Sunteți pe pagina 1din 48

Author’s Accepted Manuscript

Influence of Sm3+ ions on structural, optical and


solar light driven photocatalytic activity of spinel
MnFe2O4 nanoparticles

Shankar K. Rashmi, Halehatty Seethya Bhojya


Naik, Honnalli Jayadevappa, Chittanahalli N.
Sudhamani, Sunitha B. Patil, Manjyanaik www.elsevier.com/locate/yjssc
Madhukara Naik

PII: S0022-4596(17)30320-1
DOI: http://dx.doi.org/10.1016/j.jssc.2017.08.013
Reference: YJSSC19903
To appear in: Journal of Solid State Chemistry
Received date: 7 June 2017
Revised date: 28 July 2017
Accepted date: 9 August 2017
Cite this article as: Shankar K. Rashmi, Halehatty Seethya Bhojya Naik, Honnalli
Jayadevappa, Chittanahalli N. Sudhamani, Sunitha B. Patil and Manjyanaik
Madhukara Naik, Influence of Sm3+ ions on structural, optical and solar light
driven photocatalytic activity of spinel MnFe2O4 nanoparticles, Journal of Solid
State Chemistry, http://dx.doi.org/10.1016/j.jssc.2017.08.013
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
1

Influence of Sm3+ ions on structural, optical and solar light driven


photocatalytic activity of spinel MnFe2O4 nanoparticles
a, b
Shankar K. Rashmi , Halehatty Seethya Bhojya Naikb, *, Honnalli Jayadevappa a
,
b b b
Chittanahalli N. Sudhamani , Sunitha B. Patil , Manjyanaik Madhukara Naik
a
Department of Chemistry, Sahyadri Science College, Shimoga-577203, Karnataka, India
b
Department of Studies and Research in Industrial Chemistry, School of Chemical Sciences, Kuvempu
University, Shankaraghatta-577451, Karnataka, India
E-mail: hsb_naik@rediffmail.com; Fax: +91-8282-256255

Abstract

In this article, co-precipitation method was used to produce a series of samarium substituted

manganese ferrite nanoparticles, MnFe2-xSmxO4 (x = 0.0, 0.5, 1.0, 1.5, 2.0) with oleic acid as

surfactant. The effect of Sm3+ substitution and calcination temperature on the MnFe2O4 was

explored. The MnFe2-xSmxO4 sample calcinated at 973 K induced an enhanced optical

behaviour. The influence of Sm3+ ions on cationic distribution and optical properties of

MnFe2O4 were probed by various physicochemical techniques. The presence of secondary

orthorhombic (SmFeO3) phase with initial cubic phase was confirmed by doublet peak in the

X-ray diffraction pattern. The particle size distribution of MnFe0.5Sm1.5O4 sample was about

40 nm from HRTEM. The visible light absorption ability extended as Sm content was

increased up to x = 1.5 in MnFe2O4. At an optimum concentration of Sm as x =1.5 in

MnFe2O4, photocatalyst exhibited decrease in energy band gap (1.64 eV); as a result

effective visible light driven photocatalytic activity was achieved than the pure MnFe2O4.

The MnSm1.5Fe0.5O4 sample exhibits pronounced photodegradation towards Rose Bengal

with in 50 min under direct sunlight illumination as energy input. The broad emission band

is observed in the entire photoluminescence (PL) spectrum. Current study provides an

excellent competency of rare earth substituted ferrite as a new class of photocatalyst and they

found to be a beneficial in the field of environmental cleaning application.

Keywords: MnFe2-xSmxO4; Rose Bengal; Solar light illumination; Photocatalytic activity


2

1. Introduction

Current trends in visible light induced photocatalysis have ensued in the fabrication of

various semiconducting nano-photocatalyst towards the remediation and decontamination of

wastewater. Different environmental pollutants, such as pesticides, detergents, dyes, and

volatile organic compounds in waste water owed to rapid industrialization and growing

population. Despite of all these pollutants, dyes from the textile industries pollute the

environment and a serious problem caused due to its intense colour and carcinogenicity.

About 1–20% of total world production of dyes from textile industries is lost during the

process dyeing and printing [1]. Photocatalytic degradation is eco-friendly method for

terminating the organic pollutants; it involves photocatalyst and sunlight as the photon source

in the visible region. As a response, set of oxidation-reduction (redox) reactions between

active species formed from a photocatalyst by illumination and organic pollutants. The term

‘photocatalysis’ or ‘photocatalytic degradation’ signifies complete photocatalytic oxidation or

photomineralisation, essentially to CO2, H2O, NO3−, PO43− and halide ions [2].

Development of semiconductive material such as iron oxide nanoparticles as a

magnetic recoverable catalyst which is suitable for cleaning industrial water environment for

commercial application, has been focused on by many researchers [3-6]. Easterday R et al.

