You are on page 1of 13

Chemical Engineering Journal 359 (2019) 963–975

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Synthesis and application of stable, reusable TiO2 polymeric composites for T


photocatalytic removal of metronidazole: Removal kinetics and density
functional analysis

N. Neghia, Mathava Kumara, , Danil Burkhalovb
a
Environmental and Water Resources Engineering Division, Department of Civil Engineering, IIT Madras, Chennai 600 036, India
b
Department of Chemistry, Hanyang University, 222 Wangsimni-Ro, Seoul 04763, Republic of Korea

H I GH L IG H T S G R A P H I C A L A B S T R A C T

• Polymeric composites were synthe-


sized and applied for metronidazole
removal.
• cles
PVA-CS-TiO was stable after 15 cy-
2
of UV experiments.
• multi-cycle
No loss of TiO from PVA-CS-TiO in
2
photocatalytic experi-
2

ments.
• Density functional analysis was used
to find metronidazole removal by ad-
sorption.
• Modified Langmuir-Hinshelwood
model was used to estimate rate of
photocatalysis.

A R T I C LE I N FO A B S T R A C T

Keywords: TiO2-polymeric composites, i.e. chitosan supported TiO2 (CS-TiO2) and polyvinyl alcohol/chitosan blend sup-
Chitosan-based TiO2 composites ported TiO2 (PVA-CS-TiO2), were synthesized via precipitation in an alkali/solvent medium and applied for
UV/TiO2 photocatalysis photocatalytic removal of metronidazole (MNZ) in a batch reactor. The composites have produced 100% MNZ
Adsorption removal (10 mg L−1) within 120 min at a catalyst loading of 0.3 g L−1. However, a maximum TOC removal of
Reusability studies
76.1% and 63.7% was observed in PVA-CS-TiO2 and CS-TiO2 systems, respectively. In comparison, complete
DFT modelling
MNZ removal in suspended TiO2 system was observed at 8.3 times higher dosage under similar conditions. The
stability of the composites in water and adsorption mechanism was modelled through density functional theory
(DFT). The energy of formation for CS-TiO2, PVA-TiO2 and PVA-CS in water were found to be −0.144, −0.153
and −0.220 eV, respectively. It was inferred that the presence of CS significantly enhanced MNZ adsorption and
moreover, the removal was due to synergetic removal by adsorption (pseudo-second-order) and photocatalysis
(pseudo-first-order) as evidenced via kinetic models. The hierarchy of adsorption capability of different ad-
sorbents towards MNZ is as follows: PVA-CS-TiO2 > CS-TiO2 > CS > TiO2 > PVA. The prepared composites
showed consistent MNZ removal rates and structural integrity over 15 photocatalytic treatment cycles.
Moreover, a kinetic model was developed to identify a best combination of retention time and TiO2 content in
PVA-CS-TiO2 to predict 100% MNZ. Overall, MNZ removal along with enthalpy values indicated that the pre-
pared composites could be applied for pharmaceutical removal on a long-term basis.


Corresponding author.
E-mail address: mathav@iitm.ac.in (M. Kumar).

https://doi.org/10.1016/j.cej.2018.11.090
Received 27 August 2018; Received in revised form 3 November 2018; Accepted 11 November 2018
Available online 12 November 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

1. Introduction to adsorb metal ions [38,39], dyes [40], inorganic ions [41], organic
compounds such as caffeine, phenols, pharmaceuticals [42–44]. How-
The trace organics removal from secondary effluents has become a ever, instability due to the poor mechanical resistance was observed to
difficult task due to the penetration of many emerging contaminants be the major setback in the application of adsorption. In order to
including pharmaceuticals, personal care products, pesticides, herbi- overcome this limitation, a variety of polymer blends (a mix of two or
cides, etc. to wastewater. Antibiotics usage in human and veterinary more polymers) have been developed which complement each other
medications has increased several folds, which has indirectly in- rather than as individuals. PVA-CS blend is one such combination
troduced their traces in aqueous and soil environments [1,2]. These which is found to have better hydrophilicity, mechanical properties and
compounds can generate antimicrobial resistance (AMR) and eco-toxi- biocompatibility, and it has been applied as (a) an adsorbent for metal
city [3]. Metronidazole (MNZ, C6H9N3O3) is an imidazole antibiotic, removal [45], (b) food packaging material [46], and (c) biomedical and
used to treat anaerobic bacteria, bacteroides and protozoal infections in tissue replacement material [47]. It is also known that PVA is bio-
humans and also been used as feed additives in poultry and aquaculture compatible and non-toxic when administered orally [48]. Therefore,
farms. The presence of MNZ was detected in drinking water, ground/ such polymeric blends are potential candidates for producing highly
surface water, soil sediments and wastewater from one to several functional materials and interfaces. On the other hand, the information
ng L−1, globally [4–7]. Moreover, it has been reported that imidazole- regarding the application of TiO2 polymeric composites in tertiary
resistant genes occur very commonly in Clostridium difficile-positive wastewater treatment is less [19]. Moreover, the characteristics and
patients indicating that MNZ has a very high resistance rate [8]. trace organics removal mechanisms (i.e. both adsorption and photo-
Therefore, removal of MNZ and similar other trace organics from water catalysis) in such composites has not been discussed in detail.
systems is necessary considering its carcinogenicity and the increasing Therefore, the major aim of this investigation is the application of
AMR. CS and PVA-CS blend as support for the preparation of supported
In this context, advanced oxidation systems such as chemical/elec- photocatalysts, i.e. CS-TiO2 and PVA-CS-TiO2 on the light of following
trochemical oxidation [9], Fenton-based [10], ozone-based [11], ra- parameters: (i) aqueous and photo stability of the composites using
diation-based such as UV light [12], visible light [13], E-beam [14], different characterization techniques, (ii) efficiency of the hybrid pho-
gamma [15] and sonolysis [16] processes have been developed and tocatalyst in MNZ removal and its recoverability/reusability, (iii) cal-
investigated for the complete removal of antibiotics. Photocatalysis culate MNZ removal in composite and suspended TiO2 systems using
employing titanium dioxide nanoparticles (TiO2 NPs) possess the a) modified Langmuir-Hinshelwood kinetic model developed, and (iv)
efficiency of mineralization of trace organics [17] by the production of calculate interaction energies between MNZ and different components
hydroxyl radicals (OH.; E0=+2.8V ) and other reactive species as shown of the composites, bonding energies of the composites in water using
in Eq. (1), b) possibility of catalyst reuse and c) applicability in wide pH density functional theory.
range. Anatase TiO2 NPs (101 facet) in water has proven to be an ideal
photocatalyst owing to its crystallinity, surface area for reaction, high 2. Materials and methods
photocatalytic activity, chemical inertness, adsorption of OH/(OH.) and
cost effectiveness [18]. Chitosan (CS, deacetylation: ≥95%, viscosity 200–600 mPa·s), poly
(vinyl alcohol) (PVA, alcoholysis degree: 86–89%,
Trace Organics (R) + hv (UV254 − TiO2) → OH./O2− ./R. → Oxidation Mw ∼ 124,000 g mol−1), methanol and TiO2 (Mean particle size
→ CO2 + H2 O + Intermediates (1) ∼25 nm) were procured from Tokyo chemical industries ltd, Japan, SD
fine chemicals ltd, India and Avra synthesis, India, respectively. All
Intense research on TiO2 has been carried out in two major di- other chemicals used were of analytical grade and used as received.
mensions. One aspect is to shift its applicability from UV light to visible
light region by reduction in the band gap of TiO2 [19,20] and the other 2.1. Synthesis of CS-TiO2 and PVA-CS-TiO2 composites
is development of recoverable TiO2 composites [21] which can facil-
itate construction of large-scale photochemical reactors for air and The CS solution (2%) was obtained by dissolving 2 g of CS powder in
water purification [22]. In the context of developing recoverable pho- 100 mL of dilute acetic acid (2.58% v/v) at 60 °C for 3 h under stirring
tocatalyst, several investigations have focussed on TiO2 coating on conditions. The obtained CS and TiO2 solutions (1% w/v) were
different substrates such as glass sphere [23], quartz tube [24], natural homogenized by ultrasonication for 20 min. Subsequently, the gel blend
zeolite [25], activated carbon [26], concrete surface and also on was cross linked in NaOH (0.5 M) to obtain CS-TiO2 beads. The syn-
membrane surface [27,28] for addressing the recovery issues. It was thesized beads were thoroughly washed twice with DI water in 6 h
observed that coating TiO2 on different surfaces could reduce the cycles, oven-dried at 100 °C for 2 h and used for the degradation stu-
available surface area of TiO2, difficulty in maintaining TiO2 loading dies.
and detachment of TiO2 from substrates. Therefore, embedding TiO2 in For the synthesis of PVA-CS-TiO2 composite, CS solution (2%) and
the support could address the issues of the coating process. Based on the PVA solution (8%, stirred at 80 °C for 5 h) were prepared separately as
support chosen, the synergy of adsorption and photocatalysis greatly reported earlier [49]. The two solutions were blended for 12 h at 80 °C
improves the performance of such photocatalytic composites [29–31]. and at room temperature for another 12 h. Further, the obtained PVA-
Recently, investigations were carried out to embed TiO2 in polymer CS blend was mixed with TiO2 solution (2%) by ultrasonication for
matrices such as polyaniline [32], polyvinyl alcohol [33], chitosan 30 min and the gel blend was introduced as droplets in NaOH-methanol
[34], cellulose [35], polyethylene [36] in order to retain NPs in the solution (0.25 M) under slow stirring conditions. The synthesized beads
matrix and reuse it for multiple photocatalytic treatment cycles. Cor- were washed with DI water, oven-dried for 2 h, stored in sealed con-
respondingly, the present study aims to prepare a hybrid photocatalytic tainers and used for the degradation studies.
material that (a) performs better or equal efficiency as that of TiO2
suspended systems in short as well as long term operation in water and 2.2. Characterization of photocatalytic composites
UV environment, (b) easily recoverable, (c) possess better retention of
TiO2 within the support and (d) biodegradable/non-toxic in the ex- The functionalization of the composites, TiO2, chitosan and PVA
posed environment [37]. In this context, two photocatalytic TiO2 were examined by attenuated total reflectance Fourier-transform in-
composites were developed using chitosan (CS) and polyvinyl alcohol- frared spectroscopy (ATR-FTIR). The FTIR spectra were collected in
chitosan (PVA-CS) polymer blend matrices. CS, the most abundant 4000–550 cm−1 range, with a resolution of 4 cm−1 at room tempera-
natural biopolymer after cellulose, is used in the form of hydrogel beads ture by using a Perkin Elmer spectrometer. Raman spectroscopy was

