Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Applied Surface Science 449 (2018) 371–379

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Photocatalytic degradation of 2,4-dichlorophenoxyacetic acid - A


comparative study in hydrothermal TiO2 and commercial TiO2
Sandeep S a , K.L. Nagashree b , T. Maiyalagan c , G. Keerthiga a,∗
a
Department of Chemical Engineering, BMS College of Engineering, Bangalore, India
b
Department of Chemistry, BMS College of Engineering, Bangalore, India
c
Department of Chemistry, SRM research Institute, SRM University, Kattankulathur, India

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

Article history: Photoremediation of pesticides under natural sunlight will not be distant a dream any more due to rapid
Received 29 October 2017 development in the field of catalysis and its related technology. 2,4-dichlorophenoxyacetic acid (2,4-
Received in revised form 28 January 2018 D) is a common broad-leaved weeds, found in surface and groundwater at concentrations above the
Accepted 5 February 2018
maximum containment level, is chosen as model pollutant in this study. Hydrothermal synthesis of TiO2
Available online 13 February 2018
under sol-gel route resulted in altered band gap material with good crystallinity and particle size. Further,
the overview of its degradation is compared with commercial TiO2 such as P25, Hombicat UV-100 and
Keywords:
commercial brand of TiO2 (C-TiO2 ). More importantly, commercial viability of the catalyst was assessed
2,4-D degradation
Hydrothermal TiO2
in natural sunlight for its photocatalytic activity and selectivity. The chosen catalyst was characterized
P25 using XRD, SEM, EDAX, Raman spectroscopy, FTIR and DRS-UV for its structural, functional and electronic
Hombicat properties. Hydrothermal catalyst (H-TiO2 ) showed degradation of 96% under UV light and 83% under
UV light solar light due to its altered band gap (3.16 eV), selective anatase phase ratio appending to similar particle
Solar light size and crystallinity as that of P25. The major intermediates of 2,4-D degradation were identified for H-
TiO2 and its selective degradation mechanism was highlighted. Thus, this work not only puts forward
the importance of material characteristics for mitigation of model pollutant but also its feasibility for
practical implication under natural sunlight.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction ization and scale-up. Titanium dioxide (TiO2 ), as a photocatalysts


is promising due to its merits of strong photo-oxidizing power,
Intensification of agricultural and industrial activities to suit chemical stability, non-toxicity and low cost. Unmodified TiO2
global demand culminates in contaminating air, water, soil and suffers from narrow light-response range, difficult catalyst recov-
aquatic ecosystems [1]. Petroleum, petrochemical, food and fer- ery with low efficiency and fast recombination [5]. However,
tilizers based industries discharge waste waters characterized by it exhibits unique structural, thermal and electronic properties
a perceptible content of organics and hence gains major scientific based on the preparation methods where the size, shape, crystal
interest [2]. Increasing public concerns for water reclamation and structure and phases are tuned for specific applications. Impro-
reuse, imposes stringent regulation on wastewater disposal and vising the characteristics of TiO2 by affecting its synthesis route
mandates the need of efficient treatment technologies. The recent by doping, impregnation and band gap engineering have evolved
research on solar photocatalysis is gaining interest as it uses avail- nanostructured TiO2 for dye degradation, pesticide redemption,
able solar energy for redemption of pollutants [3]. micropollutants mitigation and for solar cells [6–8]. Amongst
Among the different strategies developed, heterogeneous pho- the different routes, hydrothermal method offers advantages for
tocatalysis has gained attention for its complete mineralization engineering the material crystallinity by controlling the solution
under ambient temperature and pressure [4]. Attentions to envi- temperature, pressure and composition [9]. Nevertheless, better
ronmental application by photocatalysts gather interest due to understanding of TiO2 for its large scale commercialization, catalyst
captivating properties, such as enhanced reactivity, commercial- recovery, and increased efficiency are still unanswered.
Interest towards degradation of 2,4-dichlorophenoxyacetic acid
(2,4-D), a plant hormone regulator, is due to its existence as an
active constituent in numerous widely used herbicides. Few stud-
∗ Corresponding author.
ies report its carcinogenic behavior and deliberated to be harmful
E-mail address: gkeerthiga.che@bmsce.ac.in (G. Keerthiga).