[7] reported the iron oxide nanoparticles as efficient catalysis for the hydrogenation of

nitrobenzene via the incorporation of ruthenium into its crystal structure. The same group has

discussed the incorporation of Pd in to the iron oxide for the dehydrogenation of alkyne

alcohols [8]. It is well known that iron oxide nanoparticles are chemically active and are

easily oxidised in air, resulting in loss of magnetism and dispersibility. Therefore, it requires

the proper coating to keep the stability of magnetic iron oxide nanoparticles. Consequently,

developing the photocatalyst which are chemically stable and capable of using sustainable

solar energy effectively is a current challenge.


3

The magnetic spinel nanoferrites (MFe2O4) are efficient visible light responsive nano-

photocatalyst with a narrow band gap of ∼ 2.0  0.5eV [9, 10]. Manganese ferrites, is one of

the utmost important family of ferrite materials with partially inverse spinel structure [11],

have suitable band gap (1.74 eV) for the photodegradation of dye [12, 13]. They exhibit an

excellent optical absorption over low energy photon which makes it capable of absorbing

visible light radiation and huge advantage of using ferrite as photocatalyst is due to their

magnetic properties, making these materials recoverable from the catalytic systems, in order

to reuse them in other degradation processes. Substitution effects have been evidenced to be

very active in tunning the optical property of the ferrite composite [14-16]. Many strategies

have been made to improve the capability of using solar radiation to a great extent by

substituting divalent and trivalent transition metal in manganese ferrite [17, 18].

Nowadays, transition metal ferrites are extended to the inner transition metal

substituted ferrites. Rare earth ions that have partially filled 4f shells, if incorporated into

suitable ferrite matrixes, their intra-4f optical transitions become possible due to the splitting

induced by the crystal field of the matrix with enhanced photocatalytic activity. Many groups

has been reported the effect of rare earth ion substitution on structural, magnetic and

electrical properties of ferrite and some of them are on La, Nd, Gd, Eu, Y, and Pr [19-21].

Rare earth elements having its 4f orbital totally screened by 5s and 5p orbitals interact with

3d electrons of transition metal [22]. This plays an important role in deciding the electrical

and magnetic properties of the ferrites.

However, the fabrication and photocatalytic properties of rare earth ion substituted

ferrite nanoparticles are still lacking in study. In the literature, rare earth ions substituted

spinel ferrite structure, subjected to this study, are offered as materials with a wide range of

application in various photocatalytic processes. As part of the dawn, Nd substituted Ni ferrite

was reported by our group [23], where the partial replacement of Fe3+ ions by Nd3+ ions
4

resulted in the formation of metastable energy states originated by Nd 4f electrons with in the

energy gap, which resulted in the reduce of optical band gap of nickel ferrite with enhanced

photocatalytic activity. Besides this doping of rare earth ion (Sm, Dy and Nd) in magnesium

ferrite was reported by Thankachan Smitha et al [24] with excellent catalytic activity in the

degradation of 4-chlorophenal, these results were possible due to the doping of rare earth

ions, which facilitates the large availability of charge carriers and mobility in case of doped

nanoparticles. These literature reports indicate that rare earth ion as a dopant proves to be a

new and more superior class of dopant in case of ferrite nanoparticles with supreme solar

light driven photocatalyst for environmental issues.

Numerous techniques have been used for the fabrication of ferrite, among these

technique homogeneous precipitation is a versatile technique used to produce nanosized

metal ferrites because of its simplicity, low cost and bulk production compared to other

methods [21, 25]. The crystallisation of spinel structure involves elevated temperature with

longer duration. Specifically, inner transition metal substituted ferrites have higher thermal

stability than the pure one and more energy is needed to complete grain crystallization and

growth [26, 27]. Albeit of several studies on manganese ferrite and its supremacy over

photocatalytic properties, to this date, there is no such investigation was reported on the

influence of Sm3+ ion on the optical and photocatalytic properties of the manganese ferrite.

In the present work, we have synthesised pure and samarium substituted manganese

ferrite nanoparticles via co-precipitation route and has utilized inexpensive available sunlight

as a photon source for photocatalytic activity, on Rose Bengal (RB) as probe reaction in

aquatic environment. This study provides the further substantiation on effect of different

calcination temperature on structural properties of Mn ferrites. The information gained from

this work will be useful for further design of solar light driven rare earth substituted ferrite in

the treatment of dye-containing effluents.


5

2. Experimental

2.1. Chemicals

Manganese chloride hexahydrate (MnCl3.6H2O) (99.9%, Aldrich), iron nitrate

nonahydrate (Fe(NO3)3.9H2O) (99.9%, Aldrich), samarium chloride hexahydrate

(SmCl3.6H2O) (99.9%, Aldrich), sodium hydroxide (NaOH) (Merck) and Rose Bengal (RB)

(99.9%, Aldrich) were of analytical grade and used directly as received without additional

purification. Ultrapure de-ionized water was used as the reaction medium in all the synthesis

steps for dilution and sample preparation.

