You are on page 1of 16

International Journal of Environmental Analytical

Chemistry

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/geac20

Facile synthesis of N-MgSb2O6 trirutile antimonate


and its enhanced photocatalytic performance.

Arunkumar Nagarajan & Saraschandra Naraginti

To cite this article: Arunkumar Nagarajan & Saraschandra Naraginti (2020): Facile synthesis of N-
MgSb2O6 trirutile antimonate and its enhanced photocatalytic performance., International Journal of
Environmental Analytical Chemistry, DOI: 10.1080/03067319.2020.1842389

To link to this article: https://doi.org/10.1080/03067319.2020.1842389

View supplementary material

Published online: 06 Nov 2020.

Submit your article to this journal

Article views: 35

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=geac20
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY
https://doi.org/10.1080/03067319.2020.1842389

Facile synthesis of N-MgSb2O6 trirutile antimonate and its


enhanced photocatalytic performance.
Arunkumar Nagarajana and Saraschandra Naragintib
a
PG & Research Department of Chemistry, Saraswathi Narayanan College, Madurai, India; bSchool of
Biological and Chemical Engineering, Anhui Polytechnic University, Wuhu, China

ABSTRACT ARTICLE HISTORY


In this study, nanocrystallinetrirutile N-doped MgSb2O6 was pre­ Received 10 September 2020
pared by a facile solution combustion method and its photocataly­ Accepted 16 October 2020
tic activity was reported for the first time. Products obtained with KEYWORDS
different oxidant to fuel ratios (O/F) were in crystalline single phasic Metal oxide; optical
N-doped MgSb2O6 confirmed by X-ray diffraction. The prepared properties; Raman
powders were well characterised by using scanning electron micro­ spectroscopy; degradation;
scopy, energy dispersive X-ray analysis, transmission electron p-bromophenol
microscopy, and X-ray photoelectron spectroscopy. Raman spectra
showed all the characteristic peaks of trirutile structure with sig­
nificant shift and broadening in the nanocrystalline product com­
pared to that of microcrystalline product. UV-visible DRS analysis
confirmed the decrease in band gap with an increase in O/F which
is in the range of 4.05 eV–3.30 eV. Nearly 96% of p-bromophenol
degradation (p-BP, 10 mg/L) was achieved over nanocrystalline
N-doped MgSb2O6 after 60 min of irradiation at a rate of
0.069 min−1. Furthermore, radical quantification experiments and
electron spin resonance (ESR) analysis revealed that •OH and •O2
− were the main ROS responsible for photodegradation of p-BP.
Hence, this study presented the enhanced photocatalytic activity of
combustion synthesised nanocrystalline N-doped MgSb2O6 for the
first time and its activity was found to be correlated with N-doping,
cation ordering, particle size, and crystallinity.

1. Introduction
Inorganic metal oxide semiconductor-based photocatalysts have been explored exten­
sively for environmental remediation and hence, are widely employed for abatement of
a variety of organic contaminants, wastewater treatment and also for water splitting [1].
Semiconductor oxide materials act as photocatalysts towards light-induced photochemi­
cal reactions due to its electronic structure characterised by a filled valence band (VB) and
an empty conduction band (CB) separated by an appropriate bandgap [2]. Such metal
oxide-based semiconductor materials is investigated as potential materials not only for
photocatalytic applications but also for many other applications such as optical devices,
solar cells [3]. In general, MgO and Sb2O3were know as good catalysts for organic
compound degradation [4]. In the same way, we found new efficient AB2O6 type complex

CONTACT Saraschandra Naraginti nsaraschandra@gmail.com


Supplemental data for this article can be accessed here.
© 2020 Informa UK Limited, trading as Taylor & Francis Group
2 A. NAGARAJAN AND S. NARAGINTI

oxides for efficient degradation of organic pollutant. Antimonate-based oxides of the