964
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 1. FTIR spectra of (a) CS, (b) TiO2, (c) CS-TiO2 before treatment (P0), and (d) CS- TiO2 after 15 cycles of photocatalytic treatment (P15).

Fig. 2. FTIR spectra of (a) CS, (b) PVA, (c) TiO2, (d) PVA-CS-TiO2 before treatment (P0), and (e) PVA-CS- TiO2 after 15 cycles of photocatalytic treatment (P15).

performed by Bruker RFS 27 spectrometer provided with Nd:YAG laser emission gun based scanning electron microscopy (FEG SEM, Quanta
(1064 nm) as the excitation source. The Raman spectrum was collected 400) at an accelerating voltage of 15 kV. Moreover, the elemental
in the spectral range of 4000–50 cm−1, with the resolution of 4 cm−1. composition of the samples was quantitatively determined using energy
The powder X-ray diffraction (XRD) patterns were done with Cu Kα dispersive spectroscopy (EDX, Bruker X Flash 6I 10). To further analyse
radiation (λ = 1.5418 Å) at a scan rate of 0.1° min−1 from 10° to 80° the morphological characteristics, high resolution transmission electron
following 40 kV of accelerating voltage and 30 mA of applied current. microscopy (HRTEM with EDX) was performed using JEOL 3010 at an
Prior to the measurement, the dried composite flakes were crushed into accelerating voltage of 200 kV. Surface analysis by XPS was accom-
powder form. The morphology of samples was observed by field plished by irradiating the sample with monoenergetic soft X-rays, Mg

965
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

using Brunauer–Emmett–Teller (BET) method. Prior to the measure-


ment, both the composites were degassed at 150 °C overnight.

2.3. Photocatalysis using composites

Photocatalytic experiments were conducted using CS-TiO2 and PVA-


CS-TiO2 composite flakes (size range of 1.5–3 mm) in a batch reactor as
designed in our previous work [50]. MNZ solution (2L) having the re-
quired concentration (10, 1 and 0.1 mg L−1) was added with 0.3 g L−1
of photocatalyst (CS-TiO2 or PVA-CS-TiO2). The reactor was provided
with 32 W UV power (λmax = 254 nm), continuous stirring (200 rpm)
and operated for 120 min without pH and temperature controls. At
regular intervals of 30 min, samples were collected from the reactor and
analyzed for MNZ concentration. In addition, control experiments were
conducted without the catalyst but with UV (32 W) for 120 min.

2.4. Reusability studies

In order to find the reusability of the composites, photocatalytic


experiments (120 min) were repeated 15 times (P1 to P15) with the
same composite retained (using a metallic tea strainer mesh) after each
cycle. At the end of each cycle, the samples were collected and analyzed
for MNZ concentration. Additionally, blank runs were performed with
Fig. 3. XRD spectra of CS-TiO2 before and after 15 cycles of photocatalytic and without UV illumination at similar catalyst loadings, but in the
treatment. absence of MNZ, to evaluate the stability of the composites in the
photocatalytic reactor.

2.5. Adsorption studies

Adsorption experiments were conducted in a round bottom flask


with 100 mL MNZ solution (10, 1 and 0.1 mg L−1 concentration) and
0.03 g of photocatalyst as adsorbent. The sorbed amount of MNZ per
unit weight of adsorbent (qe, mg g−1) at equilibrium time (t) was cal-
culated using Eq. (2).
V
qe = (C0 − Ce )
W (2)

where C0, Ce, V and W represent initial MNZ concentration (mg L−1),
MNZ concentration at equilibrium (mg L−1), reaction volume (L) and
weight of the adsorbent (g), respectively. Moreover, the data was used
to estimate maximum MNZ adsorption capacity of CS-TiO2 and PVA-CS-
TiO2 using linearized expressions of Langmuir (Eq. (3)), Freundlich (Eq.
(4)), Temkin (Eq. (5)) and Dubinin-Radushkevich (Eq. (6)) isotherms.
1 1 1
+
qe = qmL qmL KL Ce (3)

1
logqe=logKF + logCe
n (4)

RT
qe =Bln (AT ) + Bln (Ce ); B =
bT (5)
Fig. 4. XRD pattern of PVA-CS-TiO2 before and after 15 cycles of photocatalytic
treatment. lnqe = lnqm − (KDR ε 2) (6)

In Eq. (3), qmL is the amount of adsorption corresponding to the com-


Kα (1253.6 eV) using ESCA probe TPD system in an ultra-high vacuum plete monolayer coverage and KL is the Langmuir isotherm constant. In
system. Thermogravimetric analyses (TGA) of the synthesized compo- Freundlich isotherm equation, 1/n and KF are the adsorption intensity
sites, chitosan and PVA were performed in SDT Q 600 TGA (T.A) in- and Freundlich isotherm constant, respectively. In Eq. (5), AT is the
struments under nitrogen atmosphere (100 mL/min) from 25 °C to Temkin isotherm equilibrium binding constant (L mg−1), bT is the
700 °C at a heating rate of 10 °C min−1. A LAMBDA950 UV–vis diffuse Temkin isotherm constant, R is the universal gas constant
reflectance spectrometer was used to measure the light absorption (8.314 J mol−1 K−1), T is the temperature at 298 K and B is the constant
properties of TiO2 and composite materials. BaSO4 was used as the related to heat of sorption (J mol−1). KDR refers to mean free energy of
reference material, and the wavelength range was 250–800 nm. The
adsorption (mol2 kJ−2) and ε = RT ln ⎡1 + C ⎤ is Polyani potential (Eq.
1
specific surface area of the composites was obtained by nitrogen ad- ⎣ e⎦
(6)). MNZ adsorption kinetics on the composites were analyzed by
sorption–desorption isotherm at 77 K using Micromeritics ASAP 2020
plotting Lagergren’s pseudo-first order (Eq. (7)), pseudo-second order
porosimeter. Specific surface area of the composites was evaluated
(Eq. (8)) and intraparticle diffusion (Eq. (9)) models.