2.2. Sample preparation

Pure manganese ferrite nanoparticles were synthesised by chemical co-precipitation method.

The detailed synthesis procedure was as follow: manganese chloride and ferric nitrate are

weighed in the ratio (1:2) and dissolved in the ultrapure de-ionized water by sonicating it for

30 min then mixed together and stirred to form homogeneous mixture. Sodium hydroxide

(precipitating agent) of 3 M was added drop wise to the preheated homogeneous mixture

under continuous stirring till the permanent precipitate is formed. Adjusting the pH of the

medium less than 10 (pH<10), results in the formation of brown precipitate, then 4-5 drops of

oleic acid was added as surfactant. The chemical reaction was brought to a temperature of

about 353 K for an hour. Then precipitate was washed repeatedly with de-ionised water and

ethanol to extract the surfactant followed by acetone to speed up the drying. The obtained

brown solid product was filtered and dried at 333 K in the oven for 12 hour and calcinated at

673 K, 873 K, 973 K and 1373 K in alumina crucible for 8 hour in vacuum and furnace

cooled to the room temperature. The resulting powder was again grounded and recalcinated at

the same temperature for another 2 hour.

The similar experimental conditions were used for the preparation of MnFe2-xSmxO4 where (x

= 0.5, 1.0, 1.5 and 2.0) and calcinated at 973 K.


6

2.3. Characterization

The photoabsorption measurement of the ferrite samples were analysed by UV-Vis

spectrophotometer (Shimadzu, UV-1650 PL model) dispersed in ethylene glycol by

ultrasonication. The identification of phase composition and crystalline structure of ferrite

powders, calcinated at different temperature were done using X-ray diffraction (PANalytical

Xpert Pro X-ray diffractometer) with Cu-Kα (λ = 1.5406 Å) radiation. The samples fixed in

the reflection mode were analysed in the ambient atmosphere with a scanning rate of 0.01/s

over 2θ range of 10 to 80. The crystallite size was calculated by standard Scherer formula.

The surface morphology and chemical composition of the ferrite samples annealed at 973 K

were examined by field emission scanning electron microscopy (FESEM-Carl ZEISS, Supra

40VP) and energy dispersive X-ray spectroscopy (EDS). The high resolution transmission

electron microscopy, FEI, Technai G2, F30 (accelerating potential 300kV, resolution point:

2.0 Å line: 1.0 Å) was employed to probe the particle size distribution and high resolution

imaging of MnFe0.5Sm1.5O4 sample calcinated at 973 K. The photoluminescence (PL)

spectrum of ferrite samples were recorded using a Jasco FP-8500 fluorescence spectrometer

with a 450 W Xenon lamp as a light source. The FTIR spectra were recorded using a Nicolet

IR200 FT-IR spectrometer by KBr pellet technique.

2.4. Photocatalytic activity

The photocatalytic performance of MnFe2-xSmxO4 (x = 0, 0.5, 1.0, 1.5 and 2.0)

photocatalysts were evaluated by the degradation of RB under sun light illumination. The

experiment was executed with 100 ml of RB solution using deionised water. In each reaction

suspension, 0.08 g of photocatalyst was added in 100 ml solution of RB with an initial

concentration of 10 mg L-1. The suspension was kept in dark for 30 min with vigorous

stirring to attain adsorption-desorption equilibrium between dye and photocatalyst at room

temperature, then the solution was illuminated under direct sun light on sunny day of March
7

between 11.30 a.m. to 2 p.m. During illumination, stirring was maintained in order to keep

the mixture in suspension. At regular interval of time, 5 ml of sample was withdrawn and

photocatalyst was magnetically separated to be reuse for additional runs. The change in the

concentration (absorbance) of solution was monitored on UV-Vis spectrophotometer by

measuring the absorbance in 200–800 nm range using deionised water as reference. The

absorbance of RB solutions was determined at 548 nm and it is used to monitor the

degradation of dye. The temperature of the solution was between 301 K and 305 K in acidic

and basic medium. These measurements were also done for all the samples in dark and under

sunlight. Recyclability of MnFe2-xSmxO4 solid product remained after each catalytic run was

dried and again used as photocatalyst. The MnSm1.5Fe0.5O4 sample was tested as a

comparison catalyst, with varying the pH of the suspension following the same procedure to

evaluate the importance of H+/OH- on the photocatalytic activity. The pH of the suspension

was adjusted by adding HCl or NaOH.