formula MSb2O6 (M = Zn, Cd, Ca, Sr, and Ba) having a wide bandgap are an important class
of photocatalysts [5]. MgSb2O6 is useful as an electrode material for dye sensitised solar
cells [6]. MSb2O6 crystallises in two types of structures, trirutile and PbSb2O6 structure.
Ionic radii of divalent A-site cation less than 0.8 Å belong to trirutile structure (for example,
MSb2O6; M = Zn, Mg, Co, Cu, Ni, Fe), while increasing the ionic radius of A-site cation
favours PbSb2O6 structure (M = Cd, Ca, Sr, Ba, Pb) [6]. Mineralogically, MgSb2O6 is known
as bystromite and crystallises in trirutile-type structure, which has the space group P41
/mnm. Its structure is built up from two edge-sharing SbO6octahedra along the c axis,
sharing their corners with two other SbO6octahedra. M atoms are octahedrally coordi­
nated by six O atoms in each rutile unit. The features are (1) Dominant Sb 5 s orbitals at the
bottom of the conduction band (2) Edge-shared MO6 octahedron (M metal) aligned
regularly in the trirutile-type structure [7]. All these antimonates based photocatalyst
contain distorted Sb–O polyhedron which is attributed for their photocatalytic activity.
Thus, it would be worthwhile to investigate the feasibility of MgSb2O6 as a significant
photocatalyst for degradation of p-bromophenol. Furthermore, heteroatom doping (N) is
anticipated to further enhance the catalytic activity for oxidation-reduction reactions.
To the best of our knowledge, this is the first report on the synthesis of nanocrystalline
trirutileN-doped MgSb2O6 by solution combustion method and its photocatalytic activity.
The catalyst performed excellently in the degradation of p-BP. Furthermore, radical
quantification and ESR analysis was carried out to identify the main ROS responsible for
photodegradation of p-BP and the plausible photocatalytic degradation mechanism was
also proposed.

2. Experimental
2.1. Materials
Starting materials for the synthesis of MgSb2O6 were Antimony Trioxide 99.99% (Sigma
Aldrich), and Magnesium oxide 99.99% (Sigma Aldrich).Terephthalic acid (TA), Nitro blue
tetrazolium chloride (NBT), and p-bromophenol was obtained from SD-Fine Chemicals.

2.2. Preparation of microcrystalline MgSb2O6 by solid-state synthesis


The standard solid-state method was followed to synthesise MgSb2O6 [6]. Briefly, the MgO
and Sb2O3were has taken in stoichiometric quantities, and mixed thoroughly followed by
heating at 600°C for 12 h and finally at 900°C for 6 h with mediate grindings using
a programmable thermolyne furnace (Model No: F46110CM-33) at a heating rate of 4°
C/min.

2.3. Preparation of nanocrystalline N-doped MgSb2O6 by combustion synthesis


Initially magnesium nitrate was dissolved in small aliquot of deionised water. To this,
SbCl3 mixed with NH4NO3 (in 1:3 ratio) was then added followed by addition of urea (fuel).
Stirring was continued for 30 min to get a clear dispersion and then introduced into
a preheated furnace and kept at 400°C [7]. The volume of the solution rapidly decreased
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 3

Figure 1. Schematic representation of combustion synthesis of N-doped MgSb2O6.

with boiling and combustion takes place quickly with intense flame along with evolution
of gases yielding the products. Figure 1 shows the pictorial representation of combustion
process employed in this study. The Oxidant-to-Fuel (O/F) is one of the important para­
meters which plays a crucial role in combustion process where product formation is varied
with different ratios [8].In the present study three different (O/F) ratios namely 0.5, 1.0 and
2.0 were attempted for combustion process.

2.4. Characterisation
X-ray diffraction analysis was carried out by using Bruker D8 Advance X-ray diffractometer
(CuKα radiation Kα λ = 1.5406 Å).SEM (FESEM Supra 55 – CARL ZEISS, Germany) analysis
was carried out to study the surface morphology and microstructure. EDS (Oxford
Instruments, Liquid Nitrogen-free SDD X MAX 50 EDS) analysis was done to study the
cationic ratio. JEOL JEM 2100Transmission Electron Microscope was used for particle size
measurement. The standard BET (Micromeritics ASAP 2020) method was employed for
surface area analysis. X-Ray Photoelectron Spectra (XPS) were recorded using K-Alpha
instrument (XPS K-Alpha surface analysis, Thermo fisher scientific, UK). Raman spectra
were recorded on a laser Raman Spectrometer (Horiba Scientific/Lab RAM HR). Optical
absorption characteristics of the products were measured by using JASCO UV-Vis-NIR
V-670 spectrophotometer.