966
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 5. TEM images of CS-TiO2 (a) before, and (c) after 15 photocatalytic treatment cycles and PVA-CS-TiO2 (b) before, and (d) after 15 photocatalytic treatment
cycles (Inset shows the lattice spacing (d = 0.36 nm) corresponding to (1 0 1) anatase TiO2).

k1 positions was carried out. During this optimization, the ion cores were
log (qe − qt ) = logqe − t
2.303 (7) described by norm-conserving nonrelativistic pseudo potentials [54]
with cutoff radii 1.14, 1.48, 1.47, 2.42 and 1.25 a.u. for C, N, O, Ti and
t 1 1
= + t H, respectively; the wave functions were expanded with a double-ζ plus
qt K2 qe2 qe (8) polarization basis of localized orbitals for non-hydrogen atoms and with
qt = Kid t 1/2 + c a double-ζ basis for H. Optimizations of the force and total energy were
(9)
performed with the accuracies of 0.04 eV/Å and 1 meV, respectively.
where qe and qt are the adsorption capacities at equilibrium and at time For modeling of CS, three members of CS chains were considered
(t) (mg g−1). k1 is the pseudo-first order rate constant (min−1), k2 is the (Fig. S1a). Similarly, for the modeling of anatase TiO2 and PVA, a slab
pseudo-second order rate constant (g mg−1 min−1), kid is the intra- made from 4 × 2 × 2 TiO2 supercell with surface passivated by hy-
particle diffusion constant (g mg−1 min−0.5) and C is the intercept that droxyl groups (Fig. S1b) and three members of CH2–CH–OH chain (Fig.
were obtained by plotting Eqs. (7)–(9), respectively. S1c) was considered, respectively.
The first step in the modeling is evaluation of the composite stability
2.6. Analytical techniques in water. The second step is, taking into account the stability of com-
posites in water environment; the model evaluates the energy of the
MNZ concentration was analyzed using high performance liquid adsorption of molecules from water. Usually the enthalpy of formation
chromatography (HPLC, Dionex, USA) fitted with C18 column and and/or adsorption is calculated by Eq. (10).
UV–Visible detector. Acetonitrile and deionized water (60:40) was used E form/ads = Ehost + guest − (Ehost + Eguest) (10)
as the mobile phase. The column flow-rate was maintained at
1 mL min−1 and wavelength of UV detector was set at 254 nm. The where Eguest is the total energy of compound in empty space. None-
reaction mixture pH and temperature were measured using a multi- theless, in this case the parts of the composite and the guest molecules
parameter analyzer (OAKTON, India) connected to the probes of the exist before and after reactions in water environment. Thus, Eq. (10)
batch reactor. The total organic carbon remaining in the samples was should be corrected for the case of composites formation as,
analyzed by total organic carbon analyzer equipped with non-destruc-
E form = Ecomposite − (E1part + E2part + …) − (E1water + E2water + …)/n (11)
tive infrared analyzer (Shimadzu, Japan).
where Epart1 is the total energy of the parts of composites in empty
2.7. Modelling using DFT space, Ewater1 is the energy of interaction of selected part of composites
with water, and n is the coefficient which equals 2 when the parts of
Modeling was performed using density functional theory (DFT) and composite participate in the bond formation with guest molecules. The
carried out by means of the pseudo potential code SIESTA [51]. All formation of these new bonds generally originated from one side of the
calculations were performed using the generalized gradient approx- composite (Fig. S1d). It should be also noted that, other part of the
imation (GGA-PBE) including spin polarization [52] with taking into composites generally interacts with water. Thus, n = 1 and the com-
account van der Waals correction [53]. Full optimization of the atomic posites form periodic structure with breaking of all previous

967
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 6. XPS spectra of CS-TiO2 (P15) (a) survey scan, (b) Ti 2p, (c) C 1s, (d) O 1s narrow scans (Inset: XPS spectra of CS-TiO2 before photocatalytic treatment).

interactions of the parts of composite with water. In case of adsorption, MNZ within the boundary conditions [56].
Eq. (12) can be applied.

Eads = Ehost + guest − −(Ehost + Eguest) − −E water (guest) (12) 3. Results and discussion

where Ewater (guest) is the energy of interaction of adsorbed molecules 3.1. Characterization of photocatalytic composites
with water. It is necessary to evaluate the energies of interactions with
water. This is because in an aqueous environment multiple configura- The FT-IR spectra of CS-TiO2 and PVA-CS-TiO2 before (P0) and after
tions of water molecules were possible. Therefore, building of the exact 15 cycles of photocatalytic treatment (P15) were analyzed including the
model for aqueous environment is difficult and expensive process. In pristine materials (CS, PVA, and TiO2) were shown in Figs. 1 and 2. For
this regard, the energy of water have been estimated by summarizing CS, PVA and its composites, the presence of broad peak at 3355 cm−1
the interaction energies of each hydrophilic groups of TiO2 and CS with corresponds to the stretching of OeH and NeH bonds. The peaks at
water (about 0.1 eV) [55]. For MNZ, a correction value of 0.44 eV was 2877 and 2906 cm−1 are due to CeH stretching of the alkyl groups
used for incorporating the energetics of water. Initially, in order to (Figs. 1c and 2d). The peaks seen at 1644 and 1565 cm−1 (as in Figs. 1c,
check the characteristics of the interaction of MNZ with water, total d and 2d) corresponds to CeO stretching in the primary amide group of
energy of interaction between MNZ molecules in solid phase was con- chitosan and bending of NeH of secondary amide [57]. The peak lo-
sidered. Therefore, MNZ stack molecules within periodic boundary cated at 1374 is ascribed to CeH and OeH in plane bending vibration
conditions were built (Fig. S1e). The value of interaction energy be- [58].
tween MNZ molecules was found to be −0.641 eV. The magnitude of Compared with CS, the IR spectrum of CS-TiO2 showed similar
this energy was about one and half time (1 ½) larger than the estimated peaks except the prominent eOeH bond peaks at 3355 and 1374 cm−1.
value of the energy of interaction of MNZ with water. These results are This indicates that intermolecular hydrogen bonding of TiO2 with CS
in agreement with experimentally detected strong hydrophobicity of has occurred. The signature area of chitosan structure indicating

968
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 7. XPS spectra of PVA-CS-TiO2 (P15) (a) survey scan, (b) Ti 2p, (c) C 1s, (d) O 1s narrow scans (Inset: XPS spectra of PVA-CS-TiO2 before photocatalytic
treatment).

Table 1
Performance of TiO2, CS-TiO2 and PVA-CS-TiO2 system in MNZ removal.
Type of photocatalysts Catalyst MNZ removal (%) at different MNZ MNZ removal rate (×10−2 min−1) at different MNZ removal rate (×10−3 min−1) at different
loading concentration (mg L−1) after 120 min MNZ concentration (mg L−1) after 120 min as MNZ concentration (mg L−1) after 120 min as
(g L−1) per Eq. (13) per Eq. (15)

0.1 1 10 0.1 1 10 0.1 1 10

TiO2 0.3 33.8 37.8 17.8 0.43 0.34 0.16 0.1 0.1 0.05
CS-TiO2 0.3 99.9 99.9 98.2 4.6^ 4.1 3.3 1.4^ 1.4 1.1
PVA-CS-TiO2 0.3 99.9 99.9 99.7 4.2^ 4.6 4.8 1.4^ 1.5 1.6
Earlier TiO2 2.5 99.7 98.2 96.2 4.8 3.2 2.7 0.02 0.01 0.01
investigation*

* Results of our earlier investigation [50].


