You are on page 1of 14

DE GRUYTER International Journal of Chemical Reactor Engineering.

2018; 20170136

K Sathish Kumar1 / K R Rohit Narayanan1 / S Siddarth1 / R Pavan Kumar1 / R Badri Narayan1 /


R Goutham1 / V Samynaathan1

Synthesis of MgO/TiO2 Nanocomposite and Its


Application in Photocatalytic Dye Degradation
1 Department of Chemical Engineering, Sri Sivasubramaniya Nadar College of Engineering, Rajiv Gandhi Salai (OMR),

Kalavakkam 603110, Tamil Nadu, India, E-mail: sathishkumark@ssn.edu.in

Abstract:
The application of nanoparticles in dye degradation is one of the trending arenas of research in the present
day world. In this work, we report a novel route to synthesise MgO/TiO2 metal oxide nanocomposite by mi-
croemulsion technique and its application in photocatalytic dye degradation. Oil in water microemulsion was
prepared using Span 80 and Tween 80 as surfactants whose proportion was regulated using hydrophilic and
lipophilic Balance (HLB). The obtained microemulsion was then mechanically stirred and calcined to obtain the
nanocomposite. The as-prepared nanocomposites were characterized using X Ray Diffraction (XRD), Fourier
transform infrared spectroscopy (FTIR) and scanning electron microscopy (SEM). The prepared nanoparticles
exhibited remarkable potential to degrade azo dye (Methyl red) under UV-Visible light irradiation. The effect
of catalyst in the degradation was studied for different concentrations of dye (20, 40, 60, 80, 100 mg/L) and
different loadings of the catalyst (0.4, 0.6, 0.8, 1.0, 1.2 g/L) so as to determine the optimum catalyst load. The
consistency of the obtained data was compared with the first order reaction rate expression. Quasi steady state
model was used in fitting the data and the kinetic constants were evaluated. Also, the degradation efficiency of
MgO/TiO2 nanocomposite was compared with the degradation efficiency of TiO2 nanoparticles synthesized
by microemulsion method.
Keywords: nanocomposite, microemulsion, micelles, quasi steady state model, photocatalytic dye degradation,
azo dye
DOI: 10.1515/ijcre-2017-0136
Received: July 10, 2017; Revised: May 9, 2018; Accepted: May 13, 2018

1 Introduction
These days, numerous industries are faced with problems involving the treatment and disposal of effluent
streams containing organic compounds, some of which are well known environmental pollutants. The drainage
of these materials in aqueous environment leads to a decrease in dissolved oxygen levels, which in turn has ir-
revocable damage on aquatic ecosystem (Mandal et al. 2010; Mittal et al. 2009; Saravanan et al. 2015; Srikanth et
al. 2017). In case of textile and dyeing industries, these pollutant compounds manifest themselves in the form of
dyestuffs/pigments. These compounds present the challenge of being stable and persisting in the environment
over longer periods of time (Muhd Julkapli, Bagheri, and Bee Abd Hamid 2014; Van Driel et al. 2016; Adamek
et al. 2013; Channei et al. 2014; Chen et al. 2017). Additionally, the recalcitrant nature of such dyestuffs or their
intermediates produced in due course of their mineralisation in environment can have a negative and serious
effect on the environment owing to their potential mutagenic, carcinogenic and toxic properties (Narayan et
al. 2016). This provides the rationale for developing strategies for removal dyestuffs from effluents emanating
from industries before they can be discharged. Methyl Red (CAS number 493–52–7), is a commercial dye, that
is largely used in textile industry. Nonetheless, it is regarded as a persistent organic pollutant (Zhang and Gao
2013; Yang et al. 2015). Due to its thiazine structure, it is a representative member of the class of organic com-
pounds that are stable to oxidation. Additionally, the presence of large concentrations of MR dyes in surface
waters can inhibit photosynthesis by light absorption (Izawa 1962). Till date, a large number of methods have
been developed to remove organic contaminants from aquatic media, including conventional water treatment
methods namely—biological treatment and physical treatment (which is generally achieved by sorption over
activated carbon) (Narayan et al. 2016). The disadvantage of employing biological treatment is that, the pres-
ence of non-biodegradable and toxic intermediate compounds in aquatic media can be recalcitrant for biological
systems or may increase the time to mineralise these compounds (Kurt, Apaydin, and Gonullu 2007; Jorfi et al.

K Sathish Kumar is the corresponding author.


© 2018 Walter de Gruyter GmbH, Berlin/Boston.