2.5. Photocatalytic degradation of p-bromophenol


Photocatalytic study was carried out at room temperature by using medium pressure
mercury lamps (8 W, 254 nm) as UV light source. In each degradation experiment 20 mg of
the photocatalyst was dispersed in 100 mL of 10 mg/L p-BP solution. The suspension was
then stirred in the dark for 30 min to reach the adsorption/desorption equilibrium
followed by UV light irradiation while stirring. Samples were collected at regular intervals
and was eventually analysed by using a high-performance liquid chromatography
4 A. NAGARAJAN AND S. NARAGINTI

(Shimadzu, Japan) equipped with Eclipse XDBC18 (4.6 x 150 mm, 5 µm) reverse phase
column. The UV detector wavelength was 280 nm, while the mobile phase consisted of
water (60%) and methanol (40%) fed at a flow rate of 1 mL min.−1

2.6. Radical quantification experiments


The generation of •OH was analysed by Photoluminescence (PL) technique (HITACHI
(F-7000) Fluorescence Spectrophotometer) using terephthalic acid (TA) as the probe.
For this experiment, 20 mg of the photocatalyst was added to a 50 ml mixture of TA
solution (3 mmol) and NaOH (10 mmol). After UV irradiation at regular time intervals
samples were collected and corresponding PL intensities were recorded. Similarly, for
estimation of •O2−, the photocatalytic reaction was carried out in NBTsolution at
a concentration of 10 mg/L. NBT can be peculiarly reduced by photogenerated •O2− by
forming an insoluble purple formazan compound in the aqueous solutions. The concen­
tration of NBT in the aliquots was monitored by measuring the absorbance at λmax259 nm
in a UV-Visible spectrophotometer (Hitachi – 2910)

3. Results and discussion


3.1. Structural characterisation
The XRD pattern confirmed the single phasic (JCPDS No.037–1470) microcrystalline
MgSb2O6 ordered trirutile structure of solid state synthesised MgSb2O6which is having
the characteristic reflections of (002), (101) and (112) that showed high structural ordering
with the lattice parameters of a = 4.6507 Å; c = 9.2319 Åand the space group of P42/mnm
(Figure 2a). Further, the combustion synthesised products with different fuel ratios con­
firmed that the as prepared products itself are in nanocrystalline single phase with an
average crystallite size D ~ 20 nm possessing disordered trirutile structure (absence of the
above characteristic reflections)(Figure 2b). More importantly, all the as prepared pro­
duces calcined at above 500°C showed significant peak broadening indicating nanocrys­
talline nature and crystallising in tetragonal unit cell with parameters, a ~ 4.6692 Å,
c ~ 3.0784 Å. No additional diffraction peaks was observed for N doped phases. In ordered
trirutile structure, Mg2+ and Sb5+were arranged in a 1:2 order along z-axis so that the ‘c’
cell parameter is equal to three times that of disordered rutile structure, giving some
characteristic reflections such as (002), (101), and (112) due to cationic ordering. The
structural disordering observed in as synthesised nanocrystalline-doped MgSb2O6 may be
due to the low synthetic temperature employed as ordered trirutile structure was attained
only at higher processing temperature (Figure 2b). In order to improve the network
(structural) ordering in nanocrystalline N-doped MgSb2O6 and further to achieve the
disorder-order transition, the as synthesised nanocrystalline product obtained at O/F
ratio of 1.0 was calcined at different temperatures (500°C, 600°C, 700°C and 900°C) for
12 h (Figure 2c). It is evident that at calcinations less than 700°C, no ordering was
observed. Above 700°C ordering takes place as revealed by the appearance of character­
istic reflections in the XRD patterns. Mostly, increase in calcination temperature increases
the crystallite size which is further indicated by the decrease in peak broadening.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 5

Figure 2. XRD patterns of (a) MgSb2O6 (solid state synthesis), (b) nanocrystalline N-doped MgSb2O6
with varying O/F ratio (combustion synthesis), (c) nanocrystalline N-doped MgSb2O6calcined at
various temperature for O/F = 1.
6 A. NAGARAJAN AND S. NARAGINTI

Figure 3. SEM and EDAX pattern (a&b); TEM image and SAED pattern (c&d)of nanocrystallineN-doped
MgSb2O6.