^
MNZ removal rates calculated for 90 min.

saccharide linkages at 1151, 1022 and 897 cm−1 could be observed in vibrations in the spectra of photocatalytic composites (Figs. 1c and 2d)
Fig. 1 [59]. It is observed that the saccharide structure remains un- [60].
disturbed in the synthesis of CS-TiO2 but not in PVA-CS-TiO2. More- In PVA-CS-TiO2 composite, the intensity at 3357 and 2877 (in CS)
over, the absorption peak located at 571 cm−1 corresponds to Ti-O-Ti has reduced and shifted to lower and higher frequency, i.e. at 3342 and

969
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Table 2
MNZ adsorption isotherm parameters calculated for CS-TiO2 and PVA-CS-TiO2 and the pristine materials used for the synthesis of photocatalytic composites.
Isotherm models Parameters TiO2 CS PVA CS-TiO2 PVA-CS-TiO2

−1
Langmuir qm (mg g ) 1.33 2.21 2.11 5.9 5.1
KL (L mg−1) 0.19 0.13 0.09 0.03 0.05
r2 0.9796 0.9947 0.99 0.571 0.3981

Freundlich KF (mg g−1) 0.17 0.22 0.15 0.17 0.24


1 0.8 0.87 0.89 0.9 0.96
n
r2 0.9976 0.9975 0.9986 0.9992 0.9794

Temkin AT (L mg−1) 0.67 0.04 0.68 0.57 0.5


B (J mol−1) 5.48 0.62 4.96 3.15 2.91
r2 0.8642 0.8224 0.8496 0.8339 0.9023

Dubinin-Radushkevich qm (mg g−1) 0.394 0.5 0.38 0.48 0.75


KDR (mol2 KJ−2) 6.0E−08 6.0E−08 7.0E−08 8.0E−08 9.0E−08
E(KJ mol−1) 2887 2887 2672 2500 2357
r2 0.8589 0.8625 0.8511 0.7817 0.9006

Table 3
MNZ kinetic parameters calculated for CS-TiO2 and PVA-CS-TiO2 and the pristine materials used for the synthesis of photocatalytic composites.
Type of photocatalysts MNZ conc (mg L−1) Kinetic models

Pseudo-first-order Pseudo-second-order Intraparticle diffusion

−1 −1 −1 −1 −1
qe (mg g ) k1 (min ) r 2
qe (mg g ) k2 (g mg min ) r 2
kid (g mg−1 min−0.5) r2

TiO2 0.1 0.02 0.05 0.874 0.02 1.43 0.359 0.002 0.895
1 0.13 0.05 0.998 0.05 1.78 0.966 0.017 0.996
10 0.79 0.03 0.998 0.87 0.08 0.772 0.099 0.979

CS 0.1 0.02 0.05 0.861 0.02 0.13 0.328 0.002 0.883


1 0.17 0.1 0.894 0.18 0.35 0.647 0.02 0.862
10 1.43 0.1 0.984 1.06 0.2 0.968 0.14 0.955

PVA 0.1 0.05 0.05 0.994 0.02 6.02 0.878 0.002 0.796
1 0.09 0.03 0.873 0.1 1.72 0.946 0.015 0.769
10 0.96 0.09 0.995 0.84 0.22 0.981 0.11 0.946

CS-TiO2 0.1 0.14 0.24 0.855 0.02 6.4 0.889 0.002 0.976
1 0.24 0.1 0.919 0.1 1.85 0.967 0.015 0.769
10 1.87 0.12 0.952 1.28 0.35 0.993 0.17 0.896

PVA-CS-TiO2 0.1 0.32 0.46 0.834 0.03 13.1 0.986 0.004 0.896
1 0.1 0.58 0.866 0.19 1.35 0.991 0.022 0.839
10 3.26 0.16 0.952 1.45 0.25 0.989 0.12 0.919

2906 cm−1, respectively, indicating that crosslinking of PVA with CS


has occurred [61,62]. On the other hand, the intensity of the peak at
1644 cm−1 has decreased in PVA-CS-TiO2 (Fig. 2d and e) which in-
dicates that the amino group of CS was taking part in the formation of
the composites. In addition, Fig. 2e (i.e. PVA-CS-TiO2 after 15 cycles of
treatment) showed new signals at 1318 and 1420 cm−1 which indicates
CeOeNO2 group [63]. This might be due to the bonding of MNZ with
the composite after treatment. Overall, it is inferred that the structural
integrity of the composites is maintained even after 15 cycles of pho-
tocatalytic treatment.
For further validation of FTIR results, Raman spectroscopy was
performed for the composites for P0 and P15. It is evident that TiO2 was
intact in the anatase form embedded/phase separated in the composites
(Figs. S2 and S3). The characteristic Raman bands of anatase TiO2
present in both the composites around 142, 394, 513 and 639 (cm−1),
well matched out of the six Raman active modes of anatase TiO2 [64].
Moreover, a peak with medium signal at 2898 cm−1 is assigned to CH2
stretching modes, which arise from CS and PVA structure.
The XRD analysis helps to identify the crystallinity of TiO2 present
in the composite for P0 and P15. As illustrated in Figs. 3 and 4 diffrac-
tion peaks clearly show anatase phase structure of TiO2, namely, the
Fig. 8. TGA curves of composites.
planes (1 0 1), (0 0 4), (2 0 0), (1 0 5), (2 1 1), (2 0 4), (1 1 6), (2 2 0) and
(2 1 5) at 2θ (°) values of 25.24, 37.84, 48.03, 53.88, 55.08, 62.57,
68.72, 70.4 and 74.9, respectively. This indicates that all 2θ values are

970
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 9. Removal profiles throughout 15 cycles using CS-TiO2 and PVA-CS-TiO2 to degrade 10 mg L−1 of MNZ.

Table 4
Calculated energies of formation (eV) of the pairs of the various components of
PVA-CS-TiO2 (EComposite), the energy of adsorption of MNZ onto the composite
(EMNZ) and distance (d) of the shortest hydrogen bonds (in Ǻ) between MNZ and
substrate for single adsorbents and the composites.
Substrate EComposite (eV) EMNZ (eV) d

TiO2 – −0.235 2.82


CS – −0.370 2.60
PVA – −0.053 2.88
CS + PVA −0.220 −0.346 2.62
TiO2 + PVA −0.153 0.125 3.31
TiO2 + CS −0.144 −0.013 2.94

in good agreement with (JCPDS-900-8213) [65]. Moreover, XRD pat-


terns obtained for CS-TiO2 and PVA-CS-TiO2 matched with JCPDS-500-
0223 (compared with TiO2 as base). Similarly, XRD patterns of both the
composites retained after 15 cycles of photocatalytic treatment mat-
ched with JCPDS-900-8213. All the standard cards mentioned above
were representing anatase form of TiO2 (Fig. S4) [66]. Therefore, it
proves that TiO2 phase structure is not disturbed in the composites after
treatment.
The SEM images of both the composites show that TiO2 NPs are
distributed on its surface (Figs. S5 and S6). The morphology of PVA-CS-
TiO2 shows smooth surface compared to CS-TiO2. The energy dispersive
X-ray (EDX) spectrum confirms the presence of Ti, O, N and Na in the
synthesized composite apart from C. It is also observed that %Na was
higher in CS-TiO2 (chemically cross linked in alkaline medium) com-
pared to PVA-CS-TiO2 (chemically cross linked in solvent medium (Fig.
S6) and was absent after photocatalytic treatment cycles in both the
composites. Moreover, a slight increase in %N was seen after treatment
indicating that MNZ adsorption onto composites has occurred. On the
other hand, the both flat (spherical shaped) and sharp (hexagonal
shaped) TiO2 NPs can be seen in the TEM images and also it reveals that
the NPs were well dispersed with limited agglomeration in the polymer
matrix and having an average particle size between 50 and 100 nm. The
insets of Fig. 5 illustrates the well-defined fringes with lattice spacing
(d = 0.36 nm) corresponding to the (1 0 1) plane of anatase TiO2. It can
also be noticed that few TiO2 NPs stack over (blue framed portion) in
both the composites. Additionally, EDX analysis reveals that TiO2 (as
%Ti) present on the surface of PVA-CS-TiO2 is higher than on CS-TiO2, Fig. 10. Optimized atomic structures of MNZ adsorbed on CS-TiO2 composite.
which could be attributed to the enhanced binding nature of PVA-CS
blend towards TiO2 [67]. It was also noted that there was loss in TiO2
NPs by 40% in CS-TiO2 but not in PVA-CS-TiO2 after 15 cycles of
treatment (Fig. S7).

971
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Fig. 11. Schematic illustration of the immobilization of TiO2 in CS and PVA-CS matrix by hydrogen bonding.