https://doi.org/10.1016/j.apsusc.2018.02.051
0169-4332/© 2018 Elsevier B.V. All rights reserved.
372 S. S et al. / Applied Surface Science 449 (2018) 371–379

light, superior accessibility of treatment reagent and for its feasi-


ble operating conditions. Reports on TiO2 include novel magnetic
imprinted TiO2 [13], novel TiO2 photocatalyst based on the mag-
netic floating fly-ash cenospheres [14], nitrogen doped TiO2 [15],
tourmaline coated TiO2 [5], TiO2 -platinum composite [16], P25
doped graphene [17] have been previously studied for optimiz-
ing the photocatalytic degradation of pesticides such as 2,4-D and
micropollutants. However, studies comparing the degradation effi-
ciency of market available catalyst such as P25 Degussa, Hombicat
UV-100, commercial grade of TiO2 and hydrothermally prepared
TiO2 for 2,4-D degradation are scare.
Degussa P25 is a commercially available standard with higher
anatase to rutile ratio with specific surface area and particle size.
Conversely, Hombicat UV 100 having pure anatase phase and
improved surface area finds applications as commercial catalyst.
Likewise, a locally marketed commercial TiO2 (henceforth abbrevi-
ated as C-TiO2 ) was included for testing 2,4-D degradation. Though
these commercial catalyst have been studied for photocatalytic
degradation of dyes, pesticides and treatment of volatile organic
compounds [18], the behavior of the catalyst for photocatalytic
Fig. 1. Pictorial representation of Photocatalytic reactor. degradation of 2,4-dichlorophenoxy acetic acid has not been explic-
itly reported in literature.
Hence, the objective of the present work is to compare the cat-
alytic activities of different commercial TiO2 and hydrothermally
prepared TiO2 towards 2,4-D degradation. Three commercial sam-
ples of TiO2 such as P25, Hombicat (UV −100), and commercial
grade of TiO2 (C-TiO2 ) are chosen along with TiO2 prepared by facile
hydrothermal enriched sol-gel route. The photocatalytic degrada-
tion of 2,4-D under UV and visible solar light is studied on all four
above mentioned catalysts. The kinetics, effect of aeration, photo
stability of the catalyst and degradation mechanism are attempted
for a better insight on 2,4-D redemption.

2. Materials and methods

2.1. Materials

Titanium tetra iso-propoxide (TTIP) (99.98%), 2,4-


dichlorophenoxyacetic acid (99%) used in these experiments
were from Sigma Aldrich. Analytical grade chemicals and solvents
were used without any further purification and double distilled
water was used throughout the experiments.

Fig. 2. XRD of different photocatalyst (* represent Rutile phase). 2.2. Synthesis of TiO2 nanomaterial

Table 1 The hydrothermal TiO2 (H-TiO2 ) was synthesized using the


Crystal size and phase ratio of the catalyst. facile hydrothermal enriched sol-gel route by stirring titanium tetra
iso-propoxide (1 M) and acetic acid (4 M) to attain homogeneity
Catalyst Crystal size (nm) WA (%) WR (%)
with drop wise addition of water followed by aging (24 h) [9]. Sub-
P25 23.7 86 16
sequently, hydrothermal reaction was carried out for 18 h at 60 ◦ C
Hombicat 6.0 100 0
C-TiO2 63.6 100 0 by transferring the contents to Teflon coated stainless steel auto-
H-TiO2 21.6 100 0 clave. The resulting mixture was centrifuged, washed and dried at
60 ◦ C followed by calcination at 600 ◦ C for 1 h.

on humans and animals [10]. Residues of 2,4-D after its usage in 2.3. Characterization
farmland or by improper disposal, contaminate and endanger water
bodies. Biodegradation of 2,4-D in water is very slow (half-life XRD patterns of the commercial and hydrothermal samples
6–170 days) mandating the intentional exclusion in water bod- were obtained with high resolution X-ray diffractometer (PAN-
ies [11]. Hence, the interest towards the redemption of 2,4-D by alytical X’Pert PRO), using Cu K␣ radiation at 45 kV and 20 mA.
◦ ◦
advances in material and technology needs attention. The scan rate was 0.03 min−1 in the range 2␪ = 10–90 . The mor-
Mitigation by adsorption, biodegradation [11] and photocat- phology of as-synthesized and commercial catalyst were obtained
alytic degradation [12] has been investigated in literature, but the using a SEM (TESCAN-VEGA3), equipped with an Energy Dispersive
commercialization is limited by its scale up, cost and efficiency Spectroscopy and High Resolution Analytical 200 KV Transmis-
attained within the time frame in each process. On the other hand, sion Electron Microscope (JEOL 2100). Raman spectrum was
Advanced Oxidation Process (AOP) shows promising venture for recorded (RAM-PKP-785) in the range of 700–200 cm−1 with a
pollutants redemption as it assists in scale up under natural sun spectral resolution of 2 cm−1 . The FT-IR spectrum was recorded
S. S et al. / Applied Surface Science 449 (2018) 371–379 373