3.2. Surface morphology and microstructure


The morphology and microstructure of nanocrystalline N-doped MgSb2O6 (O/F = 1) was
investigated by SEM analysis and indicated the spherical-shaped particles of the products
(Figure 3a). EDAX analysis confirmed the elemental composition and the cationic ratio
was equal to the nominal compositions (Figure 3b). Furthermore, TEM images of the as
prepared nanocrystallineN-doped MgSb2O6showed spherical-shaped particles with an
average particle size of 10–20 nm with agglomeration and this is consistent with crystal­
lite size calculated from XRD (Figure 3c). The formation of small polycrystalline grains of
the prepared particles was confirmed by SAED (selected area electron diffraction) pattern
by indicating diffused continuous ring pattern (Figure 3d). Furthermore, BET surface area
analysis of microcrystalline MgSb2O6and nanocrystalline N-doped MgSb2O6 (as prepared
and calcined at 500°C, 600°C and 700°C) indicated that the combustion synthesised
compounds were nanocrystalline with higher surface area (Table 1).

3.3. Raman spectroscopy analysis


Trirutile structures belong to space group P42/mnmand the corresponding Raman bands
are observed in the range of 200 to 800 cm−1for microcrystalline MgSb2O6and nanocrys­
talline N doped MgSb2O6 (Figure 4). Bands observed in the range of 800–600 cm−1were
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 7

Table 1. BET surface area of microcrystal­


line MgSb2O6 and nanocrystalline
N-doped MgSb2O6.
O/F ratio BET surface area (m2/g)
#2 35.8
#1 36.7
#0.5 32.9
*500 °C 28.8
*600 °C 26.3
*700 °C 24.7
solid state 2.3
#as prepared *O/F = 1

corresponded to simple bridging bond vibrations of Sb−Ob−Sb, bands in the range of

Figure 4. Raman spectra of Microcrystalline and Nanocrystalline MgSb2O6at O/F = 1.

600–500 cm−1are corresponded to Sb−Ocyc vibrations and Sb–Ob−Sb vibrations coupled


with M2+−O vibration, while, bands in the range of 500–200 cm−1 are corresponded to M2
+
−O vibrations [9].

3.4. XPS analysis


XPS analysis was carried out in order to determine the chemical composition and
elemental states. The high resolution XPS spectra of nanocrystallineN-MgSb2O6-x (O/
F = 1.0) showed peaks at 530.5, 531.5 and 533.4 eV (Figure 5a). The peak at 530.5 eV
could be assigned to Sb 3d5/2 [10], while at 540.5 eV could be corresponded to Sb 3d3/2
[10] (Figure 5b). The peak at binding energy 531.5 eV is attributed to O2- ions in the
8 A. NAGARAJAN AND S. NARAGINTI

Figure 5. XPS spectra of N-doped MgSb2O6.

oxygen-deficient regions [11] within the matrix of MgSb2O6-x. The peak at 533.4 eV could
be assigned to O 1s which is loosely bound surface oxygen [11]. The binding energy peak
at 1306 eV is corresponded to Mg 1s (Figure 5c) [12]. The quantification of oxygen amount
was restricted by the fact that O 1s and Sb 3d5/2 peaks were overlapped at 530.5 eV.
Furthermore, XPS spectrum of N 1s showed peaks at 395.3, 399.5 and 406.1 eV corre­
sponded different chemical states. From the literature, four types of N species could be
possible in N doped oxides i.e., as surface adsorbed NH3, O – Mg – N form of oxynitride, in
the form of NO, NO2 and N3- ion which bounds to surface oxygen sites. No peak at
397.6 eV was observed in our composition which ruled out N3- ion. However, peaks
observed at 395.3, 399.5 and 406.1 eV confirmed the presence of N in the form of NH3
species [13], oxynitride (O – Mg – N) [14], NO or NO2 like species respectively [15].