1
Sathish Kumar et al. DE GRUYTER

2017; Saleh and Gupta 2014). In case of adsorption, the use of expensive commercial activated carbons or spe-
cially prepared adsorbents is not economically feasible for industrial level applications. This motivates to look
out for advanced oxidation processes (AOPs) as a viable alternative for treatment of pollutants from wastewa-
ters. Photocatalysis as a AOP is attractive in this regard, as the reactions can be activated by using inexpensive
illumination sources in presence of suitable quantities of nano-photocatalysts. Photocatalysis works to com-
pletely degrade organics into CO2 , H2 O, NH3 , mineral salts and acids that can be easily discharged without
causing much of harm to the ecosystem (Khan, Adil, and Al-Mayouf 2015). Micro-emulsions are a novel means
to synthesise ultrafine particles in view of the capacity to control the size of resultant nanoparticles and inhibit
their aggregation (Lu, Wu, and Kale 2008). In this work, water-in-oil (W/O) microemulsion was employed for
preparing the catalysts. This microemulsion technique can be employed to produce very fine particles of cata-
lysts, which are efficient for photocatalytic processes, which is largely depend on surface morphologies of the
catalyst particles (Yu et al. 2003; Martin Andersson, Lars Österlund, and Sten Ljungström, Anders Palmqvist
2002). In this study, MgO-TiO2 nanoparticles in the range of 0.63 nm and band gap energy of ~4.3 eV (Bayal
and Jeevanandam 2014; Bandara, Hadapangoda, and Jayasekera 2004) were synthesized using microemulsion
technique. The aim of this work was to determine the applicability of the as synthesised nanoparticles as pho-
tocatalysts for wastewater remediation.

2 Materials and methods


2.1 Chemicals and reagents

Analytical grade titanium isopropoxide, magnesium sulphate, Span 80 and Tween 80, n-pentanol and n-octane
were procured from Sigma Aldrich and used as such without further processing. Other reagents used include
ammonium hydroxide, ammonia solution and acetone. Methyl red dye purchased from Sigma Aldrich was
used as a representative azo dye for degradation studies.

2.2 Instrumentation

The Fourier Transfom Infra-Red Spectroscopy (FTIR) analysis was done using Perkin Elmer 100S FTIR at Na-
tional institute of Technology, Warangal. KBr pelletisation of the sample was done before the analysis to fa-
cilitate sharp peak formation and reduction of noise. The X- Ray Diffraction Analysis (XRD) was done using
Bruker KAPPA APE XII at Indian Institute of Technology, Madras. Cu Kα radiation and high speed wide angle
lynx eye detector for fast collection of data were used for the analysis. The Scanning Electron Microscopy Anal-
ysis was done using the COXEM CX 200 TM scanning electron microscope (SEM) at Microlabs, Chennai. The
range of magnification of SEM was 30× to 300000× Magnification with SE (Secondary Electron) & BSE (Back
scattered Electron) imaging. The amount of degradation of the dye was measured by using UV Visible Spec-
troscopy by measuring the absorbance values of different samples. The characterisation was done using Jasco
V-630 UV-Visible spectrophotometer at Department of Chemical Engineering, SSN College of Engineering.

2.3 Synthesis of TiO2 nanoparticles

The titanium oxide nanoparticles were prepared by microemulsion technique. Firstly the microemulsion was
prepared by mixing 2 mL of titanium iso-propoxide with 10 mL water to get a solution of 0.7 mol/L Ti4+ .
About 7 mL of Span 80 and 7.3 mL of Tween 80 was added to the contents of the beaker in order to maintain
49:51 ratio of Span 80 to Tween 80. 10 mL of n-pentanol and 10 mL of n-octane were added as solvents and
co-solvents respectively. In another beaker, a buffer solution of 2 grams NH4 HCO3 and 3 mL NH3 was mixed
and the same amounts of Span 80 and Tween 80, n-pentanol and n-octane were added to the mixture (Parvathi
V et al. 2016; Sanchez-Dominguez et al. 2015). In the second step, the two solutions were mixed and stirred at a
constant temperature of 25o C until the mixed solution turned from clear to turbid and after aging for 12 hours
the microemulsion was broken up with 5 mL acetone. The precipitate obtained was washed three times with
alcohol, and then dried at 60o C for 8 h to obtain the nanoparticles (Bethi et al. 2016; Bickley et al. 1991; Zhang
et al. 2016).

2
DE GRUYTER Sathish Kumar et al.

2.4 Synthesis of MgO-TiO2 nanocomposite

The experiment was conducted using titanium iso-propoxide because of its ability to form TiO2 when reacted
with water. The first step was the formation of microemulsion. In one beaker 1.19 grams of MgSO4 was taken
and dissolved in 10 mL water to get a solution of 1 mol/L Mg2+ . About 2 mL of Titanium iso-propoxide was
taken with a dropper and mixed with 10 mL water to get a solution of 0.7 mol/L Ti4+ . This was mixed with
the Mg2+ solution. The ratio of Mg2+ to Ti4+ maintained was 5:3. About 7 mL of Span 80 and 7.3 mL of Tween
80 was added to the contents of the beaker in order to maintain 49:51 ratio of Span 80 to Tween 80. 10 mL of
n-pentanol and 10 mL of n-octane were added as solvents and co-solvents respectively. In another beaker, a
buffer solution of 2 grams NH4 HCO3 and 3 mL NH3 was mixed and the same amounts of Span 80 and Tween
80, n-pentanol and n-octane were added to the mixture. In the second step, the two solutions were mixed and
stirred at a constant temperature of 25o C until the mixed solution turned from clear to muddy, and after aging
for 12 hours the microemulsion is broken up with 5 mL acetone. The precipitate obtained was washed three
times with alcohol, and then dried at 70o C for 12 h to obtain the nanocomposite (Fu et al. 2015; Juma et al. 2017;
Tan et al. 2011).