Fig. 3. (a) SEM images of various catalyst. (b) TEM images of P25 and H-TiO2.
374 S. S et al. / Applied Surface Science 449 (2018) 371–379

(PerkinElmer-L160000A) in the range of 4000–400 cm−1 . The


energy band gap was determined through diffuse reflectance spec-
tra of UV–Vis (Jasco) spectrophotometer using Kubelka-Munk
theory. Carbon content before and after degradation was ana-
lyzed using TOC analyzer. The intermediates of 2,4-D photocatalytic
degradation were analyzed by quadruple liquid chromatography
with mass spectrophotometer (LCMS-Agilent) of appropriate ion-
ization source.

2.4. Experimental setup

Kinetic experiments on photocatalytic degradation of 2,4-D


(50 ppm) was carried out using a glass fabricated double jack-
eted photo-reactor (1000 mL) with a UV-light (Philips ∼265 nm,
11 W) as pictorially represented in Fig. 1. Prior to photocatalytic
experiments, the adsorption experiment was carried out with mag-
Fig. 4. Raman spectra of various photocatalyst.
netic stirring in the dark with subsequent UV spectrophotometer
analysis for every 15 min and the rate of adsorption was calcu-
lated [8]. Water circulation was provided in the outer jacket where
an aeration of 2.5 L/min was maintained for all experiments. The
degradation of 2,4-D under solar visible light was carried out on
sunny days from 10 am to 3 pm under similar conditions. Uniform
catalyst loading of 0.2 g/L was maintained in all experiments where
degradation of 2,4-D (50 ppm) was quantified by using UV–vis
spectrophotometer (Jasco-V530) for the chromophore observed at
232 nm. All reactions were carried out at room temperature and the
samples analyzed at regular intervals of time. To check the stability
of the catalyst, the catalyst at the end of the reaction was recovered,
washed and dried followed by its reuse in subsequent runs as indi-
cated by Lu et al., [19]. The calibration curve was plotted at 232 nm
to find the minimum detectable limit of unconverted 2, 4-D (Fig.
A1).

3. Results and discussions Fig. 5. FTIR spectra of different photocatalyst.

The morphological and physiological properties of the catalyst, Fig. 3b shows HRTEM image obtained for P25 and hydrothermal
their correlation with the photocatalytic activity are discussed in catalyst where the catalyst showed uniform hierarchical spherical
the following sections. structure with good dispersibility. The SAED (Fig. 3b iii) patterns
with rings assigned to corresponding hkl values for P25 and H-TiO2
3.1. XRD analysis are represented in the plot.

Fig. 2 shows the XRD patterns of commercial catalyst and com- 3.3. Raman spectroscopy
pared with hydrothermal TiO2 . The major peaks obtained at 2␪
positions of 25.33◦ , 38.56◦ , 48.06◦ , 53.89◦ can be ascribed to the Raman spectroscopy was used to find the information required
(101), (004), (105), (211) anatase planes of hydrothermally synthe- for TiO2 structural fingerprint via inelastic scattering principle [21].
sized TiO2 respectively, which is in agreement with the literature Fig. 4 shows the Raman spectra of the commercial and synthe-
[9]. The phase ratio and crystal size determined from the XRD data sized catalysts, where the Eg band corresponds to the symmetric
are shown in Table 1. Complete anatase phase (100%) was found for stretching vibration of O Ti O while the B1g band corresponds to
hydrothermal TiO2 , Hombicat and commercial grade of TiO2 except the symmetric bending vibration of O Ti O. Furthermore, the A1g
P25 with 84% anatase and 16% rutile. The crystallographic particle band position corresponds to the anti-symmetric bending vibration
size by hydrothermal method was 21 nm comparable to 23.7 nm of O Ti O [22]. The varying ratio of these bands from commer-
for P25, 6 nm for Hombicat and 63 nm for commercial TiO2 as cal- cial to hydrothermal TiO2 shows increasing degree of interaction.
culated from the Scherer’s formula. The particle size of commercial Sharp peaks in hydrothermal catalyst at ≈202 cm−1 and 645 cm−1 ,
catalyst was in comparison with literature [20] whereas range of shoulder peaks at 397 cm−1 corresponds to the Eg and B1g planar
particle size obtained for hydrothermally synthesized TiO2 and P25 vibrations of TiO2 [16].
are similar.
3.4. FTIR
3.2. SEM/TEM
Fig. 5 shows FTIR spectra of the commercial and as-synthesized
Fig. 3a depicts the comparative SEM images of the commercial catalysts. The bands at 3437 cm−1 , 2921 cm−1 , 1726 cm−1 ,
TiO2 compared with hydrothermal TiO2 . Spherical morphology has 1627 cm−1 , and 1258 cm−1 can be ascribed to O H, C H, C O, C C,
been reflected in all samples with the hydrothermally prepared and Anatase O Ti O stretching vibrations respectively. The band
TiO2 samples showing a mixture of aggregates than P25 and C-TiO2 . at 664 cm−1 is due to the Ti O vibration modes of the Ti3+ O Ti4+
The EDAX collected at representative sites for all samples showed framework [23]. The observations are in coherence with the results
enriched Ti and O composition with minimal carbon impurity. of XRD where no rutile peaks were observed.
S. S et al. / Applied Surface Science 449 (2018) 371–379 375