3.5. Optical band gap


UV-vis DRS spectra analysis was carried out at the wavelength range 200–1200 nm and
shown in Figure 6. The reflectance was increased with increase in fuel content. The
reflectance was transformed to absorbance by using Kubelka – Munk conversion (1)

ðα=SÞ ¼ ð1 RÞ2 =2R (1)


INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 9

Figure 6. (a) Diffused reflectance spectra (b) KM Absorbance spectra (c) Tauc’s plot for ordered and
disordered N-doped MgSb2O6.
10 A. NAGARAJAN AND S. NARAGINTI

Where α = absorption coefficient, S = scattering coefficient and R = diffused reflectance at


particular energy [16].Microcrystalline MgSb2O6absorption spectra showed the absorp­
tion edge at 320 nm whereas for N-doped MgSb2O6-x the absorption edge was in the
range of 320–430 nm based on the O/F ratio (Figure 6a). Simultaneously, a marked
absorption edge was observed for nanocrystalline N-doped MgSb2O6-xin visible and
near infrared regions. This could be due to the oxygen vacancies created by various
N-doping amounts that affected through altering fuel amounts during combustion [17].
In nanocrystallineN-doped products, the absorption edge was shifted to higher wave­
length indicating a substantial reduction in band gap compared to that of microcrystalline
MgSb2O6.The optical band gap was determined from absorption spectrum using Tauc’s
relation, Equation (2) [18].
ðαhυÞn ¼ Aðhυ Eg Þ (2)
Where α represents absorption coefficient, hνrepresents discrete photo energy, A is constant,
Eg is the band gap and exponent n depends on type of transition, n = 2 for direct transition.
The band gap of microcrystalline MgSb2O6wasabout 4.05 eV while it was in the range of
3.30–3.80 eV in the case of nanocrystalline N-MgSb2O6depends on O/F ratio, and the band
gap significantly decreased with increase in O/F ratio (Figure 6b). This facile urea mediated
combustion synthesis resulted in about 0.5 eV of band gap reduction in MgSb2O6 due to
doping of N compared to that of microcrystalline MgSb2O6. Therefore a higher wavelength
shift was observed at about 100 nm in the absorption edge. The colours of the compounds
obtained with varying fuel contents are consistent with the band gaps (Inset of Figure 6b).
Doping of N could lead to formation of localised energy level/defects within the band gap
closer to VB while on top of VB causing band gap reduction [19]. Furthermore, to understand
the ordering-disordering influence of cation in trirutile structure the band gaps of disordered
and ordered phases of N-MgSb2O6were calculated (Figure 6c). The disordered N-MgSb2O6 (O/
F = 1) was innanocrystallinephase with a band gap of 3.45 eV. To improve the cation ordering,
the disordered phase was calcined at 700°C for12h and the resulted ordered, more crystalline
N-MgSb2O6 has a band gap of 3.85 eV. The latter exhibited a significant blue shift in the
absorption edge when compared to the former as indicated by DRS spectra and their colours.
In latter case, it indicated that the elimination of N-doping could be due calcination at 700°C
and the disordered N-MgSb2O6 changed to ordered MgSb2O6 which leads to increase in band
gap and crystallinity.

3.6. Photocatalytic activity and plausible degradation mechanism of p-BP


Some of the metal antimonite were reported previously by few researchers as potential
photoactive materials for pollutants degradation [20]. Hence, we anticipated that nanocrystal­
line N-doped MgSb2O6could be a promising photocatalytic material under UV irradiation
based on the band gap values obtained from DRS spectra and evaluated the photocatalytic
activity of micron-sized ordered MgSb2O6, nanocrystalline N-doped disordered MgSb2O6 (O/
F = 1, as prepared) and ordered trirutilenanocrystalline N-doped MgSb2O6 (700°C calcined)
during the degradation of p-bromophenol. The sequential changes of p-BP concentration
during photodegradation were recorded, and the degradation efficiency was found to be
highest for ordered trirutilenanocrystalline N-doped MgSb2O6compare to other products
after 60 min of irradiation (Figure 7a) (Table S2). The ordered N-doped MgSb2O6phase
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 11

Figure 7. (a) Temporal changes of p-BP concentration at different time intervals (b) degradation
activity of N-doped MgSb2O6prepared at different O/F ratios.

showed better performance than the disordered phase although its surface area (24.7 m2/g)
lower than the disordered phase (36.7 m2/g), indicated that crystallinity was the main
parameter for photocatalytic activity [21]. To compare the reaction kinetics of the photode­
gradation of p-BP quantitatively, Langmuir-Hinshelwood (L-H) kinetics model was applied as
per pseudo-first-order Equation (3) [22].
In ðC=Co Þ ¼ kt (3)