2.5 Photocatalytic dye degradation

Different concentrations of methyl red dye samples were prepared (20, 40, 60, 80 and 100 mg/L). A calibration
curve between absorbance and concentration was drawn. 200 mL of the 100 mg/L dye sample was taken in the
photocatalytic reactor. As per literature review (Lachheb et al. 2002; Pandit et al. 2015; Kaushik and Malik 2009),
the optimized 1 g/L catalyst loading was taken and accordingly calculated for 200 mL sample . The UV lamp
was switched on and magnetic stirrer was set at 1200 rpm. For every 10 minutes 5 ml sample was taken from
photocatalytic reactor as shown in Figure 1.

Figure 1: Schematic representation of photocatalytic reactor.

This procedure was repeated till 60 minutes. For the samples taken absorbance of Methyl red is noted at
410 nm. The above procedure was repeated for different concentrations (80, 60, 40 and 20 mg/L). Using the
absorbance data and the calibration curve, percentage degradation was calculated. For a given concentration,
different catalyst loading was tried out (i. e. 0.4, 0.6, 0.8, 1.0, 1.2 g/L) and the above procedure was repeated for
that concentration. Using this, the optimized catalyst loading was calculated for that concentration.

2.6 Chemical Oxygen Demand (COD) analysis

COD analysis was performed by the standard open reflux method. About 5 mL of the sample was drawn from
the reaction mixture and was diluted suitably for the analysis. Reduction in the COD level of the end sample
proved that the dye molecules were degraded. The amount of COD removed after degradation was expressed
as %COD removal.

3
Sathish Kumar et al. DE GRUYTER

3 Results and discussion


3.1 FTIR analysis of TiO2 nanoparticle

FTIR analysis of TiO2 is shown in Figure 2. A broad band associated with the characteristic vibrational modes
of TiO2 appeared between 880 and 380 cm−1 for the sample. Ti–O–Ti bond vibrations were found at 547.91
cm−1 .This large band appeared mainly due to nanoparticles with large size distribution. The nanoparticle shape
and state of aggregation of the nanocrystals can modify the infrared powder spectra inducing shifts, enlarge-
ments and superposition of vibrational modes.

3.2 XRD analysis of TiO2 nanoparticle

The X Ray Diffraction Analysis of TiO2 nanoparticle was carried out to confirm the presence of titanium dioxide
and also to determine the size of the unit cell. The XRD Analysis is as shown in Figure 3.

Figure 2: FTIR Analysis of TiO2 nanoparticle.

Figure 3: XRD Analysis of TiO2 nanoparticle.

The XRD pattern of TiO2 showed well defined peaks at 2θ values of 25o , 37o and 48o indicating the presence
of titanium oxide. The peak values corresponded well with the standard JCPDS File No #21–1272 for titanium
oxide nanoparticle. The broad diffraction lines in the XRD pattern of TiO2 related to the nanocrystalline anatase
phase. The width reduction of the peaks associated to an increase in the crystallite size. The unit cell Size
determination of the TiO2 nanoparticle was done by using the standard h,k and l values for the corresponding
peaks. For TiO2 nanoparticle 25°, 37°, 48° peaks had the (h,k,l) coordinates of (1 0 1), (0 0 4), (2 0 0) respectively.
The unit cell size (α) was determined as 0.6312 nm by using Bragg’s law and the symmetry was predicted as
cubic. Using this data Derby –Scherer Formula was used to find the size of the prepared nanoparticle. The
evaluation yielded a size of 42 nm which confirmed that the dimensions were in the nanoscale.

3.3 SEM analysis of TiO2 nanoparticle

The SEM analysis of TiO2 nanoparticles was done to define the surface morphology of the nanoparticles. As
shown in Figure 4, the nanoparticles of titanium oxide have an irregular morphology. An increased magnifica-
tion showed a spherical shape for the nanoparticles. The images corresponded well with the SEM analysis of
TiO2 prepared by sol-gel method.

4
DE GRUYTER Sathish Kumar et al.

Figure 4: (a–d). SEM Analysis of TiO2 nanoparticles.

3.4 FTIR analysis of MgO-TiO2 nanocomposite

FTIR spectrum of MgO/TiO2 nanocomposite was taken and analysed as shown in Figure 5. FTIR analysis was
done in the range of 500 cm−1 to 4500 cm−1 . Almost equally spaced bonds are shown in the picture. The O-
H stretching vibration bonds represented in the range of 4000 cm−1 to 3400 cm−1 are due to the presence of
water molecules. From 3400 cm−1 to around 1800 cm−1 there is not any significant bond which characterizes
the desired composite. At 1088 cm−1 the peaks obtained with the effect of C = O bond is due to the possible
stretching bond of Ti-O-Mg. The strong Mg-O stretching bond is observed at 700 cm−1 .

Figure 5: FTIR Analysis of MgO/TiO2 nanocomposite.

3.5 XRD analysis of MgO-TiO2 nanocomposite

X Ray Diffraction Analysis plays a pivotal role in determining the unit cell dimensions and also in confirming
the presence of the desired compounds. Figure 6 shows the XRD analysis of prepared MgO-TiO2

Figure 6: XRD Analysis of MgO/TiO2 nanocomposite.

