous photocatalytic system in principle, the eral, all RC/TiO2 nanocomposite membrane exhibited similar
reaction presumably occurs on the surface of the catalyst surface. photocatalytic activity pattern in UV and visible light irradiation,
Therefore, the adsorption of the phenol is crucial to promote high respectively. However, the photocatalytic activity under UV was
photocatalytic reaction. Prior to the photodegradation experiment, higher than under visible light irradiation. This might be due to
a longer period of 60 min was considered in order to reached the TiO2 nanorod in the RC absorbing more photon fractions in
adsorption/desorption equilibrium in the dark. the UV region. The photocatalytic of the phenol in aqueous solution
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 213

Table 3 morphology of the nanocomposite membrane with enhanced pho-


Phenol degradation percentage at different TiO2 loading wt% under UV and visible tocatalytic properties.
light irradiation.
For comparison study, 0.5 wt% of commercial anatase TiO2 from
Samples Ads (%)a kapp (min1) Degradation (%)b Sigma–Aldrich (TAA) and Degussa P25 (TP25) also had been incor-
UV light Visible light UV light Visible light porated within the RC membrane using NaOH/Urea aqueous solu-
tions system as described in Section 2.4. The resultant membranes
RC 19 3.0  104 3.0  104 8.0 8.0
RC/TiO2-0.3 24 3.8  103 1.4  103 85.0 38.0 were denoted as RC/P25 for Degussa P25 and RC/AA for anatase
RC/TiO2-0.5 30 9.8  103 4.8  103 96.6 78.8 Sigma–Aldrich. The photocatalytic activity of these membranes
RC/TiO2-0.7 26 4.9  103 2.8  103 89.0 52.9 was compared with that of RC/TiO2-0.5 under UV and visible irra-
Photolysis – 4.0  105 4.0  105 1.8 1.8 diation, as shown in Fig. 12(c and d) respectively. The finding
a
Adsorption of phenol in the dark within 60 min. showed that RC/TiO2-0.5 exhibited the highest photocatalytic
b
Photocatalytic activity under UV and visible light irradiation for 360 min activity, followed by RC/P25 and RC/AA with kapp values of
obtained from Eq. (4). 9.8  103, 7.5  103, and 2.9  103 min1, respectively, under
UV irradiation, as shown in Fig. 12(c). This result indicated that
the TiO2 nanorods prepared possessed the highest photocatalytic
can be regarded as pseudo-first-order kinetics model, thus it can be
activity with the RC membrane as compared to the commercial
assumed as follows: [6,71]
P25 and AA photocatalyst. Furthermore, the photocatalytic activi-
dC t ties of RC/TiO2-0.5 and RC/P25 were comparable and higher than
r¼ ¼ kC t ð5Þ RC/AA due to the fact that both membrane consisted of anatase/
dt
rutile mixed phase TiO2 photocatalyst due to synergistic effect
 
C0 between the two components as confirmed in XRD analysis [73].