Fig. 6. (a) Absorbance spectra of the photocatalyst. (b) KM plot of photocatalyst.

Fig. 7. (a) Adsorption capacities of different samples. (b) UV degradation plot of 2,4-D. (c) Kinetic plot of 2,4-D degradation under UV light.

Table 2
Comparison of UV and solar rate constant of the TiO2 catalysts.

Catalyst UV degradation Solar degradation


−1 −2
Rate constant (k) min 10 R 2
Rate constant (k) min−1 10−2 R2

P25 1.00 0.94 0.52 0.91


Hombicat 0.59 0.85 0.49 0.98
C-TiO2 0.98 0.94 0.24 0.93
H-TiO2 1.97 0.98 0.70 0.98
376 S. S et al. / Applied Surface Science 449 (2018) 371–379

rium within 30 minutes of adsorption. It is inferred that 2,4-D is


predicted to be adsorbed, desorbed and again adsorbed on the cat-
alyst surface [26,8]. Hence, the photocatalytic experiments were
performed after an initial dark equilibration time of 30 min.
Fig. 7b shows percentage degradation with time for the cata-
lysts under study. H TiO2 shows an average degradation of 96%
compared to P25 (76%), Hombicat (61%) and C TiO2 (65%). The
higher efficiency for hydrothermal TiO2 may be due to altered band
gap, phase being purely anatase with high surface to volume ratio
coherent with the previously reported results of Abdennouri et al.
[16].
The kinetics of the reaction was evaluated by plotting ln (C0 /C)
with respect to time (Fig. 7c) [19]. Pseudo first-order kinetics was
assumed where C0 being the initial concentration of 2,4-D and C
being the concentration at any time t, k is the rate constant eval-
uated from the slopes of the plot [27]. The rate constants with R2
values are given in Table 2. The increasing order of rate constant for
Fig. 8. Percentage TOC removal of synthesized catalysts. H-TiO2 (1.97E–2 min−1 ) > P25 (1E–2 min −1 ) > C TiO2 (0.98E–2 min
−1 ) > Hombicat (0.59 E–2 min −1 ) indicates higher degradation per-