The XRD diffraction pattern of MgO-TiO2 shows well defined peaks indicating the presence of MgO, TiO2
and magnesium substituted Titania compounds. The XRD pattern of MgO indicates peaks at 2θ values of 36o,
43o and 62o which are in well accordance with the standards. The peaks for rutile and anatase TiO2 appeared
at 2θ values of 23.71o , 24.93o , 29.89o and 31.05o respectively. The peaks at 28.58o and 40.00 indicate the pres-
ence of magnesium substituted titania compound. This substituted compound (MgTiO3 ) was formed by the
substitution of titanium ions by the magnesium ions in the crystal lattice of TiO2 during composite formation.
The high intensity peaks at the start of the diffraction pattern indicates that the composite consists of array of

5
Sathish Kumar et al. DE GRUYTER

nano-crystalline grains. The determination of unit cell dimensions involves the calculation of miller indices of
the crystal lattice and substituting in the diffraction law equation. The miller indices obtained conformed to
the values of indices in the standard JCPDS files for MgO and TiO2 . The low intensity peaks at higher values
of 2θ shows the amorphous and coarse grain nature of the compound. The average unit cell dimension was
calculated as 1.0190 nm using Bragg’s law and the size was calculated as 23 nm using Derby-Scherer Formula.
Likewise, the bandgap of the prepared MgO/TiO2 nanocomposites were found to be around 4.3 eV using pho-
toluminesence (PL) spectroscopy.

3.6 SEM analysis of MgO-TiO2 nanocomposite

Different magnifications of SEM image were obtained as shown in Figure 7. A uniform dispersion of MgO and
TiO2 were observed. Comparing with MgO-TiO2 synthesised by sol-gel method yielded similar results. 100K×
magnification clearly indicated the dispersion of mixed oxide nanoparticles which showed its amorphous na-
ture. The images confirmed the fact that nanoparticles prepared through micro-emulsion have a defined mor-
phology compared to other methods of synthesis.

Figure 7: (a–d). SEM Analysis of MgO-TiO2 nanocomposite.

3.7 Dye degradation studies

3.7.1 Dark reaction

To verify the significance of UV light in photocatalytic reactions, the study was performed at dark conditions in
the absence of any light source. 200 mL of 100 mg/L dye solution was taken in the photocatalytic reactor. Cata-
lyst was loaded at 1 g/L and the whole reaction was carried out without switching on the UV lamp at absolute
dark conditions inside the reactor for 1 hour. Even after such a long period of time, there was no significant
change in the concentration of the sample. This clearly indicates that there was no significant degradation in
the absence of UV light. Hence the reaction proceeds only in the presence of UV light and the degradation of
the dye is only through photocatalysis.

3.7.2 Effect of concentration of dye solution

The effect of concentration of MR was studied to predict the efficiency of the catalyst as shown in Figure 8.
It was observed that the amount of dye degraded increases with increase in concentration of dye solution.
For example, it was observed that 100 mg/L dye degrades to 25 mg/L after 50 minutes whereas 20 mg/L dye
became 3 mg/L in the same time span. Also, the characteristic curves decreased in the following order: 100, 80,
60, 40, 20 mg/L. For any given time, the concentration corresponding to an initial dye concentration of 100 mg/L
was higher than that of 80, 60, 40 and 20 mg/L and this began to converge as time increased.

6
DE GRUYTER Sathish Kumar et al.

Figure 8: Effect of concentration of dye solution.

Also the degradation efficiencies were calculated for various concentrations at different time. Graphs were
plotted between percentage degradation versus time and amount of dye degraded versus time as shown in
Figure 9. It was observed that even though the amount of dye degraded for 100 mg/L was more than that of 80,
60, 40, 20 mg/L yet the degradation efficiency of 100 mg/L was found to be the lowest. Degradation efficiency of
20 mg/L was found to be the highest. Even though there is not much difference in the percentage degradation
curve for different concentrations, yet the difference increases as time increases.

Figure 9: Degradation efficiency for different concentrations.

The photo-degradation reaction was further proved by performing COD analysis for the given range of
initial concentrations of MR (20, 40, 60, 80 and 100 mg/L). Figure 10 shows the COD removal efficiency for
different initial concentrations of MR It can be seen that the COD removal efficiency decreased when the initial
concentration of MR samples were increased. The maximum COD removal of 75 % was achieved at an initial MR
concentration of 20 mg/L and the minimum COD removal of 61 % was achieved at an initial MR concentration
of 100 mg/L.

Figure 10: COD removal data for different initial dye concentrations.

7
Sathish Kumar et al. DE GRUYTER

3.7.3 Effect of catalyst loading

The effect of catalyst loading was studied to find the optimal loading at which the maximum degradation
occurs. From Figure 11, the concentration of the dye was found to decrease exponentially with increase in
loading of the photocatalyst. The concentration curve at 1.2 g/L almost saturated to the value at 1.0 g/L which
showed that at 1.0 g/L was the optimised amount of loading of the catalyst.
The degradation efficiency for different loading of the catalyst was studied and was shown in Figure 12. The
degradation efficiency curve for 1.0 g/L increased exponentially above all other amounts of catalyst loading
which indicated that the degrading efficiency was the maximum for 1.0 g/L loading of the photocatalyst. The
experimentation with 1.2 g/L yielded a lesser degradation efficiency than 1.0 g/L confirming that the latter
was the optimum amount of catalyst loading. The reason is because that the increased photocatalytic activity
with increase in photocatalysts mass happens, which may be due to increasing accessibility of photocatalyst
sites and the diminish of catalytic activity after the saturation region is correlated to rising of light scattering
by the much excess of photocatalyst particles.
Figure 13 shows the COD removal efficiency for different catalyst loading values. As expected from the
above results, when the concentration of catalyst increased the COD removal also increased till a concentration
value of 1.0 g/L. When the loading was increased beyond 1.0 g/L, the %COD removal decreased ensuring that
the optimum catalyst loading value was 1.0 g/L. From the figure the maximum COD removal of 75 % was
obtained at 1.0 g/L catalyst loading value and the minimum COD removal of 67 % was obtained at 0.4 g/L
catalyst loading value.