The valence states of the five elements (C, O, N, Ti, and Na) present 3.2. Photocatalytic studies
in both composites after 15 cycles of photocatalytic treatment (P15)
were analyzed using X-ray photoelectron spectroscopy (XPS) (Figs. 6 MNZ removal in different photocatalytic treatment systems are
and 7). The XPS spectra of CS-TiO2 and PVA-CS-TiO2 before treatment provided in Table 1. At a photocatalytic dosage of 0.3 g L−1 and MNZ
(P0) are provided as insets of Figs. 6 and 7. Accordingly, the binding concentration of 10 mg L−1, 37.8%, 98.2% and 99.7% MNZ removal
energies of Ti 2p levels (Figs. 6b and 7b) present in CS-TiO2 and PVA- was observed in TiO2, CS-TiO2 and PVA-CS-TiO2 systems. Similar MNZ
CS-TiO2 indicate two distinct peaks of Ti (as in TiO2). The C 1s spectra removal pattern was observed at lower MNZ concentrations, i.e. 1 and
of CS-TiO2 indicates the chitosan structure [68]. On comparison of the 0.1 mg L−1. The profile of degradation under different MNZ con-
insets of Figs. 6(c) and 7(c), the peak corresponds to C]O is absent in C centrations and photocatalyst loadings are provided in Fig. S10. A
1s of PVA-CS-TiO2. This indicates that glycosidic bond present in CS has comparison of our earlier photocatalytic degradation experiments with
interacted with the CeH present in PVA during the formation of PVA- TiO2 (2.5 g L−1) and the present study revealed that CS-TiO2 and PVA-
CS blend. The binding energy of O 1s level in both the composites at CS-TiO2 systems have produced almost complete MNZ removal at 8.3
530 eV (Figs. 6(d), 7(d) shows the presence of metal oxide (TieO) on times lesser catalyst loading. The reason being MNZ adsorbed on pho-
the surface of the composite. It was observed that Na 1s (∼1067 eV) tocatalytic composites embedded with TiO2 has the synergistic action of
was absent in both the composites after treatment. The N 1s spectra both adsorption and photocatalytic degradation by the hydroxyl radi-
before and after treatment were provided in Fig. S8. The XPS spectra of cals. The comparison of MNZ removal by adsorption using TiO2, CS-
CS-TiO2 after treatment (P15) has shown the absence of C]O peak in C TiO2 and PVA-CS-TiO2 shows that MNZ adsorption on PVA-CS-TiO2 was
1s (Fig. 6(c)) indicating that saccharide structure of chitosan was dis- the highest and faster (Fig. S10). This could have facilitated the faster
turbed in the presence of UV; whereas PVA-CS-TiO2 after 15 cycles of removal of MNZ in PVA-CS-TiO2. Using the data, MNZ removal rates in
photocatalytic treatment (Fig. 7) revealed no major structural changes all the systems were calculated based on Langmuir-Hinshelwood (L-H)
on the surface of the material. kinetic model as shown in Eq. (13).
The band gap of TiO2, CS-TiO2 and PVA-CS-TiO2 was estimated
from UV–vis-diffuse reflectance spectra (Fig. S9(a)) and the corre- Ct
ln = −k1t
sponding Tauc plot (Fig. S9(b)). Based on Fig. S8, the band gap of TiO2, C0 (13)
CS-TiO2 and PVA-CS-TiO2 was estimated to be 3.16, 3.2 and 3.17 eV,
respectively. Additionally, it was observed that optical absorption of where C0, Ct and k1 are initial MNZ concentration (mg L−1), MNZ
TiO2, CS-TiO2 and PVA-CS-TiO2 falls within the range of 350–400 nm. concentration at 120 min (mg L−1) and pseudo-first-order-rate constant
These observations indicate that the polymeric moieties have no sig- (min−1), respectively, at time (t). The slope of the plot between ln (Ct/
nificant impact on the performance of the photocatalytic oxidation of C0) and t gives the value of k1. Table 1 shows the MNZ removal rates in
MNZ in UV-C range. different systems and it can be noticed that MNZ removal rate was
On the other hand, it was reported that protonated amino groups maximum in PVA-CS-TiO2 (0.048 min−1) compared to CS-TiO2
and amide groups of CS can attract electrons from C and Ti atoms when (0.033 min−1). Moreover, PVA-CS-TiO2 system showed almost steady
they are present in close proximity. Moreover, it was reported that the MNZ removal rates (∼0.046 min−1) at varying concentrations of MNZ.
presence of PVA can increase the conductance of PVA-CS film due to the It can be noticed in Table 1 that complete MNZ removal was
intra- and intermolecular hydrogen bonding (eOH) [69]. Therefore, it achieved in TiO2 slurry systems and also in photocatalytic composite
is anticipated that the hybrid photocatalyst prepared in this work could containing ‘x’% TiO2, for a fixed UV power and irradiation time. It is
have enhanced the charge separation of TiO2 and also improved the important to understand how the rate differs in both the conditions. In
rate of MNZ degradation. order to address this, Eq. (13) is modified as shown in Eq. (14) by in-
corporating concentration of the photocatalyst.