formance of H TiO2 . This could be due to effect of synthesis route


3.5. DRS spectra
which facilitates the generation of free • OH and • O2 − superoxide
radicals on the surface. Moreover, band gap alteration resulted in
Diffuse reflectance spectra elucidates the electronic structure of
increased active sites on the surface with lower recombination.
the material from estimation of optical band gap. The UV–vis diffuse
Under similar experimental conditions, reduction in total organic
reflectance spectrum of as-synthesized H TiO2 shows a sharp edge
content was analyzed (Fig. 8) where a reduction upto 24% was
absorption peak at 391 nm compared to 380 nm and 376 nm for P25
reported for H TiO2 than other catalysts appending to the bet-
and Hombicat respectively. The absorption peaks less than 300 nm
ter degradation performance of H TiO2 . The variation in the rate
in Fig. 6a could be due to the elastic scattering of particles by UV
constants of the catalyst can be ascribed to decreased penetra-
light and phase transformation of Ti4+ to Ti3+ [24]. The band edge of
tion depth due to varying specific surface area of the catalyst and
catalyst could be estimated according to the following equation
2/n
  increased elastic scattering of the incident light. The behavior of
(Ah␯) = k h␯ − E g (1) H-TiO2 is in comparison with the results of Kisch et al. [28].
Where A, h, , k and Eg are absorption coefficient, Planck
constant, light frequency, proportionality constant and band gap, 3.7. Photocatalytic degradation in solar light
respectively [19]. Further, the band gaps (Eg) are obtained by
extrapolating the straight line to the (h) axis intercept as shown Solar degradation has implications towards commercial viabil-
in Fig. 6b for P25, Hombicat and H TiO2 . The calculated band gap ity of the prepared catalysts. Fig. 9a shows the percentage removal
of the P25 (3.26 eV) and Hombicat (3.29 eV) can be compared with of 2,4-D under solar spectrum where the percentage degradation
that of H TiO2 (3.16 eV) where the reduction in the band gap for for H TiO2 is 83% while it is 78%, 63% and 50% for P25, Hombicat
H TiO2 may be due to localized Ti3+ gap states and the oxygen and C TiO2 respectively. The kinetics of the photocatalytic degra-
vacancies brought by hydrothermal synthesis [25]. dation assuming pseudo first order is plotted in Fig. 9b where the
rate constant with R2 is reported in Table 2. Higher degradation per-
3.6. Photocatalytic degradation in UV light formance of H TiO2 compared to commercially catalysts in solar
light could be due to the altered band gap affected by its synthesis
The photocatalytic degradation of 2,4-D on various synthesized [9].
and commercial catalyst were performed in a fabricated reactor for Further, Fig. 10 shows the comparative percentage degradation
120 min prior to which adsorptive behavior of the catalysts were of 2,4-D under UV and solar ambience. The degradation efficacy of
studied. As shown in Fig. 7a, the catalyst chosen attains equilib- H TiO2 is higher in UV irradiation as compared to the solar spec-

Fig. 9. (a) Solar degradation plot of 2,4-D. (b) Kinetic plot of 2,4-D degradation under solar light.
S. S et al. / Applied Surface Science 449 (2018) 371–379 377

Fig. 11. Comparison of effect of aeration for the degradation of 2,4-D.

Fig. 10. Comparison of percentage removal of 2,4-D in UV and solar ambience.


results are shown in Fig. 11. The percentage degradation with and
without aeration was 96% and 59% respectively indicating the fact
trum. This can be ascribed to the higher threshold energy of UV that dissolved oxygen act as an electron sink in the photodegrada-
rays and the lower quanta of UV radiations in the solar spectra. The tion process [32] and retards the electron hole recombination. The
UV rate constants of the catalysts are in general higher than that molecular oxygen in the inlet air can either recombine or react with
of the solar rate constants. The possible explanation for this is the electron donors or acceptors which are adsorbed on the H TiO2
monochromatic behavior of UV leading to efficient n-␲* electron surface or trapped within the surrounding electrical double layer of
transition and presence of acute amount of UV in solar spectra (8%) the charged particles leading to the production of hydroxyl radicals
which poses insufficient threshold energy to push the electrons which promotes the oxidation of 2,4-D.
leading to the 2,4-D degradation [29]. The photo stability of the catalysts was tested by the proce-
Table 3 shows the cumulative literature summary of degra- dure proposed by Lu et al. [19]. The photo catalysts were separated,
dation efficiencies obtained at various parameters such as lamp washed and reused for degradation of 2,4-D upto 3 cycles. Fig. 12
intensity, degradation time, catalyst dosage and the percentage a demonstrates the degradation performance of H TiO2 and P25
efficiency attained for degrading 2,4-D [5,12,16,30,31]. It can be where reproducibility of the degradation rate was within 10% when
inferred that, by using same lamp intensity, hydrothermally syn- observed for 3 repeated cycles. XRD pattern of the catalyst (Fig. 12b)
thesized TiO2 catalyst as reported in this study can degrade recorded after reuse was similar with that of prepared catalyst,
2,4-D (50 ppm) to a maximum of 96% within 2 h than compared demonstrating the stability of the chosen catalyst.
to graphene/TiO2 for 100% degradation after 15 h as reported
by Huang et al. [12]. The attainment of 96% efficiency for 2,4- 3.9. Photocatalytic degradation mechanism
D degradation could be due to its uniform mixing, easy and
facile adsorption and desorption of the intermediates of 2,4-D on In the presence of catalyst, light and aeration, holes and protons
hydrothermal TiO2 surface, altered band gap and active Ti3+ states. are generated which then reacted with dissolved oxygen to produce
The reason for longer degradation time on graphene/TiO2 in Table 3 hydroxyl radicals (• OH), resulting in initiation of degradation of 2,4-
could be due to stronger adsorption of 2,4-D followed by its miti- D. Fig. 13 shows the mechanism of degradation as understood from
gation. the intermediates formed during the hydrothermal degradation of
TiO2 . The procedure for LCMS/MS analysis, the residence time of
3.8. Effect of aeration and photo stability measurements various intermediates and its m/z ratio are reported in the support-
ing information (Fig. B1–B4). The attack of an electron and hydroxyl
In order to check the effect of aeration, the experiments were radical on 2,4-D yielded chlorinated hydrophenoxyacetic acid and
conducted with and without aeration in UV light for H TiO2 and the 2,4 dichlorophenol respectively. Due to the electrophilic attack, 2,4

Table 3
Comparison of the degradation parameters for 2,4-D.