Figure 11: Effect of catalyst loading.

Figure 12: Degradation efficiency for different loading of the catalyst.

8
DE GRUYTER Sathish Kumar et al.

Figure 13: COD removal data for different catalyst loading values.

3.7.4 Model fitting

For a first order reaction

c
In ( ) = - kt
c0

Hence, plotting a graph between ln(C/C0 ) and time by setting intercept as zero, yielded a linear plot. Figure 14
shows the First order kinetics for various dye concentrations. The slope of the linear plot gives us the numerical
value of rate constant for a given initial concentration. It was also found that coefficient of determination (R2 )
value for the linear plots is above 0.97 indicating that linear fit to be a good fit. Hence, the degradation of dye
of different concentration using MgO/TiO2 can be modelled as a first order reaction

3.7.5 Kinetic model for photocatalytic degradation of methyl red

During photocatalysis of Methyl red (MR), the dye molecules decomposes as per the following mechanism

𝐾𝑀𝑅
𝑀𝑒𝑡ℎ𝑦𝑙𝑟𝑒𝑑 + 𝑂𝐻 • → 𝐸𝑛𝑑 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 (1)

Being a batch process, the system can be considered as a closed system. Therefore quasi-steady-state assumption
can be used to illustrate the kinetics of photocatalysis of Methyl red. Hence,

𝑑[𝑑𝑦𝑒]
𝑟𝑀𝑅 = − (2)
𝑑𝑡

𝑟𝑀𝑅 = 𝑘𝑀𝑅 .[𝑑𝑦𝑒].[𝑂𝐻 • ] (3)

Where [dye], OH• and KMR are Methyl red concentration, hydroxyl radical concentration and rate constant
respectively. Concentration of hydroxyl radicals present mainly depends upon the following three reactions,

1. Photocatalytic dissociation of water

𝑘𝑑
𝐻2 𝑂 → 𝑂𝐻 • + 𝐻 • (4)

2. Formation of hydrogen peroxide

𝑘𝑟
𝑂𝐻 • + 𝑂𝐻 • ⇔ 𝐻2 𝑂2 (5)
𝑘𝑓

3. Possible side reaction

9
Sathish Kumar et al. DE GRUYTER

𝑘1
𝐻2 𝑂2 + 𝑂𝐻 • → 𝐻𝑂•2 + 𝐻2 𝑂 (6)

Where kd , kf , kr and k1 are rate constants.


Mass balance equations for [H2 O2 ] and [OH• ] are given as

𝑑[𝐻2 𝑂2 ]
= 𝑘𝑓 [𝑂𝐻 • ]2 − 𝑘𝑟 [𝐻2 𝑂2 ] − 𝑘1 [𝐻2 𝑂2 ][𝑂𝐻 • ] (7)
𝑑𝑡

𝑑[𝑂𝐻 • ]
= 𝑘𝑑 − 𝑘𝑓 [𝑂𝐻 • ]2 + 𝑘𝑟 [𝐻2 𝑂2 ] − 𝑘1 [𝐻2 𝑂2 ][𝑂𝐻 • ] − 𝑘𝑀𝑅 [𝑑𝑦𝑒][𝑂𝐻 • ] (8)
𝑑𝑡
Under quasi steady state assumption, Eq.A8 becomes,

𝑘𝑑 − 𝑘𝑓 [𝑂𝐻 • ]2 + 𝑘𝑟 [𝐻2 𝑂2 ] − 𝑘1 [𝐻2 𝑂2 ][𝑂𝐻 • ] − 𝑘𝑀𝑅 [𝑑𝑦𝑒][𝑂𝐻 • ] = 0 (9)

𝑘𝑑
[𝑂𝐻 • ] = (10)
𝑘1 [𝐻2 𝑂2 ] + 𝑘𝑀𝑅 [𝑑𝑦𝑒]

Substituting Eq.A10 in Eq.A3,

𝑘𝑑 .𝑘𝑀𝑅 .[𝑑𝑦𝑒]
𝑟𝑀𝑅 = (11)
𝑘1 [𝐻2 𝑂2 ] + 𝑘𝑀𝑅 [𝑑𝑦𝑒]

𝑘𝑞 [𝑑𝑦𝑒]
𝑟𝑀𝑅 = (12)
1 + 𝐾𝑞 [𝑑𝑦𝑒]

Where,

𝑘𝑑 𝑘𝑀𝑅
𝑘𝑞 =
𝑘1 [𝐻2 𝑂2 ]

𝑘𝑀𝑅
𝐾𝑞 = .
𝑘1 [𝐻2 𝑂2 ]

The rate equation given by Eq. (A12) was found to fit the experimental data well and the results are shown
in Figure 15. Kinetic parameters kq and Kq values were obtained with MATLAB 7.12 and values are 0.01967
min−1 and 0.001864 L/mg, respectively. The kinetic equation for photocatalytic degradation of Methyl red is
given as

0.01967[𝑑𝑦𝑒]
𝑟𝑀𝑅 =
1 + 0.001864[𝑑𝑦𝑒]

Figure 14: First order kinetics fit for effect of dye concentration.