The photocatalytic degradation kinetics of microcrystalline MgSb2O6, nanocrystalline


N-doped MgSb2O6 and ordered N-doped MgSb2O6 followed the expected first order
which showed a linear correlation (R2 > 0.97) between ln (C/Co) and irradiation time
(Figure 7a & Figure 7b) (Table S1). Thus, the results indicated that the ordered N-doped
MgSb2O6 exhibited highest photocatalytic activity than disordered N-doped MgSb2O6
and microcrystalline MgSb2O6 in degrading p-bromophenol.
Photocatalytic activity of the semiconductor materials is mainly attributed to ROS
(reactive oxygen species) specifically •OH and •O2− from aqueous solution during light
irradiation. Hence, radical quantification and ESR techniques were carried out to under­
stand and study the possible degradation pathway over best performed catalyst (nano­
crystalline N-doped MgSb2O6). In quantification of •OH, the terephthalic acid
photoluminescence probing technique (TA-PL) was employed since TA readily reacts
with •OH and from a fluorescent compound 2-hydroxyterephthalic acid (TAOH) (Scheme
S1) [23]. PL intensity was increased with increase in time which is proportional to the
amount of •OH (Figure 8a). This is due to easy charge transfer, and improved holes on the
VB of N-doped MgSb2O6to produce more •OH. Furthermore, the ESR spectrum showed
four characteristic peaks of DMPO•OH species in aqueous dispersions which further
confirmed the generation of •OH (Figure 8b). At the same time, •O2− quantification was
also carried out thorough NBT transformation during photocatalytic reaction (Scheme S2)
[24]. The much higher transformation percentage of NBT was observed with increase in
time (Figure 8c) and strong characteristic peaks of DMPO •O2− species was also observed
in methanol dispersion further confirmed the •O2− generation (Figure 8d).
Based on the above results a possible degradation pathway of p-BP under UV light by
N-doped MgSb2O6 was proposed (Fig. S1). In general, the para-substituted phenols
12 A. NAGARAJAN AND S. NARAGINTI

Figure 8. (a) Fluorescence spectra of TAOH, (b) DMPO spin-trapping ESR spectra of •OH, (c) Spectra of
NBT transformation and (d) DMPO spin-trapping ESR spectra of O2•− at ambient temperature using
N-doped MgSb2O6under UV light.

experience degradation by generating 4-hydroquinone as an intermediate compound by


attack of •OH. In the degradation pathway, the para position of p-BP attacked by •OH to
eliminate bromide ion and forms 4-hydroquinone, owing to stronger para-directing hydroxyl
group than bromine (Fig. S1). The 4-hydroquinone further reacts with •OH to undergo
opening of the aromatic ring to generate various aliphatic alcohols and acids which will
further degrade to shorter chain organic acids and finally CO2presumably based on photo-
kolbe reaction [25].

3.7. Photocatalytic degradation mechanism of N-doped MgSb2O6


The plausible photocatalytic degradation mechanism of N-doped MgSb2O6 is proposed and
shown in Figure 9. The band gap values of nanocrystalline disordered N-doped MgSb2O6 and
microcrystalline MgSb2O6were about 3.45 and 4.05 eV respectively. Thus, these compounds
could generate electron-hole pairs on activation by UV light which are responsible for
photodegradation of pollutants. Upon irradiation, the photogenerated electron from valence
band (VB) of N-doped MgSb2O6 excites to conduction band (CB) to react with the dissolved
oxygen in the solution adsorbed on the surface of the photocatalyst to yield •O2− ,while h+
reacts with the OH−to generate •OH [22]. Furthermore, the radical quantification and ESR
results indicated that the easy charge transfer resulted in improving the number of holes in
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 13

Figure 9. Possible mechanism for photocatalytic degradation by MgSb2O6.

the VB of N-doped MgSb2O6to generate more •OH and O2•− which further resulted for the
degradation of p-BP into smaller molecules.