References Light source Catalyst Time (min) Catalyst dosage 2,4-D dosage Degradation
(%)

Del et al. [16] 250 W Visible light TiO2 doped with 180 100 mg 40 ppm 83
Nitrogen
Kundu et al.[30] Portable germicidal TiO2 /UV/O3 75/180 40 ppm 39/96
lamp, 15 W
Elsalamony et al. [31] UV lamp, 254 nm TiO2 doped with 150 1 g/L 50 ppm 46
copper
Bian et al. [5] UV lamp, 254 nm Tourmaline coated 40 500 mg/L 20 ppm 90 above
(200–1000 W/m) TiO2
Abdennori et al. [16] Medium pressure Pt/TiO2 90 120 mg/L 20 ppm 90 above
mercury lamp (400 W)
Huang et al. [12] 11 W, UV lamp Graphene/TiO2 15 h 0.8 g/L 50 ppm 100
Present work 11 W, UV lamp Hydrothermal TiO2 2h 0.2 g/L 50 ppm 96
378 S. S et al. / Applied Surface Science 449 (2018) 371–379

Fig. 12. (a) Photodegradation rate and (b) XRD patterns of H-TiO2 and P25 for different cycles.

Fig. 13. Mechanism of degradation of 2,4-D.

dichlorophenol yielded 2 chlorophenol followed by hydroxylation and retarded fast electron-hole pair recombination effected by aer-
resulting in the formation polyhydroxylated intermediates such as ation. The degradation followed pseudo first-order kinetics in both
1,2,4 benzenetriol, chlorohydroquinone and 2,4 dichlorocatechol UV and solar irradiation conditions whereby the mineralization of
(Fig. B2) [12]. Further, these intermediates undergo ring opening the organic pollutant is illustrated from the intermediates identi-
mechanism where the surface catalyzed reaction occurs by bond fied from LCMS/MS study.
hydrolysis of C Cl and C O followed by the mineralization of the
organic pollutant [13,14,16,17] to get degraded to CO2 , H2 O and Acknowledgements
other gaseous products. The amendments in photocatalytic activity
for hydrothermally synthesized TiO2 were explicitly discussed by The authors acknowledge funding from TEQIP-II (Technical
comparing with commercially available TiO2 for degradation of 2,4- Education Quality Improvement Programme) of BMS College of
D. Further analysis of effect of light intensity, catalyst dosage, pH Engineering, Bangalore and also acknowledge Department of Sci-
and toxicity studies will guide us in large scale commercialization. ence and Technology – Science and Engineering Research Board
(DST-SERB) New Delhi, India for establishing facilities through Early
4. Conclusion Career Research Award (ECR/2016/000841). The authors sincerely
thank Indian Institute of Technology Madras (IITM) and National
TiO2 was synthesized by hydrothermal method and character- Centre for Catalysis Research (NCCR) and Agilent Technologies for
ized by XRD, SEM, TEM, DRS-UV, FTIR and Raman spectroscopy. The their analytical assistance.
degradation of 2,4-D was studied for hydrothermally synthesized
TiO2 and compared with P25, Hombicat and commercial grade TiO2 . Appendix A. Supplementary data
Hydrothermally synthesized TiO2 catalyst showed a maximum pol-
lutant degradation of 96% and 83% under UV light and visible solar Supplementary data associated with this article can be found,
light respectively. This enhanced photocatalytic activity can be in the online version, at https://doi.org/10.1016/j.apsusc.2018.02.
attributed to the pure anatase phase formation, reduced band-gap 051.
S. S et al. / Applied Surface Science 449 (2018) 371–379 379

References [16] M. Abdennouri, A. Elhalil, M. Farnane, H. Tounsadi, F.Z. Mahjoubi, R.