ln ¼ kapp t ð6Þ It has been suggested that, the anatase/rutile mixed phase TiO2
Ct
photocatalyst can improve the charge carrier separation through
where r is the reaction rate, Ct the concentration of aqueous phenol electron trapping in rutile, and consequently reduce the electron
at reaction time t, C0 the initial phenol concentration and kapp the recombination to maintain the formation of radical species and
appararent rate constant. The apparent rate constant for all RC/ improve the photocatalytic activity [47,74]. However, in the test,
TiO2 nanocomposite membrane were obtained to be 9.8  103, the photocatalytic activity of RC/P25 and RC/AA were very low
4.9  103, 3.8  103, and 3.0  104 min1 in UV irradiation under visible irradiation with the kapp value of 6.0  104 and
(Fig. 12(a)) and 4.8  103, 2.8  103, 1.4  103, and 3.0  104 - 5.0  104 min1, as shown in Fig. 12(d). This was due to the both
min1 in visible irradiation (Fig. 12(b)) for RC/TiO2-0.5, RC/TiO2-0.7, P25 and AA incorporated in the RC membrane were not capable of
RC/TiO2-0.3, and RC respectively; indicative of an increasing cat- absorbing photon from visible light. The RC/TiO2-0.5 consisted of
alytic activity order of RC/TiO2-0.5 < RC/TiO2-0.7 < RC/TiO2- nitrogen dopant, which was incorporated in the TiO2 lattice struc-
0.3 < RC. The percentage of phenol degradation was obtained using ture, had introduced localized N 2p states above the valence band,
Eq. (4). The direct photolysis showed relatively the lowest degrada- as shown in Fig. 13 (enlarged image) which improved the absorp-
tion of phenol in aqueous solution, with only 1.8% degradation tion capability under visible light irradiation [75]. N doping con-
under UV and visible light irradiations for 360 min. It was con- verted some Ti4+ to Ti3+ by charge compensation, and the 3d
firmed that phenol would not be degraded by UV and visible light orbital of the Ti3+ states in TiO2 formed a donor energy below the
irradiation alone [72]. Therefore, the direct photolysis of phenol conduction band, which also contributed to the visible light
was negligible in this study. The RC/TiO2-0.5 sample exhibited the absorption in N-doped TiO2 [52]. The higher visible light photocat-
highest photocatalytic activity under UV and visible light irradiation alytic activity observed was correlated to the combined band gap,
for 360 min followed by RC/TiO2-0.7, RC/TiO2-0.3 and RC with the narrowing the effect of the nitrogen doping and electron hole sep-
degradation percentage of 96.6%, 89%, 85% and 8% in UV irradiation arating effect of anatase/rutile mixed phase. Therefore, the differ-
and 78.8%, 52.9%, 38.0%, and 8% in visible irradiation, as shown in ences in photocatalytic activity were most likely dependent on
Table 3. It was clearly shown that the photocatalytic activity of morphological and surface properties of photocatalyst used which
RC/TiO2 nanocomposite membrane increased along with the addi- have been described in our previous study [76].
tion of concentration of TiO2 increase from 0 to 0.5 in UV and visible
light irradiation, respectively. However, as the TiO2 nanoparticles
concentration increased up to 0.7, the photocatalytic activity of 4. Conclusions
RC/TiO2 nanocomposite membrane decreased. This might be due
to the agglomeration of TiO2 particles in regenerated cellulose The novel polymer–inorganic nanocomposite membrane had
membrane as shown in Fig. 3(g). The TiO2 agglomeration resulted been successfully prepared via incorporation of N-doped anatase/
in lower surface area of TiO2 for the absorption of photon to pro- rutile mixed phase TiO2 nanorods into the cellulose microfiber
mote photocatalyst activation. via phase inversion technique. The morphological structure of
Moreover, the higher concentration TiO2 led to the pore block- the prepared membrane was affected by the addition of TiO2
age, which reduced the porosity of the membrane and caused a nanorods. The presence of TiO2 nanorods served to limit the molec-
decrease in contact area between the photocatalyst and pollutants ular motion of the cellulose polymer chains, and spontaneously
in the membrane. As a result, a lower photocatalytic activity of the increased in free volume between the polymer chains. As a result,
RC/TiO2 nanocomposite membrane would be expected. In addition, simultaneous increase in the membrane porosity and mean pore
the photodegradation process occurred inside the pore of the size up to loading of 0.5 wt% was achieved. However, the mem-
nanocomposite membrane throughout the membrane level from brane porosity and mean pore size of the RC/TiO2 nanocomposite
top to bottom, as shown in Fig. 13. It was clearly revealed membrane started to decline when more than 0.5 wt% of TiO2
that the addition of 0.5 wt% of TiO2 in the RC membrane exhibited nanorods concentration was used. The presence of N-doped TiO2
the best TiO2 concentration for the highest photocatalytic activity nanorods within the RCM matrix promoted photoactivation not
under UV and visible light irradiation, respectively. Therefore, the only in UV, but also in visible light irradiation. In addition, the
amount of TiO2 nanorods in the membrane surface needed to understanding and modification on microstructures of photocat-
be optimized in order to obtain the desirable structural and alytic membrane gave a positive impact towards development of
214 M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215

high performance and effective wastewater treatment. Therefore, [17] H. Yano, Optically transparent composites reinforced with networks of
bacterial nanofibers, Adv. Mater. 17 (2005) 153–155.