3. Results and discussion

The powder X-ray diffraction pattern of pure and Sm substituted MnFe2O4

nanoparticles are manifested in Fig. 1(a-b). In order to investigate the effect of calcination

temperature on the structure and phase formation of MnFe2O4, XRD analysis was conducted

(Fig. 1a). The sharp and intense peak shows the fine crystals and high purity samples. At low

calcination temperature of 673 K, powders show pure spinel structure (JCPDS 10-0319);

possibly due to slow diffusion rate accompanying with muffled crystallization. However, at

873 K, concurrent with ferrite structure the sample discloses the appearance of secondary

Fe2O3 and Mn2O3 (JCPDS 33-0664 and 24-0508) phase due to the oxidation of MnFe2O4. The

Fe2O3 and Mn2O3 with body-centered cubic (BCC) structure cannot be dissolved in the face

centred cubic spinel structure of MnFe2O4 [28], so precipitate beside as secondary BCC
8

phases [29]. Another reason for the presence of Fe2O3 and Mn2O3 phase is due to the

decomposition pressure of MnFe2O4 is lower than the oxygen partial pressure in air,

MnFe2O4 calcinated at this particular temperature is unstable. At elevated calcination

temperature, the metal oxides get reincorporate in to the crystal lattice of MnFe2O4. The fine

and well crystallised cubic spinel phase of MnFe2O4 (JCPDS 73-1964) is formed at 1373 K,

which is thermodynamically stable. This temperature is high enough to form a pure ferrite

phase. A similar behaviour of MnFe2O4 was described by Dong et al. [30] at high calcination

temperature (1173 K), the particles start to recrystallize and grow in size. The crystallite size

of MnFe2O4 nanoparticles has been computed using Debye-Scherrer’s formulae from the

broadening of the (311) peak. [31].

D=0.9λ/βcosθ (1)

Where D is the crystallite size (nm), β is the full width half maximum intensity

measured in radians, λ is the X-ray wavelength and θ is the diffraction angle. The

experimental lattice parameter ‘aexp’ was calculated from the interplanar spacing using the

following equation for cubic structure [32].

a = dhkl (h2 +k2 +l2)1/2 (2)

The d-spacing values were calculated for the recorded peaks using Bragg’s law. Table

1 imputes the value of crystallite size and lattice parameter of MnFe2O4 for various

calcination temperatures. The crystallite size upturns from 39 to 62 nm as calcination

temperature rises. Increase in calcination temperature induces a rise in lattice parameter of

MnFe2O4 from 8.349 to 8.545 Å.

Fig 1b shows the diffraction pattern of Sm substituted MnFe2O4 with various

composition calcinated at 973 K. The reason for the calcination of MnFe2-xSmxO4 at 973 K

will be explained later in the optical section. In all Sm-substituted ferrite, a gradual shift in

the prominent (311) peak towards the smaller 2θ angle confirms the occupancy of Sm3+ ions
9

in the MnFe2O4 spinel structure [33]. Particularly for x=1.5 sample, the secondary

orthorhombic SmFe2O3 phase is formed with initial cubic spinel phase due to high reactivity

of Fe3+ with Sm3+ ions on the grain boundaries [34]. It can be explained by the solubility limit

of ions in the spinel lattice owing to substantial difference in ionic radii of Sm3+ (0.964 Å)

and Fe3+(0.648 Å) ions [35, 36]. This causes an increase in the lattice parameter, which

results in restrained distortion of the lattice. For sample x=2.0 , the Sm3+ ions are

accommodated by complete replacement of Fe3+ ions in host spinel lattice contributes in the

formation of cubic crystalline structure of MnSm2O4 accompanying with secondary Sm2O3

phase. Fig. 1c reveals the diffraction peak corresponding to MnFe0.5Sm1.5O4 nanoparticles at

different calcination temperatures (673, 873, and 973 K). The degree of crystallinity

enhanced with temperature and reduces the additional broadening. This suggest that the more

energy is needed to make Sm3+ ions enter into lattice site and form to the Sm3+ -O2- bond due

to larger ionic radii as compared to Fe3+ ions [37]. As a result, the energy required for the Sm

substituted samples to far-reaching crystallization and to grow grains is more [38] and

average crystallite size of MnFe0.5Sm1.5O4 at different calcination temperatures (673, 873,

and 973 K) are 28, 35 and 43 nm, increase in calcination temperature involved in the growth

of the crystallite size [26]. The crystallite size dives from 59 to 38 nm with cumulative

Sm3+content in MnFe2O4 (Table 2). These results specifies that the Sm3+ ions restrict the

grain growth of the ferrite phase owed to the (SmFe2O3) secondary phase formation together

with dominant MnFe2O4 phase, whereas calcination temperatures participate in their grain

growth [39]. The experimental lattice parameter aexp upturns from 8.423 to 9.108 Å with

Sm3+ insertion, since the ionic radius of Sm3+ is larger as compared to Fe3+ ions. The Sm3+

ions have strong site preference towards the octahedral B- site, which prompts the expansion

of unit cell. It endorses the occupancy of Sm3+ ion in MnFe2O4, which results in moderate
10

distortion of the lattice. The analogous result has been reported by many groups on rare earth

ion substituted Mg-ferrite [40], Mg-Mn ferrite [41] and Co-ferrite [42].