Elmoubarki, M. Sadiq, L. Khamar, A. Galadi, M. Baalala, M. Bensitel, Y. El
[1] F. Perreault, A.F. de Faria, M. Elimelech, Environmental applications of Hafiane, A. Smith, N. Barka, Photocatalytic degradation of 2,4-D and 2,4-DP
graphene-based nanomaterials, Chem. Soc. Rev. 44 (2015) 5861–5896. herbicides on Pt/TiO2 nanoparticles, J. Saudi Chem. Soc. 19 (2015) 485–493.
[2] V.R. Posa, V. Annavaram, J.R. Koduru, P. Bobbala, V. Madhavi, A.R. Somala, [17] S.M. Gupta, M. Tripathi, An overview of commonly used semiconductor
Preparation of graphene–TiO2 nano composite and photocatalytic nanoparticles in photocatalysis, High Energy Chem. 46 (2011) 1–9.
degradation of Rhodamine-B under solar light irradiation, J. Exp. Nanosci. 11 [18] J. Kirchnerova, H. Cohen, M.L. Guy, D. Klvana, Photocatalytic oxidation of
(2016) 722–736. n-butanol under fluorescent visible light lamp over commercial TiO2 , Appl.
[3] Z. Lu, Z. Zhu, D. Wang, Z. Ma, W. Shi, Y. Yan, X. Zhao, H. Dong, Li Yang, Z. Hua, Catal. A Gen. 282 (2005) 321–332.
Specific oriented recognition of a new stable ICTX@Mfa with retrievability for [19] Z. Lu, Z. Yu, J. Dong, M. Song, Y. Liu, X. Liu, Z. Ma, H. Su, Y. Yan, P. Huo, Facile
selective photocatalytic degrading of ciprofloxacin, Catal. Sci. Technol. 6 microwave synthesis of a Z-scheme imprinted ZnFe2 O4 /Ag/PEDOT with the
(2016) 1367–1377. specific recognition ability towards improving photocatalytic activity and
[4] Z. Lu, X. Zhao, Z. Zhu, Y. Yan, W. Shi, H. Dong, Z. Ma, N. Gao, Y. Wang, H. Huang, selectivity for tetracycline, Chem. Eng. J. (2017), http://dx.doi.org/10.1016/j.
Enhanced recyclability, stability, and selectivity of CdS/C@Fe3O4 cej.2017.12.115 (in press).
nanoreactors for orientation photodegradation of ciprofloxacin, Chem. Eur. J. [20] M. Mehrvar, W.A. Anderson, M. Moo-Young, Comparison of the
21 (2015) 18528–18533. photoactivities of two commercial titanium dioxide powders in the
[5] X. Bian, J. Chen, R. Ji, Degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) degradation of 1,4-dioxane, Int. J. Photo Energy 4 (2002) 140–146.
by novel photocatalytic material of tourmaline-coated TiO2 nanoparticles, [21] X. Shao, W. Lu, R. Zhang, F. Pan, Enhanced photocatalytic activity of TiO2 -C
kinetic study and model, Materials 6 (2013) 1530–1542. hybrid aerogels for methylene blue degradation, Sci. Rep. Nat. 3 (2013) 3018.
[6] J. Tifeng, H. Guo, Q. Zhang, Q. Peng, Y. Tang, X. Yan, B. Li, Reduced graphene [22] A.C. Ferrari, Raman spectroscopy of graphene and graphite: disorder,
oxide-based silver nanoparticle-containing composite hydrogel as highly electron-phonon coupling, doping and non-adiabatic effects, Solid State
efficient dye catalysts for wastewater treatment, Sci. Rep. 5 (2015) 11873. Commun. 43 (2008) 47–57.
[7] D. Mao, X. Shimin, W. Tianhui, Z. Deqiang, Z. Qian, F. Zihong, Z. Yao, J. [23] V. Šteng, S. Bakardjieva, T.M. Grygar, J. Bludská1, M. Kormunda,
Fangying, H. Qiang, X. Xuan, Preparation of a microspherical silver-reduced TiO2 -graphene oxide nanocomposite as advanced photocatalytic materials,
graphene oxide-bismuth vanadate composite and evaluation of its Chem. Cent. J. 7 (41) (2013) 1–12.
photocatalytic activity, Materials 9 (160) (2016) 1–14. [24] A.M. Shehap, D.S. Akil, Structural and optical properties of nano TiO2 /PVA for
[8] Z. Lu, Z. Yu, J. Dong, M. Song, Y. Liu, X. Liu, D. Fan, Z. Ma, Y. Yan, P. Huo, different composite thin films, Int. J. Nanoelectron. Mater. 9 (2006) 17–36.