4. Conclusions
In conclusion, nanocrystalline trirutile N-doped MgSb2O6was prepared and its photoca­
talytic activity was reported for the first time. Products prepared with different oxidant to
fuel ratios were in nanocrystalline single phasic N-doped MgSb2O6. Band gap analysis
revealed that the decrease in band gap was observed with increase in O/F. In photo­
catalytic activity, 96% of degradation was achieved in degrading p-BP using nanocrystal­
line N-doped MgSb2O6 after 60 min. In addition, radical quantification results and electron
spin resonance (ESR) analysis revealed that •OH and •O2− were the main ROS responsible
for photodegradation of p-BP. Thus, this study reported the preparation and improved
photocatalytic activity of nanocrystalline N-doped MgSb2O6 for the first time.
14 A. NAGARAJAN AND S. NARAGINTI

Disclosure statement
No potential conflict of interest was reported by the author(s).

References
[1] D. Zhu and Q. Zhou, Environ. Nanotechnol. Monit. 12, 100255 (2019). L. Chen, J. Tang, L.N.
Song, P. Chen, J. He, C.T. Au, S.F. Yin, Appl. Catal. B. Environ242, 379-388 (2019). doi:10.1016/j.
enmm.2019.100255.
[2] R. Shwetharani, H.R. Chandan, M. Sakar, G.R. Balakrishna, K.R. Reddy and A.V. Raghu, Int.
J. Hydrog. Energy 45, 18289 (2020). doi:10.1016/j.ijhydene.2019.03.149.
[3] J. Liu, N. Ma, W. Wu and Q. He, Chem. Eng. J. 393, 124719 (2020). doi:10.1016/j.
cej.2020.124719.
[4] Y. Zheng, L. Cao, G. Xing, Z. Bai, J. Huang and Z. Zhang, RSC.Adv. 9, 7338 (2019). H. Xue, X. Lin,
Q. Chen, Q. Qian, S. Lin, X. Zhang,D.P. Yang,L. Xiao, Nanoscale Res. Lett.13, 114 (2018).
[5] S. Matsushima, T. Tanizaki, H. Nakamura, M. Nonaka and M. Arai, Chem. Lett. 30, 1010 (2001).
N. Kikuchi, H. Hosono and H. Kawazoe, J. Am. Ceram. Soc.88, 2793-2797 (2005); J. Jang, S.J.
Kim, Jpn. J. Appl. Phys.5, 10NE23-1−10NE23-4 (2012); X. Lin, J. Wu, X. Lu, Z. Shan, W. Wang and
F. Huang, Phys. Chem. Chem. Phys.11, 10047-10052 (2009); S. Uma, J. Singh and V. Thakral,
Inorg. Chem.48, 11624-11630 (2009); D. Larcher, A. S. Prakash, L. Laffont,a M. Womes,b
J. C. Jumas, J. Olivier-Fourcade, M. S. Hedge, J.-M. Tarascon, J. Electrochem. Soc. 153,
A1778-A1787(2006). doi:10.1246/cl.2001.1010.
[6] H. Mizoguchi and P.M. Woodward, Chem. Mater. 16, 5233 (2004). doi:10.1021/cm049249w.
[7] K.L. Zhang, X.P. Lin, F.Q. Huang and W.D. Wang, J. Mol. Catal. A. Chem. 258, 185 (2006). Q. You,
Y. Fu, Z. Ding, L. Wu, X. Wang, Z. Li, Dalton Trans. 40,5774-5780 (2011). doi:10.1016/j.
molcata.2006.05.044.
[8] N. Arunkumar and R. Vijayaraghavan, RSC Adv. 4, 65223 (2014). K.C. Patil, Bull. Mater. Sci.16,
533-541 (1993);K. Boobalan, R. Vijayaraghavan, K. Chidambaram, U.M. KamachiMudali, B. Raj,
J. Am. Ceram. Soc.93, 3651–3656 (2010).
[9] H. Haeuseler, SpectrochimicaActa. 37A, 487 (1981). S. Bahfenne, R.L. Frost, Appl. Spectrosc.
Rev.45, 101-129 (2010); Y. Lv, Y. Liu, Y. Zhu, Y. Zhu, J. Mater. Chem. A. 2, 1174-1182 (2014).
doi:10.1016/0584-8539(81)80036-0.
[10] H. Li, L. Song, H. Liu, J. Li, A. Yang, C. Sun, R. Li, Y. Fu and C. Yu, Ceram. Int. 45, 7894 (2019),
Q. You, Y. Fu, Z. Ding, L. Wu, X. Wang, Z. Li, Dalton Trans.40, 5774-5780 (2011);O. E. Linarez
Pérez, M. D. Sánchez, M. L. Teijelo, J. Electroanal. Chem.645, 143-148 (2010).
[11] J. Wang, Z. Wang, B. Huang, Y. Ma, Y. Liu, X. Qin, X. Zhang and Y. Dai, ACS Appl. Mater.
Interfaces. 4 (4024–4030), (2012). H. Tan, Z. Zhao, W. Zhu, E.N. Coker, B. Li, M. Zheng, W. Yu,
H. Fan, Z. Sun, ACS Appl. Mater. Interfaces6, 19184-19190 (2014).
[12] S.G. Krishnan, M. Harilal, A. Yar, B.L. Vijayan, J.O. Dennis, M.M. Yusoff and R. Jose, Electrochim.
Acta 243, 119 (2017). doi:10.1016/j.electacta.2017.05.064.
[13] W. Qian, P.A. Greaney, S. Fowler, S.K. Chiu, A.M. Goforth and J. Jiao, ACS Sustainable Chem.
Eng 2, 1802 (2014). doi:10.1021/sc5001176.
[14] P.N. Kumar, A. Das and M. Deepa, J. Alloys Compd. 832, 154880 (2020).
[15] G. Yang, Z. Jiang, H. Shi, T. Xiao and Z. Yan, J. Mater. Chem. 20, 5301 (2010). C.W. Wang, W.D.
Zhu, J.B. Chen, X.H. Xu, Q. Zhang, Y. Li, J. Wang, F. Zhou, Thin Solid Films556, 440-446 (2014);
A. Bjelajac, R. Petrović, M. Popović, Z. Rakočević, G. Socol, I.N. Mihailescu, D. Janaćković, Thin
Solid Films692, 137598 (2019). doi:10.1039/c0jm00376j.
[16] P. Kubelka and F. Munk, Z. Tech. Phys. (Leipzig). 12, 593 (1931). R.O. Yathisha, Y.A. Nayaka,
Inorg. Chem. Commun.115, 107877 (2020).
[17] L. Xiao, S. Zhang and J. Huang, Powder Technol. 258, 297 (2014). F. Zhou, H. Song, H. Wang,
S. Komarneni, C. Yan, Appl. Clay Sci.166, 9-17 (2018).
[18] P.P. Sahoo and P.A. Maggard, Inorg. Chem. 52, 4443 (2013). doi:10.1021/ic302649s.
INTERNATIONAL JOURNAL OF ENVIRONMENTAL ANALYTICAL CHEMISTRY 15