10
DE GRUYTER Sathish Kumar et al.

Figure 15: Kinetic model for photocatalytic degradation of Methyl Red.

Also the first order kinetics fit for different catalyst loading was studied. Similar to the fitting procedure
adopted for different concentrations of the dye, the kinetic model fit for different catalyst loading was done as
shown in Figure 16.

Figure 16: First order kinetics fit for effect of catalyst loading.

The plot of ln(C/C0 ) and time by setting intercept as zero yielded a linear plot indicating the kinetics to be of
first order. The slope of the linear plot gave the numerical value of rate constant for a given initial concentration.
It was also found that coefficient of determination (R2 ) value for the linear plots was above 0.97 indicating that
linear fit to be a good fit. Hence, the degradation of dye of different catalyst loading using MgO/TiO2 was
modelled as a first order reaction.

3.7.6 Comparative degradation studies of MgO-TiO2 nanocomposite and TiO2 nanoparticle

The degradation of methyl red was carried out using the prepared Titanium oxide Nanoparticles. Figure
17.shows the plot of concentration of the dye with time for MgO –TiO2 Nanocomposite and TiO2  Nanoparticle
at a dye concentration of 20 mg/L and a catalyst

11
Sathish Kumar et al. DE GRUYTER

Figure 17: Concentration-time graph comparison for MgO-TiO2 nanocomposite and TiO2 nanoparticle.

loading of 1 g/L. It was observed that the concentration of dye decreased as time increased for both mixed
oxide and single oxide. The decrease appears to be an exponential decay. The decrease in concentration of
methyl red in presence of single oxide was lower when compared in the presence of mixed oxide. Or in other
words, the concentration profile of methyl red in presence of single oxide was above that of mixed oxide. This
proves that better degradation of methyl red was achieved using mixed oxide.
Also it was found that the degradation efficiency and amount of dye degraded as shown in Figure 18 in-
creased with increase in time for both single oxide and mixed oxide catalysed degradation. It was observed
that in a time span of 1 hour, the degradation efficiency achieved using a mixed oxide was more than that of
single oxide. It can also be inferred that amount of methyl red degraded by the mixed oxide was more than that
of single oxide. The first order model fit was compared for the single and mixed oxide as shown in Figure 19.
It was inferred that linear was a good fit for the given data points as it had a high coefficient of determination
(R2 ) value.

Figure 18: Percentage Degradation-time graph comparison for MgO-TiO2 nanocomposite and TiO2 nanoparticle.

Figure 19: First Order kinetics fit comparison for MgO-TiO2 nanocomposite and TiO2 nanoparticle.

Rate constants were found as the slope of the line. The rate constants for methyl red degradation in presence
of single oxide and mixed oxide were 0.024 min−1 and 0.034 min−1 respectively. This indicated that for a given
concentration (20 mg/L), the rate constant and hence the degradation rate of methyl red was found to be higher
in the presence of mixed oxide when compared with that of single oxide.

3.7.7 Reusability of the synthesised photocatalyst

As we know, it is essential to determine the stability and reusability of a photocatalyst so as to make it econom-
ical for implementation in industrial scale water treatment systems. The circulating runs in the photocatalytic
degradation of methyl red under UV-light irradiation were performed to evaluate these factors. The MgO-TiO2
catalysts were collected after each trial by centrifugation followed by washing in distilled water and drying in
hot air oven. The resultant catalyst particles were then reused under the same experimental conditions. This
methodology was continued for the subsequent cycles and the reduction in the concentration of the dye was

12
DE GRUYTER Sathish Kumar et al.

noted for each and every cycle. From Figure 20, it can be seen there was not a significant decrease in the ef-
ficiency of degradation for the first four cycles. However, when the number of cycles increased beyond four,
there was a significant reduction in the percentage dye removal. This may be attributed to the loss of active
sites on the catalyst surface and fouling of catalyst surface. In summary, this finding demonstrates that MgO-
TiO2 photocatalyst exhibits remarkable stability and reusability without undergoing photo-corrosion during
the photocatalytic reaction, which favors its long-term utilization in the removal of organic pollutants.

Figure 20: Degradation efficiency for different cycles of reusing the photocatalyst.

4 Conclusion
The study involved the use of a mixed oxide nanocomposite for degradation of methyl red dye. The nanocom-
posite synthesized was characterised using FTIR spectroscopy, XRD and SEM. The mixed oxide was found to
have a high degradation efficiency, the maximum being 84.8 % for 20 mg/L in 60 minutes for a catalyst load-
ing of 1 g/L. Quasi steady state model was used in fitting the data for the heterogeneous catalytic reaction.
The photocatalytic activity of the nanocomposite was compared with TiO2 nanoparticles and an increase in
degradation efficiency of 11.3 % was observed.

Funding
This work was supported by SSN Trust, Grant Number: Grant of Indian rupess 25,000.