MnFe2O4 possess partially inverse spinel structure, where Mn2+ ions predominantly in

the tetrahedral site (low degree of inversion) and Fe3+ ions are distributed between A- and B-

interstitial sites [11]. Based on the Neel’s two sub lattice model [32], the cation distribution

of Sm-substituted MnFe2O4 can be assumed as follows.

[Mn2+0.75Fe3+0.25][Mn2+0.25Sm3+xFe3+1.75-x]O42-

The ionic radii of tetrahedral A- sites (rA) and octahedral B- sites (rB) in the spinel structure

can be determined from the relation.

rA = [CMn r( Mn2+) + CFe r (Fe3+)] (3)

rB = 0.5[CMn r (Mn2+)+ CSm r (Sm3+)+ CFe r (Fe3+)] (4)

Where, rMn2+, rFe3+, rSm3+ are cationic radius of Mn, Fe and Sm respectively. The

proposed cation distribution and values of rA and rB are listed in Table 2. It can be concluded

from the cation distribution that the Sm3+ ions has strong preference to the octahedral site,

since the ionic radius of Sm3+ ions is larger than the other cations. The interspace of

tetrahedral site is smaller than those of octahedral site [43], it help to accommodate the Sm3+

ions of larger ionic radius. Mn2+ ions also have a small fraction of the octahedral site

occupancy with the strong preference to the tetrahedral sites. The larger Sm3+ ions (0.964 Å)

successively replace the smaller Fe3+ ions (0.648 Å) in octahedral site due to the site

preference. This results in the rearrangement of the cations between tetrahedral and

octahedral sites to minimize the free energy of the system by migrating fraction of Mn2+ ions

from B- to A- sites [44] to accommodate larger Sm3+ ions. Hence, the ionic radius of

octahedral sites increases, where slight increase in rB is correlated with the migration of Mn2+
11

ions from B- to A- site. The theoretical lattice parameter (ath) and oxygen positional

parameter were calculated by using the following equation.

(5)

(6)

Where R0 is the radius of the oxygen ion i.e. R0 = 1.32 Å. The data on aexp, ath, rA, rB,

D, and u for MnFe2-xSmxO4 are given in Table 2. The linear increase in ath can be explained

on the basis of ionic radii difference of the component ions. It is observed that experimentally

calculated aexp and theoretically calculated ath shows similar trend. In an ideal face centered

cubic spinel structure u value is close to 0.375 Å. In the present investigation, u parameter

varies from 0.3767-0.3865 Å. Increase in the value of u parameter indicates the distortion of

lattice due to the substitution of larger Sm3+ ions in octahedral sites thereby relocation of

Mn2+ ions from octahedral site to tetrahedral site and deviation from ideal case [45]. Zhang L.

Y. et al. [44] also observed the similar increasing trend in u value for Eu3+ doped MnFe2O4.

The microstructure of the MnFe2-xSmxO4 nanoparticles calcinated at 973 K was

studied by FE-SEM micrographs. The images with different resolution (1 m, 25 Kx; 200

nm, 75 Kx) of MnFe2O4 and (1 m, 25 Kx; 200 nm, 75 Kx) MnFe0.5Sm1.5O4 are depicted in

Fig. 2. The individual particles of pure MnFe2O4 appear to be platelets like fine granular

microstructure (Fig. 2a and b). The high content of Sm3+ ions might prevent the grain growth

due to the formation of SmFeO3 secondary phase, resulting in smaller particles of

MnFe0.5Sm1.5O4 sample (Fig. 2c and d). Due to the larger bond energy of Sm3+ -O2- compared

to that of Fe3+ -O2-, more energy is needed to enter the lattice site and to form the Sm3+-O2-

bond [21].
12

The compositional analysis for MnFe2-xSmxO4 nanoparticles are carried out to

confirm the Sm substitution in elemental level by EDS spectrum (Fig.3a-d). It indicates the

presence of Mn, Fe, Sm, and O atoms as major chemical components in the sample. The

atomic ratio of Mn: Sm: Fe for x=1.5 is about 11.35:14.40:4.51, which is close to

experimental stoichiometric ratio with small error, confirms the formation of MnFe0.5Sm1.5O4

nanoparticle. The elemental analysis as obtained from the EDS is in close agreement with the

predictable composition from the stoichiometry of reactant solutions used.