Construction of stable core–shell imprinted [25] M.M. Khan, S.A. Ansari, D. Pradhan, M.O. Ansari, D.H. Han, J. Leea, M.H. Cho,
Ag-(poly-o-phenylenediamine)/CoFe2 O4 photocatalyst endowed with the Band gap engineered TiO2 nanoparticles for visible light induced photo
specific recognition capability for selective photodegradation of ciprofloxacin, electrochemical and photocatalytic studies, J. Mater. Chem. A 2 (2014)
RSC Adv. 7 (2017) 48894–48903. 637–640.
[9] M. Kavitha, C. Gopinathan, P. Pandi, Synthesis and characterization of TiO2 [26] Z. Lu, X. Zhao, Z. Zhu, M. Song, N. Gao, Y. Wang, Z. Ma, W. Shi, Y. Yan, H. Dong,
nano-powders in hydrothermal and sol-gel method, Int. J. Adv. Res. Technol. 2 A novel hollow capsule-like recyclable functional ZnO/C/Fe3 O4 endowed with
(2013). three-dimensional oriented recognition ability for selectively photodegrading
[10] D. Fonseca MB, L. Glusczak, B.S. Moraes, D. Menezes, A. Pretto, M.A. Tierno, R. danofloxacin mesylate, Catal. Sci. Technol. 6 (2016) 6513–6524.
Zanella, F.F. Gonçalves, V.L. Loro, Ecotoxicol. Environ. Saf. 69 (3) (2008) [27] Z. Lu, Y. Luo, M. He, P. Huo, T. Chen, W. Shi, Y. Yan, J. Pan, Z. Ma, S. Yang,
416–420. Preparation and performance of a novel magnetic conductive imprinted
[11] O.M. Aly, S.D. Faust, Herbicides in surface waters studies on fate of 2,4-D and photocatalyst for selective photodegradation of antibiotic solution, RSC Adv. 3
ester derivatives in natural surface waters, J. Agric. Food Chem. 12 (1964) (2013) 18373–18382.
541–546. [28] H. Kisch, D. Bahnemann, Best practice in photocatalysis: comparing rates or
[12] D. Huang, T. Yang, Z. Mo, Q. Guo, S. Quan, C. Luo, L. Liu, Preparation of apparent quantum yields? J. Phys. Chem. Lett. 6 (2015) 1907–1910.
graphene/TiO2 composite nanomaterials and its photocatalytic performance [29] R. Mariselvam, A.J.A. Ranjitsingh, P.M. Selvakumar, A.A. Alarfaj, M.A.
for the degradation of 2,4-dichlorophenoxyacetic acid, J. Nanomater. (2016) Munusamy, Spectral studies of UV and solar photocatalytic degradation of azo
11, http://dx.doi.org/10.1155/2016/5858906, Article ID 5858906. dye and textile dye effluents using green synthesized silver nanoparticles,
[13] Z. Lu, F. Chen, M. He, M. Song, Z. Ma, W. Shi, Y. Yan, J. Lan, F. Li, P. Xiao, Bioinorg. Chem. Appl. (2016) (2016).
Microwave synthesis of a novel magnetic imprinted TiO2 photocatalyst with [30] S. Kundu, A. Pala, A.K. Dikshit, UV induced degradation of herbicide 2,4-D:
excellent transparency for selective photodegradation of enrofloxacin kinetics, mechanism and effect of various conditions on the degradation, Sep.
hydrochloride residues solution, Chem. Eng. J. 249 (2014) 15–26. Purif. Technol. 44 (2005) 121–129.
[14] Z. Lu, P. Huo, Y. Luo, X. Liu, D. Wb, X. Gao, C. Li, Y. Yan, Performance of [31] R.A. Elsalamony, S.A. Mahmoud, Preparation of nano structured ruthenium
molecularly imprinted photocatalysts based on fly-ash cenospheres for doped titania for the photocatalytic degradation of 2-chlorophenol under
selective photodegradation of single and ternary antibiotics solution, J. Mol. visible light, Arabian J. Chem. 10 (2) (2017) 194–205, http://dx.doi.org/10.
Catal. A: Chem. 378 (2013) 91–98. 1016/j.arabjc.2012.06.008.
[15] K. Del, Á. Sancheza, Photocatalytic degradation of 2,4-dichlorophenoxyacetic [32] S. Devipriya, S. Yesodharan, Photocatalytic degradation of pesticide
acid under visible light: effect of synthesis route, Mater. Chem. Phys. 139 contaminants in water, Sol. Energy Mater. Sol. Cells 86 (3) (2005) 309–348.
(2013) 423–430.

You might also like