0.5 wt% of TiO2 nanorods was considered as the best loading to
[18] T. Zhu, Y. Lin, Y. Luo, X. Hu, W. Lin, P. Yu, et al., Preparation and
exhibit appropriate structure and physical properties. RC/TiO2-0.5 characterization of TiO2-regenerated cellulose inorganic–polymer hybrid
exhibited the highest photocatalytic activity in the degradation membranes for dehydration of caprolactam, Carbohydr. Polym. 87 (2012)
of phenol in aqueous solution under UV and visible light irradiation 901–909.
[19] A.W. Morawski, E. Kusiak-Nejman, J. Przepiórski, R. Kordala, J. Pernak,
with percentage of degradation of 96.6% and 78.8%, respectively. Cellulose-TiO2 nanocomposite with enhanced UV–vis light absorption,
The results obtained from this study suggested that the produced Cellulose 20 (2013) 1293–1300.
membrane can be as truly green photocatalytic membrane with [20] A. Snyder, Z. Bo, R. Moon, J.-C. Rochet, L. Stanciu, Reusable photocatalytic
titanium dioxide-cellulose nanofiber films, J. Colloid Interface Sci. 399 (2013)
effective performance in a broader range of light spectrum i.e., 92–98.
from UV to visible light in the wastewater treatment. [21] X.J. Jin, J. Xu, X.F. Wang, Z. Xie, Z. Liu, B. Liang, et al., Flexible TiO2/cellulose
acetate hybrid film as a recyclable photocatalyst, RSC Adv. 4 (2014) 12640–
12648.
Acknowledgements [22] Y. Mao, J. Zhou, J. Cai, L. Zhang, Effects of coagulants on porous structure of
membranes prepared from cellulose in NaOH/urea aqueous solution, J. Membr.
Sci. 279 (2006) 246–255.
The authors acknowledge the financial support from Ministry of [23] M. Ichwan, T. Son, Preparation and characterization of dense cellulose film for
Education Malaysia (Research University Grant Scheme (GUP)) and membrane application, J. Appl. Polym. Sci. 124 (2011) 1409–1418.
Universiti Teknologi Malaysia (UTM) for the research activities [24] S. Mahmoudian, M.U. Wahit, A.F. Ismail, A.A. Yussuf, Preparation of
regenerated cellulose/montmorillonite nanocomposite films via ionic liquids,
undertaken in Advanced Membrane Technology Research Centre
Carbohydr. Polym. 88 (2012) 1251–1257.
(AMTEC). [25] E.K. Liu, W.Q. He, C.R. Yan, ‘‘White revolution” to ‘‘white pollution”—
agricultural plastic film mulch in China, Environ. Res. Lett. 9 (2014)
091001.
Appendix A. Supplementary data [26] B. Choudhury, M. Dey, A. Choudhury, Defect generation, d–d transition, and
band gap reduction in Cu-doped TiO2 nanoparticles, Int. Nano Lett. 3 (2013) 1–
8.
Supplementary data associated with this article can be found, in [27] B. Viswanathan, K.R. Krishanmurthy, Nitrogen incorporation in TiO2: does it
the online version, at http://dx.doi.org/10.1016/j.cej.2015.08.128. make a visible light photo-active material?, Int J. Photoenergy 2012 (2012) 1–
10.
[28] Y.C. Zhang, M. Yang, G. Zhang, D.D. Dionysiou, HNO3-involved one-step low
References temperature solvothermal synthesis of N-doped TiO2 nanocrystals for efficient
photocatalytic reduction of Cr(VI) in water, Appl. Catal. B Environ. 142–143
[1] P.S. Goh, B.C. Ng, W.J. Lau, A.F. Ismail, Inorganic nanomaterials in polymeric (2013) 249–258.
ultrafiltration membranes for water treatment, Sep. Purif. Rev. 216–249 [29] L. Gai, X. Duan, H. Jiang, Q. Mei, G. Zhou, Y. Tian, et al., One-pot synthesis of
(2014). nitrogen-doped TiO2 nanorods with anatase/brookite structures and enhanced
[2] S.-J. You, G.U. Semblante, S.-C. Lu, R.A. Damodar, T.-C. Wei, Evaluation of the photocatalytic activity, CrystEngComm 14 (2012) 7662.