Fig. 4a-d shows TEM, HRTEM images with particle size histogram and SAED of

MnFe0.5Sm1.5O4 nanoparticles. The surface appearance of MnFe0.5Sm1.5O4 nanoparticles are

shown in Fig. 4a and b. The magnified TEM image points that the most of the particles seems

to be spherical in shape and are agglomerated to some extent. The strong magnetic attraction

and presence of water between ferrite nanoparticles may result in such type of agglomeration

[41]. The mean crystallite size of MnFe0.5Sm1.5O4 nanoparticles from histograms is 40 nm

(Fig. 4d). The crystallite size determined from XRD technique and TEM are in good

agreement with each other. HRTEM images reveals the clear lattice fringes with a

interplanar spacing of approximately 0.263 nm corresponding to 311 crystal planes of spinel

phase.

Fig. 5 shows the FTIR spectra of MnFe2-xSmxO4 nanoparticles of two different Sm

concentrations (0 and 1.5). The bands in IR spectra consigned from inter atomic vibrations.

Ferrite shows two principle vibration bands Fe-O and M-O are analogous to the tetrahedral

(ν1 ~ 600-500 cm-1) and octahedral (ν2 ~ 450-380 cm-1) interstitial lattice site vibrations,

which are considered to be the typical bands of spinel structure. Pure MnFe2O4 shows

vibrational frequencies around 590 cm-1 (ν1) and 410, 505 cm-1 (ν2) are due to the tetrahedral

M-O vibration on A site and octahedral M-O vibration on B site [46]. MnFe2O4 possess
13

partial inverse spinel structure, where about 20% of the Mn2+ ions are on the B sites

(octahedral sites) and the rest resides on the A sites (tetrahedral sites) [11]. Generally, the Fe-

O bond length at B site is more than A site. The substitution of Sm3+ ions shows the shoulder

as well as shifting of ν2 band to higher frequency in the range (420-442 cm-1) can be seen.

These bands appear due to the formation of (Sm3+-O2-) bond in addition to (Fe3+-O2-) bond in

octahedral site [47]. Another reason for splitting of ν2 is the formation of SmFeO3 secondary

phase at B site. The slight shift in the position and intensities of ν1and ν2 bands are due to the

perturbation arising in the bond length of Fe3+-O2- by addition of Sm3+ ions. The peak at 3400

cm-1 and 1600 cm-1 are assigned to the stretching and bending vibration of adsorbed H-O-H

molecule on the surface of nanoparticle. The absorption band at 2350 cm-1 and 1000 cm-1 are

assigned to KBr and –OH groups [48]. The characteristic band of pure Sm2O3 crystals at 800

cm-1, ascribed to the stretching vibration of Sm3+–O2- group [26]. The increase in unit cell

dimensions due to replacement of Fe3+ ions by Sm3+ ions affects the Fe3+-O2- stretching

vibration in octahedral site, it results in the band shift [49].

The effect of samarium concentration on absorbance of MnFe2O4 was probed by UV-

Vis absorption spectrophotometer. The MnFe2-xSmxO4 calcinated at 973 K shows pronounced

optical behaviour. The modification in MnFe2O4 with samarium concentration significantly

affected the absorption efficiency. Visually, the incorporation of Sm3+ ions in MnFe2O4

spinel matrix was substantiated by the color change of the Sm-MnFe2O4 nanoparticles. Pure

MnFe2O4 was brown, whereas the MnFe0.5Sm1.5O4 shows typically intense black brown,

signifying an apparent change in electronic structure of substituted sample [50]. The

absorption peak of MnFe2O4, MnFe1.5Sm0.5O4, MnFe1Sm1O4, MnFe0.5Sm1.5O4 and MnSm2O4

occurred at wavelength of 463, 482, 500, 525 and 412 nm respectively. Particularly, as

shown in the absorption spectra (Fig. 6a), the absorbance peak shifted towards longer
14

wavelength (red- shift) with the increase in samarium concentration was about 80 nm. The

apparent red shift indicated that the MnSmxFe2-xO4 nanoparticle size condensed with increase

in samarium concentration, it was consistent with the result calculated by the Scherer’s

formula. The absorption of MnFe2O4 in visible region was likely due to the transition of

photoexcited electrons from O-2p level to Fe-3d level [51]. Indeed, MnFe0.5Sm1.5O4

photocatalyst exhibit greater visible light absorption than pure MnFe2O4, it states the lower

band energy of MnFe0.5Sm1.5O4 and expected to be a better photocatalytic capability under

visible region. The shift in absorption edge was due to the substitution of Sm3+ ion in