972
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

Ct compared to composites (Table 3). This might be due to fact that the
ln = −kc ·CT ·t
C0 (14) extensive functional groups such as amino and hydroxyl groups present
in the pristine materials were utilized in the formation of composites.
where kc is the modified-pseudo-first-order rate constant and CT is the
concentration of TiO2 or the composite to achieve 100% MNZ de-
gradation for a fixed UV energy power. It is conceived that ‘kc’ would be 3.4. Stability and reusability studies
a better parameter than ‘k1′ to compare the performance of TiO2 which
is either used as nanoparticle form or nanoparticle embedded in an Thermal decomposition pattern of the synthesized composites are
adsorbent matrix. Eq. (14) is similar to the Chick-Watson equation for shown in Fig. 8. In CS-TiO2, the weight loss started at 270 °C and 488 °C;
disinfection. Further, Eq. (14) was modified by incorporating the TiO2 whereas in PVA-CS-TiO2 it occurred at 270 °C, 489 °C and 538 °C. It can
content in the composite as shown in Eq. (15). be observed from Fig. 8, PVA-CS-TiO2 is more thermally stable than CS-
TiO2 due to the chemical crosslinking of CS with PVA. Moreover, it was
Ct
ln = − kc ·CT (Xa + Xb )·t also reported that increase in TiO2 NPs concentration plays a significant
C0 (15) role in the enhancement of tensile strength of PVA-CS [67].
where Xa and Xb are the proportions of adsorbent and the photocatalyst, Additionally, it was observed that both photocatalytic composites
respectively. This equation was applied and the MNZ removal rate was synthesized in this work were stable in neutral and alkaline conditions
calculated for TiO2, CS-TiO2 (contains ∼35% Ti) and PVA-CS-TiO2 for almost 15 days. In case of acidic medium, PVA-CS-TiO2 was found to
(contains ∼ 45% Ti) systems (Table 1). The comparison of MNZ re- be stable in 0.05 M HCl (pH ∼ 2) for a longer time than CS-TiO2 (Table
moval rate as per Eq. (13) for TiO2 slurry system indicate that the value S1). Subsequently, FTIR analysis was performed to check the structural
of rate constant was increased by 1.8 times in PVA-CS-TiO2 system integrity of the composites under the following conditions: (a) com-
(from 0.027 min−1 to 0.048 min−1). On the other hand, the MNZ re- posites stored in airtight glass container for 60 d, (b) composites ob-
moval rate by PVA-CS-TiO2 (0.0016 min−1) was 160 times faster than tained after 15 cycles from the blank runs with UV illumination and
the TiO2 slurry system (0.00001 min−1) as per Eq. (15). These results without MNZ, and (c) composites obtained after 15 cycles from the
indicate that when an adsorbent is chosen as a matrix to embed TiO2, blank runs without UV illumination and MNZ (Figs. S11 and S12). It
the photocatalytic rate is tremendously increased as compared to a can be observed that both the composites showed slight changes in the
slurry system. This is better understood from the modified-pseudo-first- spectra range of 1644 cm−1 which corresponds to the secondary amide
order rate constant (kc). Conversely, Eq. (15) can be used to calculate group of the composites.
the required amount of TiO2 content in the composite to attain 100% The photo stability of the composites in the reactor was evaluated
removal in shorter retention times or vice-versa. by the characterization techniques such as FTIR, FT-Raman, XRD, TEM
MNZ degradation in TiO2 system follows a series of hydroxylation, and XPS results. However, the TEM analysis revealed slight loss in TiO2
denitration steps from the parent compound. The breaking of imidazole after 15 cycles of treatment (Fig. S7(c)) in CS-TiO2. This might be the
moiety leads to small molecular weight carboxylic and alcoholic com- reason for the decrease in MNZ removal (97.9%) at 13th cycle in CS-
pounds and further mineralized to CO2 and H2O [70]. Similar pattern of TiO2; whereas PVA-CS-TiO2 showed complete MNZ removal at varying
degradation is expected in the composite system when MNZ gets en- MNZ concentrations (0.1, 1 and 10 mg L−1) throughout 15 cycles
countered with OH.. The sample collected from CS-TiO2 and PVA-CS- (Fig. 9). Additionally, XPS analysis revealed no structural changes in
TiO2 photocatalytic systems (32 W, 0.3 g L−1 of photocatalyst) was PVA-CS-TiO2 after 30 h of photocatalytic operation.
analyzed for TOC and the value was compared with CS-TiO2 and PVA-
CS-TiO2 alone systems (without UV). The TOC reduction was 63.7% 3.5. Modelling using DFT
and 76.1% in CS-TiO2 and PVA-CS-TiO2 photocatalytic systems, re-
spectively, whereas the reduction was 9.4% and 11.2% in CS-TiO2 and The stability of the photocatalytic composites in water was eval-
PVA-CS-TiO2 alone systems. uated by calculating the formation energy value using Eq. (11) as
shown in Table 4. It should be noted that, the exact combination of the
3.3. Adsorption studies parts of the composite (PVA-CS-TiO2) is unknown. Therefore, all other
possible combinations of the pairs of components to interact with each
The maximum MNZ adsorption capacity of composites including other and with MNZ were considered. From Table 4, the order of sta-
pristine materials was calculated using Eqs. (3)–(6) and the results are bility of the composites in water is found to be PVA-CS > PVA-
shown in Table 2. BET surface area of CS-TiO2 and PVA-CS-TiO2 was TiO2 > CS-TiO2. This observation is in qualitative agreement with
found to be 0.5309 m2 g−1 and 0.1813 m2 g−1, respectively. It is evi- experimental results that are shown in Table S1. The magnitude of the
dent that the specific surface area of chitosan and its composites is very formation energies is moderate, which was few times lower than the
low and the porosity is mainly contributed by TiO2 [71]. Moreover, it energies of coordination bonds ∼ 1 eV. This is because of the number of
was observed that Freundlich isotherm model well represent MNZ ad- groups such as hydroxyl groups on TiO2 surface that participates in the
sorption on CS-TiO2 and PVA-CS-TiO2 (r2 > 0.99). The value of formation of various weak bonds per unit cell. For example, the ten-
Freundlich constant (1/n) was found to be between 0 and 1, which dency to form hydrogen bonds is moderate for CS than TiO2 (Fig. 10).
indicates that MNZ adsorption on all materials is favourable. On the On the other hand, both hydrophilic and hydrophobic groups partici-
other hand, MNZ equilibrium data better fits both Langmuir and pated in the formation of bonds depends on the nature of the compo-
Freundlich adsorption isotherms for pristine materials, i.e. CS, PVA and sites.
TiO2. However, Temkin isotherm gave a reasonable fit of the equili- Secondly, MNZ adsorption on each single part of the composite and
brium data obtained in all the systems, which indicate that there exist on the pairs was evaluated using Eq. (12). The difference in the en-
an electrostatic interaction between MNZ and hydroxyl groups present ergetics of MNZ adsorption was derived from the number and flexibility
in the adsorbents. The E value calculated using Dubinin-Radushkevich of functional groups in the composites participating in the formation of
isotherm (2357–2887 KJ mol−1) reiterates the fact that chemisorption bonds (Fig. 10). It was also observed that the hydrogen bonds between
is responsible for MNZ removal by adsorption. The kinetic study re- different functional groups such as TiO2 passivated with hydroxyl
vealed that MNZ adsorption kinetics in pristine materials has followed groups (Ti-OH) and eNH2 (from CS) with MNZ were found to be
pseudo-first-order kinetics whereas it was pseudo-second-order for CS- stronger than the bonds between the same groups and water. In this
TiO2 and PVA-CS-TiO2 (r2 > 0.98). From the values of k2, it can be context, the calculated distances were in the range of 2.6–2.8 Ǻ in-
inferred that the rate of adsorption was faster in pristine materials when dicating stronger adsorption (hydrogen bonds) while the distances