References
Martin Andersson, Lars Österlund, and Sten Ljungström, Anders Palmqvist. 2002 . “Preparation of Nanosize Anatase and Rutile TiO2 by
Hydrothermal Treatment of Microemulsions and Their Activity for Photocatalytic Wet Oxidation of Phenol.” Journal of Physical Chemistry
106 (41): 10674– 10679 . DOI: 10.1021/JP025715Y.
Adamek, E., W. Baran, J. Ziemiańska, and A. Sobczak. 2013. “The Comparison of Photocatalytic Degradation and Decolorization Processes of
Dyeing Effluents.” International Journal Photoenergy 2013: 1–11. DOI: 10.1155/2013/578191.
Bandara, J., C. Hadapangoda, and W. Jayasekera. 2004. “TiO2/MgO Composite Photocatalyst: The Role of MgO in Photoinduced Charge
Carrier Separation.” Applications Catal B Environment 50, 83–88. doi: 10.1016/J.APCATB.2003.12.021.
Bayal, N., and P. Jeevanandam. 2014. “Synthesis of TiO2−MgO Mixed Metal Oxide Nanoparticles via a Sol−Gel Method and Studies on Their
Optical Properties.” Ceram International 40, 15463–15477. doi: 10.1016/J.CERAMINT.2014.06.122.
Bethi, B., S.H. Sonawane, B.A. Bhanvase, and S.P. Gumfekar. 2016. “Nanomaterials-Based Advanced Oxidation Processes for Wastewater
Treatment: A Review.” Chemical Engineering Processing Processing Intensif 109, 178–189. doi: 10.1016/J.CEP.2016.08.016.
Bickley, R.I., T. Gonzalez-Carreno, J.S. Lees, L. Palmisano, and R.J.D. Tilley. 1991. “A Structural Investigation of Titanium Dioxide Photocata-
lysts, J.” Solid State Chemical 92, 178–190. doi: 10.1016/0022-4596(91)90255-G.
Channei, D., B. Inceesungvorn, N. Wetchakun, S. Ukritnukun, A. Nattestad, J. Chen, and S. Phanichphant. 2014. “Photocatalytic
Degradation of Methyl Orange by CeO2 and Fe–Doped CeO2 Films under Visible Light Irradiation.” Sciences Reports 4, 5757. doi:
http://dx.doi.org/10.1038/srep05757.
Chen, X., Z. Wu, D. Liu, and Z. Gao. 2017. “Preparation of ZnO Photocatalyst for the Efficient and Rapid Photocatalytic Degradation of Azo
Dyes.” Nanoscale Researcher Letters 12, 143. doi: 10.1186/s11671-017-1904-4.
Fu, J.-R., J. Zheng, W.-J. Fang, C. Chen, C. Cheng, R.-W. Yan, S.-G. Huang, and C.-C. Wang. 2015. “Synthesis of Porous Magnetic
Fe3O4/Fe@ZnO Core–Shell Heterostructure with Superior Capability for Water Treatment.” Journal Alloys Compd 650, 463–469. doi:
10.1016/J.JALLCOM.2015.08.065.