antifouling and photocatalytic properties of poly(vinylidene fluoride) plasma- [30] M.A. Mohamed, W.N.W. Salleh, J. Jaafar, A.F. Ismail, Structural characterization
grafted poly(acrylic acid) membrane with self-assembled TiO2, J. Hazard of N-doped anatase–rutile mixed phase TiO2 nanorods assembled
Mater. 237–238 (2012) 10–19. microspheres synthesized by simple sol–gel method, J. Sol-Gel Sci. Technol.
[3] A. Rahimpour, M. Jahanshahi, B. Rajaeian, M. Rahimnejad, TiO2 entrapped 74 (2015) 513–520.
nano-composite PVDF/SPES membranes: preparation, characterization, [31] M.A. Mohamed, W.N.W. Salleh, J. Jaafar, S.E.A.M. Asri, A.F. Ismail,
antifouling and antibacterial properties, Desalination 278 (2011) 343–353. Physicochemical properties of ‘‘green” nanocrystalline cellulose isolated
[4] S.S. Madaeni, S. Zinadini, V. Vatanpour, A new approach to improve antifouling from recycled newspaper, RSC Adv. 5 (2015) 29842–29849.
property of PVDF membrane using in situ polymerization of PAA [32] N.M. Mokhtar, W.J. Lau, A.F. Ismail, B.C. Ng, Physicochemical study of
functionalized TiO2 nanoparticles, J. Membr. Sci. 380 (2011) 155–162. polyvinylidene fluoride–Cloisite15AÒ composite membranes for membrane
[5] O. Sacco, M. Stoller, V. Vaiano, P. Ciambelli, A. Chianese, D. Sannino, distillation application, RSC Adv. 4 (2014) 63367–63379.
Photocatalytic degradation of organic dyes under visible light on N-doped [33] M.A. Mohamed, W.N.W. Salleh, J. Jaafar, A.F. Ismail, M.A. Mutalib, S.M. Jamil,
photocatalysts, Int. J. Photoenergy 2012 (2012) 1–8. Feasibility of recycled newspaper as cellulose source for regenerated cellulose
[6] R. Mu, Z. Xu, L. Li, Y. Shao, H. Wan, S. Zheng, On the photocatalytic properties of membrane fabrication, J. Appl. Polym. Sci. 42684 (2015).
elongated TiO2 nanoparticles for phenol degradation and Cr(VI) reduction, J. [34] Y. Wei, H.-Q. Chu, B.-Z. Dong, X. Li, S.-J. Xia, Z.-M. Qiang, Effect of TiO2
Hazard Mater. 176 (2010) 495–502. nanowire addition on PVDF ultrafiltration membrane performance,
[7] H. Dzinun, M.H.D. Othman, A.F. Ismail, M.H. Puteh, M.A. Rahman, J. Jaafar, Desalination 272 (2011) 90–97.
Photocatalytic degradation of nonylphenol by immobilized TiO2 in dual layer [35] A. Alpatova, E.-S. Kim, X. Sun, G. Hwang, Y. Liu, M. Gamal El-Din, Fabrication of
hollow fibre membranes, Chem. Eng. J. 269 (2015) 255–261. porous polymeric nanocomposite membranes with enhanced anti-fouling
[8] X. Zhang, W. Chen, Z. Lin, J. Yao, S. Tan, Preparation and photocatalysis properties: effect of casting composition, J. Membr. Sci. 444 (2013) 449–
properties of bacterial cellulose/TiO2 composite membrane doped with rare 460.
earth elements, Synth. React. Inorgan. Met. Nano-Metal Chem. 41 (2011) 997– [36] G. Arthanareeswaran, T.K. Sriyamuna Devi, D. Mohan, Development,
1004. characterization and separation performance of organic–inorganic
[9] G.E. Romanos, C.P. Athanasekou, V. Likodimos, P. Aloupogiannis, P. Falaras, membranes, Sep. Purif. Technol. 67 (2009) 271–281.