[19] H. Irie, Y. Watanabe and K. Hashimoto, J. Phys. Chem. B 107, 5483 (2003). doi:10.1021/
jp030133h.
[20] S. Wu, G. Li, Y. Zhang and W. Zhang, Mater. Res. Bull. 48, 1117 (2013). D.P. Dutta, A. Ballal,
A. Singh, M. H. Fulekar, A. K. Tyagi, Dalton Trans.42, 16887-16897 (2013).
[21] S.A. Bakar, G. Byzynski and C. Ribeiro, J. Alloys Compd. 666, 38 (2016). D. Wu, L. Wang, Appl.
Surf. Sci.271, 357-361 (2013). doi:10.1016/j.jallcom.2016.01.112.
[22] H. Huang, K. Liu, K. Chen, Y. Zhang, Y. Zhang and S. Wang, J. Phys. Chem. C 118, 14379 (2014).
doi:10.1021/jp503025b.
[23] K.I. Ishibashi, A. Fujishima, T. Watanabe and K. Hashimoto, Electrochem. Commun. 2, 207
(2000). S. Naraginti, Y. Li, Y. Wu, RSC Adv.6, 75724–75735 (2016).
[24] T. Hirakawa and Y. Nosaka, Langmuir. 18, 3247 (2002). S.Naraginti, Y.C. Yong, Ecotoxicol.
Environ. Saf.170, 355–362 (2019). doi:10.1021/la015685a.
[25] J.M. Herrmann, Catal. Today. 53, 115 (1999). W. Bian, X. Song, D. Liu, J. Zhang, X. Chen,
J. Hazard. Mater.192, 1330–1339 (2011). doi:10.1016/S0920-5861(99)00107-8.

You might also like