13
Sathish Kumar et al. DE GRUYTER

Izawa, S. 1962. “Methylene Blue Inhibition Of Photosynthesis In Rhodopseudomonas Palustris.” Plant & Cell Physiology 3, 43–51. doi:
10.1093/oxfordjournals.pcp.a078941.
Jorfi, S., B. Kakavandi, H.R. Motlagh, M. Ahmadi, and N. Jaafarzadeh. 2017. “A Novel Combination of Oxidative Degradation for Benzotria-
zole Removal Using TiO2 Loaded on FeIIFe2IIIO4@C as an Efficient Activator of Peroxymonosulfate.” Applications Catal B Environment 219,
216–230. doi: 10.1016/J.APCATB.2017.07.035.
Juma, A.O., E.A.A. Arbab, C.M. Muiva, L.M. Lepodise, and G.T. Mola. 2017. “Synthesis and Characterization of CuO-NiO-ZnO Mixed Metal
Oxide Nanocomposite.” Journal Alloys Compd 723, 866–872. doi: 10.1016/J.JALLCOM.2017.06.288.
Kaushik, P., and A. Malik. 2009. “Fungal Dye Decolourization: Recent Advances and Future Potential.” Environment International 35, 127–141.
doi: 10.1016/J.ENVINT.2008.05.010.
Khan, M.M., S.F. Adil, and A. Al-Mayouf. 2015. “Metal Oxides as Photocatalysts.” Journal Saudi Chemical Social 19, 462–464. doi:
10.1016/J.JSCS.2015.04.003.
Kurt, U., O. Apaydin, and M.T. Gonullu. 2007. “Reduction of COD in Wastewater from an Organized Tannery Industrial Region by Electro-
Fenton Process.” Journal of Hazardous Materials 143, 33–40. doi: 10.1016/J.JHAZMAT.2006.08.065.
Lachheb, H., E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard, and J.-M. Herrmann. 2002. “Photocatalytic Degradation of Various Types
of Dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in Water by UV-irradiated Titania.” Applications Catal B
Environment 39, 75–90. doi: 10.1016/S0926-3373(02)00078-4.
Lu, C.-H., W.-H. Wu, and R.B. Kale. 2008. “Microemulsion-Mediated Hydrothermal Synthesis of Photocatalytic TiO2 Powders.” Journal of
Hazardous Materials 154, 649–654. doi: 10.1016/J.JHAZMAT.2007.10.074.
Mandal, T., S. Maity, D. Dasgupta, and S. Datta. 2010. “Advanced Oxidation Process and Biotreatment: Their Roles in Combined Industrial
Wastewater Treatment.” Desalination 250, 87–94. doi: 10.1016/J.DESAL.2009.04.012.
Mittal, A., J. Mittal, A. Malviya, and V.K. Gupta. 2009. “Adsorptive Removal of Hazardous Anionic Dye “Congo Red” from Wastewater Using
Waste Materials and Recovery by Desorption.” Journal Colloid Interface Sciences 340, 16–26. doi: 10.1016/J.JCIS.2009.08.019.
Muhd Julkapli, N., S. Bagheri, and S Bee Abd Hamid. 2014. “Recent Advances in Heterogeneous Photocatalytic Decolorization of Synthetic
Dyes.” TheScientificWorldJournal 2014, 692307. doi: 10.1155/2014/692307.
Narayan, R.B., R. Goutham, B. Srikanth, and K.P. Gopinath. 2016. “A Novel Nano-Sized Calcium Hydroxide Catalyst Prepared from Clam
Shells for the Photodegradation of Methyl Red Dye.” Journal Environment Chemical Engineering. doi: 10.1016/J.JECE.2016.12.004.
Pandit, V.U., S.S. Arbuj, Y.B. Pandit, S.D. Naik, S.B. Rane, U.P. Mulik, S.W. Gosavi, and B.B. Kale. 2015. “Solar Light Driven Dye Degradation
Using Novel Organo–Inorganic (6,13-Pentacenequinone/Tio 2 ) Nanocomposite.” RSC Advances 5, 10326–10331. doi: 10.1039/C4RA11920G.
Parvathi V, P., T. Jaiakumar, M. Umadevi, J. Mayandi, and G.V. Sathe. 2016. “Synergistic Effect of MgO/Ag Co-Doping on TiO2 for Efficient
Antibacterial Agents.” Materials Letters 184, 82–87. doi: 10.1016/J.MATLET.2016.08.037.
Saleh, T.A., and V.K. Gupta. 2014. “Processing Methods, Characteristics and Adsorption Behavior of Tire Derived Carbons: A Review, Adv.”
Colloid Interface Sciences 211, 93–101. doi: 10.1016/J.CIS.2014.06.006.
Sanchez-Dominguez, M., G. Morales-Mendoza, M.J. Rodriguez-Vargas, C.C. Ibarra-Malo, A.A. Rodriguez-Rodriguez, A. V. Vela-Gonzalez,
S.A. Perez-Garcia, and R. Gomez. 2015. “Synthesis of Zn-Doped TiO2 Nanoparticles by the Novel Oil-In-Water (O/W) Microemul-
sion Method and Their Use for the Photocatalytic Degradation of Phenol.” Journal Environment Chemical Engineering 3, 3037–3047. doi:
10.1016/J.JECE.2015.03.010.
Saravanan, R., F. Gracia, M.M. Khan, V. Poornima, V.K. Gupta, V. Narayanan, and A. Stephen. 2015. “ZnO/CdO Nanocomposites for Textile
Effluent Degradation and Electrochemical Detection.” Journal Molecular Liq 209, 374–380. doi: 10.1016/J.MOLLIQ.2015.05.040.
Srikanth, B., R. Goutham, R. Badri Narayan, A. Ramprasath, K.P. Gopinath, and A.R. Sankaranarayanan. 2017. “Recent Advancements in
Supporting Materials for Immobilised Photocatalytic Applications in Waste Water Treatment.” Journal Environment Manage 200. doi:
10.1016/j.jenvman.2017.05.063.
Tan, T.T.Y., S. Liu, Y. Zhang, M.-Y. Han, and S.T. Selvan. 2011. Microemulsion Preparative Methods (Overview).
Van Driel, B.A., P.J. Kooyman, K.J. Van Den Berg, A. Schmidt-Ott, and J. Dik. 2016. “A Quick Assessment of the Photocatalytic Activity of TiO2
Pigments — From Lab to Conservation Studio!.” Microchem Journal 126, 162–171. doi: 10.1016/J.MICROC.2015.11.048.
Yang, X., W. Chen, J. Huang, Y. Zhou, Y. Zhu, and C. Li. 2015. “Rapid Degradation of Methylene Blue in a Novel Heterogeneous Fe3O4
@Rgo@Tio2-Catalyzed photo-Fenton System.” Sciences Reports 5, 10632. doi: 10.1038/srep10632.
Yu, J.C., L. Wu, J. Lin, P. Li, and Q. Li. 2003. “Microemulsion-Mediated Solvothermal Synthesis of Nanosized CdS-sensitized TiO2 Crystalline
Photocatalyst.” Chemical Communicable 0, 1552. doi: 10.1039/b302418k.
Zhang, L., C. Dai, X. Zhang, Y. Liu, and J. Yan. 2016. “Synthesis and Highly Efficient Photocatalytic Activity of Mixed Oxides Derived from
ZnNiAl Layered Double Hydroxides.” Transactions Nonferrous Met Social China 26, 2380–2389. doi: 10.1016/S1003-6326(16)64360-1.
Zhang, M., and B. Gao. 2013. “Removal of Arsenic, Methylene Blue, and Phosphate by biochar/AlOOH Nanocomposite.” Chemical Engineering
Journal 226, 286–292. doi: 10.1016/J.CEJ.2013.04.077.

14

You might also like