Hybrid ultrafiltration/photocatalytic membranes for efficient water treatment, [37] R. Abedini, S.M. Mousavi, R. Aminzadeh, A novel cellulose acetate (CA)
Ind. Eng. Chem. Res. 52 (2013) 13938–13947. membrane using TiO2 nanoparticles: preparation, characterization and
[10] S.I. Patsios, V.C. Sarasidis, A.J. Karabelas, A hybrid photocatalysis–ultrafiltration permeation study, Desalination 277 (2011) 40–45.
continuous process for humic acids degradation, Sep. Purif. Technol. 104 [38] C.S. Ong, W.J. Lau, P.S. Goh, B.C. Ng, A.F. Ismail, Preparation and
(2013) 333–341. characterization of PVDF–PVP–TiO2 composite hollow fiber membranes for
[11] C.P. Athanasekou, G.E. Romanos, F.K. Katsaros, K. Kordatos, V. Likodimos, P. oily wastewater treatment using submerged membrane system, Desalin.
Falaras, Very efficient composite titania membranes in hybrid ultrafiltration/ Water Treat. 53 (2015) 1213–1223.
photocatalysis water treatment processes, J. Membr. Sci. 392–393 (2012) 192– [39] S.Y. Oh, D. Yoo, Y. Shin, H.C. Kim, H.Y. Kim, Y.S. Chung, et al., Crystalline
203. structure analysis of cellulose treated with sodium hydroxide and carbon
[12] S.K. Papageorgiou, F.K. Katsaros, E.P. Favvas, G.E. Romanos, C.P. Athanasekou, dioxide by means of X-ray diffraction and FTIR spectroscopy, Carbohydr. Res.
K.G. Beltsios, et al., Alginate fibers as photocatalyst immobilizing agents 340 (2005) 2376–2391.
applied in hybrid photocatalytic/ultrafiltration water treatment processes, [40] A. Sonia, K. Priya Dasan, Chemical, morphology and thermal evaluation of
Water Res. 46 (2012) 1858–1872. cellulose microfibers obtained from Hibiscus sabdariffa, Carbohydr. Polym. 92
[13] J. Zeng, S. Liu, J. Cai, L. Zhang, TiO2 immobilized in cellulose matrix for (2013) 668–674.
photocatalytic degradation of phenol under weak UV light irradiation, J. Phys. [41] M.K. Mohamad, S.J. Haafiz, A. Eichhorn, M. Hassan Jawaid, Isolation and
Chem. C 114 (2010) 7806–7811. characterization of microcrystalline cellulose from oil palm biomass residue,
[14] H.-Y. Yu, G.-Y. Chen, Y.-B. Wang, J.-M. Yao, A facile one-pot route for preparing Carbohydr. Polym. 93 (2013) 628–634.
cellulose nanocrystal/zinc oxide nanohybrids with high antibacterial and [42] Y. Yue, G. Han, Q. Wu, Transitional properties of cotton fibers from cellulose I
photocatalytic activity, Cellulose 22 (2015) 261–273. to cellulose II structure, Bioresources 8 (2013) 6460–6471.
[15] R.J. Moon, A. Martini, J. Nairn, J. Simonsen, J. Youngblood, Cellulose [43] S.X. Shu, C.R. Li, Fabrication and characterization of regenerated cellulose/TiO2
nanomaterials review: structure, properties and nanocomposites, Chem. Soc. nanocomposite hybrid fibers, Adv. Mater. Res. 418–420 (2011) 237–241.
Rev. 40 (2011) 3941–3994. [44] D. Bondeson, A. Mathew, K. Oksman, Optimization of the isolation of
[16] H. Takagi, Strength properties of cellulose nanofiber green composites, Key nanocrystals from microcrystalline cellulose by acid hydrolysis, Cellulose 13
Eng. Mater. 462–463 (2011) 576–581. (2006) 171–180.