MnFe2O4, which directs the change in the octahedral lattice with expansion of unit cell. The

intense black brown coloured MnFe0.5Sm1.5O4 photocatalyst own abundant surface and

interface defects, which induces an additional sub energy level in the system as a dopant

energy level [10]. In contrast to this, the absorption edge related to MnSm2O4 shows blue

shift due to the disappearance of Fe+3 ions in MnSmxFe2-xO4 [23] with decrease in

nanoparticle size [17].The band gap energy of the ferrites was resolute from their absorption

spectra. The optical band gap Eg assessed by a extrapolating the linear portion curve obtained

by plotting (αhν) 1/n vs. energy of absorbed light (hν) to α = 0, for n= 1/2 assigned for direct

allowed transition, where α is the absorption coefficient. The best linear fit observed for

n=1/2 assigned for direct allowed transition (Tauc plot) (Fig. 6b) [52]. The estimated value of

band gap for pure and sm substituted samples are 1.74, 1.72, 1.68, 1.64 and 2.0 eV

respectively. The band gap energy of MnFe2O4 is about 1.74 eV in accordance with the

previous reports [12].The band gap energy of Sm substituted samples is comparatively lower

than the MnFe2O4 (1.74 eV). As samarium content increases the energy band gap value

decreases, especially in case of x = 1.5 sample shows a red shift of about 1.64 eV contrast to

other samples (x = 0.5, 1.0, and 2.0). The red shift in absorbance is predominantly caused by

partial replacement of Fe3+ ions by Sm3+ ions owing to the introduction of Sm 4f electrons in
15

MnFe2O4 which forms a donor energy level below and closer to the conduction band, which

can be explained on the basis of surface phenomena [53, 23]. For the sample with x = 2.0, the

energy band gap increased to 2.0 eV, this increase may be due to the smaller bond length of

Sm3+-O2- when compared to the spinel ferrite, it can be explained on the basis of atomic pair

distribution function [54]. From the above result, it can be concluded that, MnSm2O4 samples

has no additional sub energy level or donar level between valence band (O-2p) and

conduction band (Sm-4f) due to the complete replacement of Fe3+ ions by Sm3+ ions. Thus,

MnFe0.5Sm1.5O4 nanoparticle have greater band gap values (1.64 eV) than the theoretical

energy which is required for the water splitting (1.23 eV) which makes them an efficient

photocatalyst with respect to visible light utilisation.

The calcination temperature is sensitive knob for tuning the optical properties of the

MnFe2-xSmxO4 photocatalyst. MnFe2-xSmxO4 photocatalyst calcinated at 873 K also shows

enhanced visible light absorption with red shift as samarium content increased (Fig. S1). The

absorption peak of MnFe2O4, MnFe1.5Sm0.5O4, MnFe1Sm1O4, MnFe0.5Sm1.5O4, MnSm2O4

occur at wavelength of 430, 435, 440, 460 and 395 nm and corresponding band gap energy of

pure and Sm substituted samples are 1.78, 1.77, 1.76, 1.73 and 2.2 eV respectively. The shift

in absorption edge toward the longer wavelength as a function of calcination temperature

results in the contraction of band energy by 0.09 eV (MnFe0.5Sm1.5O4, at 973 K). The MnFe2-

xSmxO4 calcinated at 973 K shows pronounced optical behaviour than sample at 873 K due to

the enhanced crystallite size of ferrite as calcination temperature is upturned. However, above

1073 K, MnFe0.5Sm1.5O4 is inactive in whole visible region due to further enlargement in the

crystallite size of the nanoparticle and loses its nanostructure (Fig. S2).

The photoluminescence spectra of MnFe2O4 and MnFe0.5Sm1.5O4 were recorded at room

temperature to detect the structural defects which affects the optical properties of ferrite and
16

are shown in Fig. 7. The samples were excited by single excitation wavelength of 370 nm

with the excitation source of 150W Xenon lamp. The emission band was lopsided and

broadened with multiple peaks indicating the participation of different luminescence centers

in the radiative processes. The photocatalytic activity insistently relates to light harvesting

ability, separation efficiency of photogenerated e-/h+ and the recombination rate of photo-

excited electrons and holes. Both the samples shows similar emission band pattern from 400

to 550 nm. The quenching in the intensity of the emission band with Sm3+ concentration

intends to the increasing the number of oxygen vacancies and interstitial metal defects

created in tetrahedral and octahedral sites of the above ferrite compositions. The lower

emission intensity indicates the decrease in the recombination of the photo-induced electron–

hole pairs, thus the higher photocatalytic activity can be expected [55]. This suggests that the

MnFe0.5Sm1.5O4 photocatalyst has lesser opportunity for electron–hole pairs recombination

and facilitates the migration of charge carriers more effectively than the pure MnFe2O4.

Makishima et al. [56] earlier investigated the photoluminescence emission of Sm3+ in BaTiO3

host lattice and found that some foreign ions can change the relative strength of the emissions

owing to a charge compensation mechanism. On the basis of their results, they concluded that

one series of the emissions is attributed to Sm3+ at the Ti4+ site, while the other series of

emissions is related to the presence of Sm3+ in the Ba2+ site.