973
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

greater than 2.8 Ǻ indicates weaker adsorption (dispersion forces). It A.J. Tamhankar, C.S. Lundborg, Antibiotics in wastewater of a rural and an urban
hospital before and after wastewater treatment, and the relationship with antibiotic
has also been found that the adsorption of MNZ on the different com-
use—a one year study from Vietnam, Int. J. Environ. Res. Public Health 13 (2016)
binations of composites is accompanied by synergetic combination of 588.
weak forces. [7] M. Wagil, J. Maszkowska, A. Białk-Bielińska, M. Caban, P. Stepnowski, J. Kumirska,
Determination of metronidazole residues in water, sediment and fish tissue samples,
From DFT results, it is inferred that stability of the composites in
Chemosphere 119 (2015) S28–S34.
water is predominantly due to extensive intermolecular hydrogen [8] J.A. Barkin, D.A. Sussman, N. Fifadara, J.S. Barkin, Clostridium difficile infection
bonds in CS-TiO2. This indicates that there is no compromise on the and patient-specific antimicrobial resistance testing reveals a high metronidazole
intrinsic properties of the components caused by cross-linking or re- resistance rate, Dig. Dis. Sci. 62 (2017) 1035–1042.
[9] A. Dirany, I. Sirés, N. Oturan, A. Özcan, M.A. Oturan, Electrochemical treatment of
inforcements. Moreover, due to the inherent property of anatase TiO2 the antibiotic sulfachloropyridazine: kinetics, reaction pathways, and toxicity
passivation in water matrix [72] surface functionalization was not ne- evolution, Environ. Sci. Technol. 46 (2012) 4074–4082.
cessary for its assembly in the composite. This has facilitated for limited [10] X. Liu, Y. Zhou, J. Zhang, L. Luo, Y. Yang, H. Huang, H. Peng, L. Tang, Y. Mu,
Insight into electro-Fenton and photo-Fenton for the degradation of antibiotics:
agglomeration of TiO2 on the polymer surface as evident from TEM mechanism study and research gaps, Chem. Eng. J. (2018).
analysis. On the basis of the results obtained so far, a schematic illus- [11] R.B. Marcelino, M.M. Leão, R.M. Lago, C.C. Amorim, Multistage ozone and biolo-
tration of attachment of TiO2 with CS and PVA-CS was proposed in gical treatment system for real wastewater containing antibiotics, J. Environ.
Manage. 195 (2017) 110–116.
Fig. 11. [12] M.N. Chong, B. Jin, C.W. Chow, C. Saint, Recent developments in photocatalytic
water treatment technology: a review, Water Res. 44 (2010) 2997–3027.
4. Conclusion [13] A. Truppi, F. Petronella, T. Placido, M. Striccoli, A. Agostiano, M.L. Curri,
R. Comparelli, Visible-light-active TiO2-based hybrid nanocatalysts for environ-
mental applications, Catalysts 7 (2017) 100.
Two photocatalytic polymer composites, i.e. CS-TiO2 and PVA-CS- [14] L. Szabó, J. Szabó, E. Illés, A. Kovács, Á. Belák, C. Mohácsi-Farkas, E. Takács,
TiO2, have been synthesized and compared with regard to their stabi- L. Wojnárovits, Electron beam treatment for tackling the escalating problems of
lity, reusability and efficiency towards the removal of MNZ. Complete antibiotic resistance: eliminating the antimicrobial activity of wastewater matrices
originating from erythromycin, Chem. Eng. J. 321 (2017) 314–324.
MNZ removal was achieved in all systems- TiO2; CS-TiO2 and PVA-CS- [15] P. Attri, F. Tochikubo, J.H. Park, E.H. Choi, K. Koga, M. Shiratani, Impact of Gamma
TiO2 however, it was achieved in the composites at a lower catalyst rays and DBD plasma treatments on wastewater treatment, Sci. Rep. 8 (2018) 2926.
loading. PVA-CS-TiO2 was found to be stable in UV environment and [16] S. Adityosulindro, L. Barthe, K. González-Labrada, U.J.J. Haza, H. Delmas,
C. Julcour, Sonolysis and sono-Fenton oxidation for removal of ibuprofen in (waste)
showed negligible loss of TiO2 NPs even after 15 treatment cycles. The water, Ultrason. Sonochem. 39 (2017) 889–896.
modified rate kinetic model equation was used to compare the removal [17] A. Kumar, M. Khan, L. Fang, I.M. Lo, Visible-light-driven N-TiO2@ SiO2@ Fe3O4
rates of TiO2 slurry systems and the photocatalytic composites. magnetic nanophotocatalysts: synthesis, characterization, and photocatalytic de-
gradation of PPCPs, J. Hazard. Mater. (2017).
Moreover, PVA-CS-TiO2 was comparatively much stable than CS-TiO2 [18] Y.-F. Li, Z.-P. Liu, Particle size, shape and activity for photocatalysis on titania
in all types of aqueous environment (pH 4 to 10) which is attributed to anatase nanoparticles in aqueous surroundings, J. Am. Chem. Soc. 133 (2011)
the enhanced binding nature of PVA-CS towards TiO2. Moreover, the 15743–15752.
[19] L. Liu, M. Yue, J. Lu, J. Hu, Y. Liang, W. Cui, The enrichment of photo-catalysis via
modelling work carried out by density functional theory (DFT) revealed self-assembly perylenetetracarboxylic acid diimide polymer nanostructures in-
that TiO2 passivated with hydroxyl groups assembles to CS or PVA-CS corporating TiO 2 nano-particles, Appl. Surf. Sci. (2018).
through extensive hydrogen bonding and MNZ adsorption to the com- [20] H. Wang, Y. Liang, L. Liu, J. Hu, W. Cui, Highly ordered TiO2 nanotube arrays
wrapped with g-C3N4 nanoparticles for efficient charge separation and increased
posites has occurred through synergistic action of large number of weak
photoelectrocatalytic degradation of phenol, J. Hazard. Mater. 344 (2018)
dispersive forces. The photocatalytic ability of the composites was as- 369–380.
sessed through kinetic models, and the adsorptive ability of the com- [21] Y. Zhang, W. Cui, W. An, L. Liu, Y. Liang, Y. Zhu, Combination of photoelec-
posites was contributed by CS component as evident from the DFT trocatalysis and adsorption for removal of bisphenol A over TiO2-graphene hy-
drogel with 3D network structure, Appl. Catal. B 221 (2018) 36–46.
analysis. Overall, the composites can be used on a long-term basis and [22] C. Byrne, G. Subramanian, S.C. Pillai, Recent advances in photocatalysis for en-
for cost-effective photocatalytic reactor operation for the removal of vironmental applications, J. Environ. Chem. Eng. (2017).
pharmaceuticals. [23] N. Miranda-García, M.I. Maldonado, J. Coronado, S. Malato, Degradation study of
15 emerging contaminants at low concentration by immobilized TiO2 in a pilot
plant, Catal. Today 151 (2010) 107–113.
Acknowledgements [24] K. Natarajan, T.S. Natarajan, H. Bajaj, R.J. Tayade, Photocatalytic reactor based on
UV-LED/TiO 2 coated quartz tube for degradation of dyes, Chem. Eng. J. 178
(2011) 40–49.
The present study is funded by the ICSR, IIT Madras (Grant number: [25] S. Liu, M. Lim, R. Amal, TiO 2-coated natural zeolite: rapid humic acid adsorption
CIE/14-15/832/NFIG/SMAT) and WTI, DST, India (Grant number: and effective photocatalytic regeneration, Chem. Eng. Sci. 105 (2014) 46–52.
DST/TM/WTI/2K15/192). [26] R.C. Asha, M. Vishnuganth, N. Remya, N. Selvaraju, M. Kumar, Livestock waste-
water treatment in batch and continuous photocatalytic systems: performance and
economic analyses, Water Air Soil Pollut. 226 (2015) 132.
Appendix A. Supplementary data [27] X. Zheng, Z.-P. Shen, L. Shi, R. Cheng, D.-H. Yuan, Photocatalytic Membrane
Reactors (PMRs) in water treatment: configurations and influencing factors,
Catalysts 7 (2017) 224.
Supplementary data to this article can be found online at https://
[28] S. Leong, A. Razmjou, K. Wang, K. Hapgood, X. Zhang, H. Wang, TiO2 based
doi.org/10.1016/j.cej.2018.11.090. photocatalytic membranes: a review, J. Membr. Sci. 472 (2014) 167–184.
[29] C. Mu, Y. Zhang, W. Cui, Y. Liang, Y. Zhu, Removal of bisphenol A over a separation
References free 3D Ag3PO4-graphene hydrogel via an adsorption-photocatalysis synergy, Appl.
Catal. B 212 (2017) 41–49.
[30] X. Wang, Y. Liang, W. An, J. Hu, Y. Zhu, W. Cui, Removal of chromium (VI) by a
[1] J.M. Philip, U.K. Aravind, C.T. Aravindakumar, Emerging contaminants in Indian self-regenerating and metal free g-C3N4/graphene hydrogel system via the synergy
environmental matrices–A review, Chemosphere 190 (2018) 307–326. of adsorption and photo-catalysis under visible light, Appl. Catal. B 219 (2017)
[2] J. Wilkinson, P.S. Hooda, J. Barker, S. Barton, J. Swinden, Occurrence, fate and 53–62.
transformation of emerging contaminants in water: an overarching review of the [31] Y. Li, W. Cui, L. Liu, R. Zong, W. Yao, Y. Liang, Y. Zhu, Removal of Cr (VI) by 3D
field, Environ. Pollut. 231 (2017) 954–970. TiO2-graphene hydrogel via adsorption enriched with photocatalytic reduction,
[3] A.J. Ebele, M.A.-E. Abdallah, S. Harrad, Pharmaceuticals and personal care pro- Appl. Catal. B 199 (2016) 412–423.
ducts (PPCPs) in the freshwater aquatic environment, Emerg. Contaminants (2017). [32] W. Cui, J. He, H. Wang, J. Hu, L. Liu, Y. Liang, Polyaniline hybridization promotes
[4] H.W. Leung, L. Jin, S. Wei, M.M.P. Tsui, B. Zhou, L. Jiao, P.C. Cheung, Y.K. Chun, photo-electro-catalytic removal of organic contaminants over 3D network structure
M.B. Murphy, P.K.S. Lam, Pharmaceuticals in tap water: human health risk as- of rGH-PANI/TiO2 hydrogel, Appl. Catal. B 232 (2018) 232–245.
sessment and proposed monitoring framework in China, Environ. Health Perspect. [33] P. Lei, F. Wang, X. Gao, Y. Ding, S. Zhang, J. Zhao, S. Liu, M. Yang, Immobilization
121 (2013) 839. of TiO2 nanoparticles in polymeric substrates by chemical bonding for multi-cycle
[5] Y.-C. Lin, W.W.-P. Lai, H.-H. Tung, A.Y.-C. Lin, Occurrence of pharmaceuticals, photodegradation of organic pollutants, J. Hazard. Mater. 227 (2012) 185–194.
hormones, and perfluorinated compounds in groundwater in Taiwan, Environ. [34] Y. Zhao, C. Tao, G. Xiao, H. Su, Controlled synthesis and wastewater treatment of
Monit. Assess. 187 (2015) 256. Ag 2 O/TiO 2 modified chitosan-based photocatalytic film, RSC Adv. 7 (2017)
[6] L.T.Q. Lien, N.Q. Hoa, N.T.K. Chuc, N.T.M. Thoa, H.D. Phuc, V. Diwan, N.T. Dat, 11211–11221.