M.A. Mohamed et al. / Chemical Engineering Journal 284 (2016) 202–215 215

[45] A. Kumar, Y.S. Negi, V. Choudhary, N.K. Bhardwaj, Characterization of cellulose [61] J. Gamage, W. McEvoy, Z. Cui Zhang, Degradative and disinfective properties of
nanocrystals produced by acid-hydrolysis from sugarcane bagasse as agro- carbon-doped anatase–rutile TiO2 mixtures under visible light irradiation,
waste, J. Mater. Phys. Chem. 2 (2014) 1–8. Catal. Today 207 (2013) 191–199.
[46] H. Qi, C. Chang, L. Zhang, Properties and applications of biodegradable [62] D.N. Manato, R.N. Prasad, B.K. Mathur, Surface morphological, band and lattice
transparent and photoluminescent cellulose films prepared via a green structural studies of cellulosic fiber coir under mercerization by ESCA, IR and
process, Green Chem. 11 (2009) 177. XRD techniques, Indian J. Pure Appl. Phys. 47 (2009) 643–647.
[47] D.O. Scanlon, C.W. Dunnill, J. Buckeridge, S.A. Shevlin, A.J. Logsdail, S.M. [63] J. Zhang, J. Zhang, L. Lin, T. Chen, J. Zhang, S. Liu, et al., Dissolution of
Woodley, et al., Band alignment of rutile and anatase TiO₂, Nat. Mater. 12 microcrystalline cellulose in phosphoric acid-molecular changes and kinetics,
(2013) 798–801. Molecules 14 (2009) 5027–5041.
[48] M. Soheilmoghaddam, M.U. Wahit, W. Tuck Whye, N. Ibrahim Akos, R. Heidar [64] S. Pourjafar, A. Rahimpour, M. Jahanshahi, Synthesis and characterization of
Pour, A. Ali Yussuf, Bionanocomposites of regenerated cellulose/zeolite PVA/PES thin film composite nanofiltration membrane modified with TiO2
prepared using environmentally benign ionic liquid solvent, Carbohydr. nanoparticles for better performance and surface properties, J. Ind. Eng. Chem.
Polym. 106 (2014) 326–334. 18 (2012) 1398–1405.
[49] C. Schütz, J. Sort, Z. Bacsik, V. Oliynyk, E. Pellicer, A. Fall, et al., Hard and [65] J. Yu, J.C. Yu, W. Ho, Z. Jiang, Effects of calcination temperature on the
transparent films formed by nanocellulose-TiO2 nanoparticle hybrids, PLoS photocatalytic activity and photo-induced super-hydrophilicity of
One 7 (2012) e45828. mesoporous TiO2 thin films, New J. Chem. 26 (2002) 607–613.
[50] X. Zhou, J. Lu, J. Jiang, X. Li, M. Lu, G. Yuan, et al., Simple fabrication of N-doped [66] S. Li, G. Ye, G. Chen, Low-temperature preparation and characterization of
mesoporous TiO2 nanorods with the enhanced visible light photocatalytic nanocrystalline anatase TiO2, J. Phys. Chem. C 113 (2009) 4031–4037.
activity, Nanoscale Res. Lett. 9 (2014) 34. [67] F. Shi, Y. Ma, J. Ma, P. Wang, W. Sun, Preparation and characterization of PVDF/
[51] M. Zhang, D. Lu, G. Yan, J. Wu, J. Yang, Fabrication of Mo+N-codoped TiO2 TiO2 hybrid membranes with different dosage of nano-TiO2, J. Membr. Sci. 389
nanotube arrays by anodization and sputtering for visible light-induced (2012) 522–531.
photoelectrochemical and photocatalytic properties, J. Nanomater. 2013 [68] A. Rahimpour, M. Jahanshahi, A. Mollahosseini, B. Rajaeian, Structural and
(2013) 1–9. performance properties of UV-assisted TiO2 deposited nano-composite PVDF/
[52] D.-H. Wang, L. Jia, X.-L. Wu, L.-Q. Lu, A.-W. Xu, One-step hydrothermal SPES membranes, Desalination 285 (2012) 31–38.
synthesis of N-doped TiO2/C nanocomposites with high visible light [69] G. Hasegawa, K. Kanamori, K. Nakanishi, T. Hanada, Facile preparation of
photocatalytic activity, Nanoscale 4 (2012) 576. hierarchically porous TiO2 monoliths, J. Am. Ceram. Soc. 93 (2010) 3110–3115.