974
N. Neghi et al. Chemical Engineering Journal 359 (2019) 963–975

[35] U.M. Garusinghe, V.S. Raghuwanshi, W. Batchelor, G. Garnier, Water resistant [56] J. Méndez-Díaz, G. Prados-Joya, J. Rivera-Utrilla, R. Leyva-Ramos, M. Sánchez-
cellulose-titanium dioxide composites for photocatalysis, Sci. Rep. 8 (2018) 2306. Polo, M. Ferro-García, N. Medellín-Castillo, Kinetic study of the adsorption of ni-
[36] F. Magalhães, F.C. Moura, R.M. Lago, TiO2/LDPE composites: a new floating troimidazole antibiotics on activated carbons in aqueous phase, J. Colloid Interface
photocatalyst for solar degradation of organic contaminants, Desalination 276 Sci. 345 (2010) 481–490.
(2011) 266–271. [57] P.B. Palani, K.S. Abidin, R. Kannan, M. Sivakumar, F.-M. Wang, S. Rajashabala,
[37] I. Armentano, D. Puglia, F. Luzi, C.R. Arciola, F. Morena, S. Martino, L. Torre, G. Velraj, Improvement of proton conductivity in nanocomposite polyvinyl alcohol
Nanocomposites based on biodegradable polymers, Materials 11 (2018) 795. (PVA)/chitosan (CS) blend membranes, RSC Adv. 4 (2014) 61781–61789.
[38] S. Pu, H. Ma, A. Zinchenko, W. Chu, Novel highly porous magnetic hydrogel beads [58] A. Zając, J. Hanuza, M. Wandas, L. Dymińska, Determination of N-acetylation de-
composed of chitosan and sodium citrate: an effective adsorbent for the removal of gree in chitosan using Raman spectroscopy, Spectrochim. Acta Part A Mol. Biomol.
heavy metals from aqueous solutions, Environ. Sci. Pollut. Res. (2017) 1–11. Spectrosc. 134 (2015) 114–120.
[39] M. Barakat, New trends in removing heavy metals from industrial wastewater, [59] M.Y. Chan, S. Husseinsyah, S.T. Sam, Chitosan/corn cob biocomposite films by
Arabian J. Chem. 4 (2011) 361–377. cross-linking with glutaraldehyde, BioResources 8 (2013) 2910–2923.
[40] S. Chatterjee, D.S. Lee, M.W. Lee, S.H. Woo, Congo red adsorption from aqueous [60] Y. Zhao, Y. Dong, F. Lu, C. Ju, L. Liu, J. Zhang, B. Zhang, Y. Feng, Coordinative
solutions by using chitosan hydrogel beads impregnated with nonionic or anionic integration of a metal-porphyrinic framework and TiO 2 nanoparticles for the for-
surfactant, Bioresour. Technol. 100 (2009) 3862–3868. mation of composite photocatalysts with enhanced visible-light-driven photo-
[41] S. Chatterjee, S.H. Woo, The removal of nitrate from aqueous solutions by chitosan catalytic activities, J. Mater. Chem. A 5 (2017) 15380–15389.
hydrogel beads, J. Hazard. Mater. 164 (2009) 1012–1018. [61] P. Sharma, G. Mathur, S.R. Dhakate, S. Chand, N. Goswami, S.K. Sharma,
[42] A. Zarzar, M. Hong, B. Llanos, A. Navarro, Insights into the eco-friendly adsorption A. Mathur, Evaluation of physicochemical and biological properties of chitosan/
of caffeine from contaminated solutions by using hydrogel beads, J. Environ. Anal. poly (vinyl alcohol) polymer blend membranes and their correlation for Vero cell
Chem 2 (2015). growth, Carbohydr. Polym. 137 (2016) 576–583.
[43] U. Soni, J. Bajpai, S.K. Singh, A. Bajpai, Evaluation of chitosan-carbon based bio- [62] M. Li, S. Cheng, H. Yan, Preparation of crosslinked chitosan/poly (vinyl alcohol)
composite for efficient removal of phenols from aqueous solutions, J. Water Process blend beads with high mechanical strength, Green Chem. 9 (2007) 894–898.
Eng. 16 (2017) 56–63. [63] P. Larkin, Infrared and Raman Spectroscopy: Principles and Spectral Interpretation,
[44] G.Z. Kyzas, D.N. Bikiaris, Recent modifications of chitosan for adsorption applica- Elsevier, 2017.
tions: a critical and systematic review, Mar. Drugs 13 (2015) 312–337. [64] O. Frank, M. Zukalova, B. Laskova, J. Kürti, J. Koltai, L. Kavan, Raman spectra of
[45] X. Li, Y. Li, Z. Ye, Preparation of macroporous bead adsorbents based on poly (vinyl titanium dioxide (anatase, rutile) with identified oxygen isotopes (16, 17, 18), Phys.
alcohol)/chitosan and their adsorption properties for heavy metals from aqueous Chem. Chem. Phys. 14 (2012) 14567–14572.
solution, Chem. Eng. J. 178 (2011) 60–68. [65] N. Khalid, E. Ahmed, Z. Hong, L. Sana, M. Ahmed, Enhanced photocatalytic activity
[46] S. Tripathi, G. Mehrotra, P. Dutta, Physicochemical and bioactivity of cross-linked of graphene–TiO2 composite under visible light irradiation, Curr. Appl Phys. 13
chitosan–PVA film for food packaging applications, Int. J. Biol. Macromol. 45 (2013) 659–663.
(2009) 372–376. [66] A. Altomare, N. Corriero, C. Cuocci, A. Falcicchio, A. Moliterni, R. Rizzi, QUALX2.
[47] S. Rinehart, T. Campbell, K. Burke, B. Garcia, A. Mlynarski, Synthesis and char- 0: a qualitative phase analysis software using the freely available database
acterization of a Chitosan/PVA antimicrobial hydrogel nanocomposite for re- POW_COD, J. Appl. Crystallogr. 48 (2015) 598–603.
sponsive wound management materials, J. Microb. Biochem. Technol. 8 (2016) [67] A.M. Youssef, S.M. El-Sayed, H.H. Salama, H.S. El-Sayed, A. Dufresne, Evaluation of
065–070. bionanocomposites as packaging material on properties of soft white cheese during
[48] M. Teodorescu, M. Bercea, S. Morariu, Biomaterials of Poly (vinyl alcohol) and storage period, Carbohydr. Polym. 132 (2015) 274–285.
natural polymers, Polym. Rev. 58 (2018) 247–287. [68] A. Al-Mokaram, M.A. Amir, R. Yahya, M.M. Abdi, H.N.M.E. Mahmud, The devel-
[49] L. Jin, R. Bai, Mechanisms of lead adsorption on chitosan/PVA hydrogel beads, opment of non-enzymatic glucose biosensors based on electrochemically prepared
Langmuir 18 (2002) 9765–9770. polypyrrole–chitosan–titanium dioxide nanocomposite films, Nanomaterials 7
[50] N. Neghi, M. Kumar, Performance analysis of photolytic, photocatalytic, and ad- (2017) 129.
sorption systems in the degradation of metronidazole on the perspective of removal [69] A. Islam, T. Yasin, M.J. Akhtar, Z. Imran, A. Sabir, M. Sultan, S.M. Khan, T. Jamil,
rate and energy consumption, Water Air Soil Pollut. 228 (2017) 339. Impedance spectroscopy of chitosan/poly (vinyl alcohol) films, J. Solid State
[51] J.M. Soler, E. Artacho, J.D. Gale, A. García, J. Junquera, P. Ordejón, D. Sánchez- Electrochem. 20 (2016) 571–578.
Portal, The SIESTA method for ab initio order-N materials simulation, J. Phys.: [70] N. Neghi, R.N. Krishnan, M. Kumar, Analysis of metronidazole removal and micro-
Condens. Matter 14 (2002) 2745. toxicity in photolytic systems: effects of persulfate dosage, anions and reactor op-
[52] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made eration-mode, J. Environ. Chem. Eng. (2018).
simple, Phys. Rev. Lett. 77 (1996) 3865. [71] M.H. Farzana, S. Meenakshi, Synergistic effect of chitosan and titanium dioxide on
[53] G. Román-Pérez, J.M. Soler, Efficient implementation of a van der Waals density the removal of toxic dyes by the photodegradation technique, Ind. Eng. Chem. Res.
functional: application to double-wall carbon nanotubes, Phys. Rev. Lett. 103 53 (2013) 55–63.
(2009) 096102. [72] C. Arrouvel, M. Digne, M. Breysse, H. Toulhoat, P. Raybaud, Effects of morphology
[54] N. Troullier, J.L. Martins, Efficient pseudopotentials for plane-wave calculations, on surface hydroxyl concentration: a DFT comparison of anatase–TiO2 and γ-alu-
Phys. Rev. B 43 (1991) 1993. mina catalytic supports, J. Catal. 222 (2004) 152–166.
[55] J.N. Israelachvili, Intermolecular and Surface Forces, Academic press, 2011.

975

You might also like