[53] F.N. Sayed, O.D. Jayakumar, R. Sasikala, R.M. Kadam, R.S. Bharadwaj, L. Kienle, [70] J.H. Pan, H. Dou, Z. Xiong, C. Xu, J. Ma, X.S. Zhao, Porous photocatalysts for
et al., Photochemical hydrogen generation using nitrogen-doped TiO2–Pd advanced water purifications, J. Mater. Chem. 20 (2010) 4512.
nanoparticles: facile synthesis and effect of Ti3+ incorporation, J. Phys. Chem. C [71] K. Naeem, F. Ouyang, Influence of supports on photocatalytic degradation of
116 (2012). 12467–12467. phenol and 4-chlorophenol in aqueous suspensions of titanium dioxide, J.
[54] G. Yang, Z. Jiang, H. Shi, T. Xiao, Z. Yan, Preparation of highly visible-light active Environ. Sci. 25 (2013) 399–404.
N-doped TiO2 photocatalyst, J. Mater. Chem. 20 (2010) 5301. [72] J. Xu, F. Wang, W. Liu, W. Cao, Nanocrystalline N-doped TiO2 powders : mild
[55] F. Peng, L. Cai, H. Yu, H. Wang, J. Yang, Synthesis and characterization of hydrothermal synthesis and photocatalytic degradation of phenol under
substitutional and interstitial nitrogen-doped titanium dioxides with visible visible light irradiation, Int. J. Photoenergy 2013 (2013) 1–7.
light photocatalytic activity, J. Solid State Chem. 181 (2008) 130–136. [73] Y. Luo, X. Liu, J. Huang, Heterogeneous nanotubular anatase/rutile titania
[56] X. Yang, C. Cao, L. Erickson, K. Hohn, R. Maghirang, K. Klabunde, Synthesis of composite derived from natural cellulose substance and its photocatalytic
visible-light-active TiO2-based photocatalysts by carbon and nitrogen doping, property, CrystEngComm 15 (2013) 5586.
J. Catal. 260 (2008) 128–133. [74] B. Ohtani, O.O. Prieto-Mahaney, D. Li, R. Abe, What is Degussa (Evonik) P25?
[57] L. Wan, J.F. Li, J.Y. Feng, W. Sun, Z.Q. Mao, Improved optical response and Crystalline composition analysis, reconstruction from isolated pure particles
photocatalysis for N-doped titanium oxide (TiO2) films prepared by oxidation and photocatalytic activity test, J. Photochem. Photobiol. A Chem. 216 (2010)
of TiN, Appl. Surf. Sci. 253 (2007) 4764–4767. 179–182.
[58] M. Zhang, J. Wu, D. Lu, J. Yang, Enhanced visible light photocatalytic activity for [75] K. Yang, Y. Dai, B. Huang, Study of the nitrogen concentration influence on N-
TiO2 nanotube array films by codoping with tungsten and nitrogen, Int. J. doped TiO2 anatase from first-principles calculations, J. Phys. Chem. C 111
Photoenergy 2013 (2013) 1–8. (2007) 12086–12090.
[59] Q. Wang, Z. Jiang, Y. Wang, D. Chen, D. Yang, Photocatalytic properties of [76] M.A. Mohamed, W.N.W. Salleh, J. Jaafar, A.F. Ismail, N.A. Mohamad, Nor,
porous C-doped TiO2 and Ag/C-doped TiO2 nanomaterials by eggshell Photodegradation of phenol by N-Doped TiO2 anatase/rutile nanorods
membrane templating, J. Nanoparticle Res. 11 (2008) 375–384. assembled microsphere under UV and visible light irradiation, Mater. Chem.
[60] F. Dong, S. Guo, H. Wang, X. Li, Z. Wu, Enhancement of the visible light Phys. 162 (2015) 113–123.
photocatalytic activity of C-doped TiO2 nanomaterials prepared by a green
synthetic approach, J. Phys. Chem. C 115 (2011) 13285–13292.

You might also like