A series of experiment was performed to evaluate the effect of Sm Substitution in

MnFe2O4 on photodegradation process. The photocatalytic activity of MnFe2O4,

MnFe1.5Sm0.5O4, MnFe1Sm1O4, MnFe0.5Sm1.5O4 and MnSm2O4 photocatalyst were examined

by the photodegradation of RB in aqueous solution under day light illumination for 90, 70,

60, 50, and 100 min (Fig. 8). Generally, photo-oxidation takes place at the surface of

photocatalyst [57], for this reason, the adsorption characteristics of dye/photocatalyst system

was probed. In order to correlate the adsorption behaviour with larger surface area of the
17

MnFe2-xSmxO4 nanoparticles, additional experiments were conducted to determine the ability

of photocatalyst to adsorb the RB molecule in the absence of light for 60 min. There are no

such consequential adsorption effect was observed for all the samples. Under same

experimental conditions, (photolysis) without catalyst under solar light illumination, only

7.89% of reduced in RB concentration was observed. This clearly specifies that this reaction

follows photocatalytic degradation. The absorption edge of RB solution occurs at a

wavelength of 518 and 548 nm. Fig.8. shows that the quenching of peak intensity of the dye

with time directs the photodegradation of RB over MnFe2-xSmxO4 photocatalysts under solar

light irridation. The degradation of RB over MnFe2O4 photocatalyst reaches 89 % in 90 min.

Fig. 8(a-e) shows that the Sm-substitution in manganese spinel ferrite lattice results in

improved degradation rate with condensed time from 90 min (x=0.0) to 50 min (x=1.5).

Comparably, the higher photodegradation behaviour of MnFe0.5Sm1.5O4 (x=1.5) photocatalyst

of all sample under solar light illumination was renowned, which completely degraded the

RB (97.8 %) in 50 min. this may be due to the smaller particle size and narrow band gap

energy (~1.64 eV) owing to the partial replacement of Sm3+ in octahedral lattice site of

MnFe2O4. Meanwhile, the substitution of Sm3+ at x=1.5 is optimal dosage in MnFe2O4 which

makes them as effeceint photocatalyst in decolorasiation of various types of dyes. Ensuing in

significant utilization of visible light and enhanced separation of photoinduced charge

carriers, reflects as potentially higher photocatalytic activity [58]. Above the optimal limit

(Sm3+), x=2.0 in MnFe2-xSmxO4, the degradation rate diminished with time. The complete

replacement of Fe3+ by Sm3+ ions in octahedral site of MnFe2-xSmxO4, leads to the

recombination of photogenerated charge carrier, this implies that the MnSm2O4 photocatalyst

has wide band gap than the prior and shows lower photocatalytic activity. The percentage

photocatalytic degradation was calculated by following equation [59]

Photodegradation % = (C0 - Ct)/Ct 100 (7)


18

Where C0 is the initial RB concentration and Ct is the concentration of RB at the time t. the

percentage degradation of the dye using all samples are shown in fig. 8f.

Photocatalytic degradation mechanism of RB dye is as shown below. A series of

mechanism has been proposed by many research groups [60, 61] for visible light driven

photodegradation, on the basis of these reports and our result we have demonstrated a simple

mechanism (Scheme 1). In comparison with MnFe2-xSmxO4, pure MnFe2O4 also excited by

solar light but it has higher recombination of photoinduced electron-hole pair, which leads to

the low visible light photocatalytic activity for the degradation of organic pollutants. Another

reason could be attributed to the larger band gap of the MnFe2O4 (1.74 eV). The energy band

gap of photocatalyst determines the wavelength of light that can be the absorbed and leading

to the generation of electron-hole pairs. The substitution of the Sm3+ ion in MnFe2O4 results

in the enhanced visible light photocatalytic activity due to the formation of metastable Sm-4f

energy levels closer to the lower edge of the conduction band of MnFe2O4, which points the

reduction in the band gap. Under solar light illumination, electrons gets excited from valence

band (VB) to new energy level (MnFe2-xSmxO4) conduction band (CB) of the system and

simultaneously, the same amount of holes are generated in the valence band react with

surface water or hydroxyl ion to produce OH• radical species, which is a potential oxidant for

the degradation of RB and concurrently, electrons in the conduction band reacts with

adsorbed oxygen molecule to produce active oxygen species O2•. It further, combines with

H+ to produce HO2• [62], which react with trapped electron to generate OH• [63]. It can be

concluded that OH•, HO2•, O2• and h+VB are active species involved in the complete

oxidation process leads to the photodegradation of RB molecule. Based on the above

analysis, the photochemical reaction for the degradation of RB under solar light illumination

over MnFe2-xSmxO4 photocatalyst was expressed as follows.

Photoexcitation: MnFe2-xSmxO4 + hν MnFe2-xSmxO4 (e + h+)


19

Oxygen ionosorption: O2 + e O2•

Ionization of water: H2O H+ + OH

Protonation of superoxides: O2• +H+ HO2•

h++ OH OH•

2 e + HO2• + H+ OH• + OH-

HO2•, OH•, O2•, h++ RB degraded products

Fig. 9a illustrates the photodegradation of RB in the presence of MnFe2-xSmxO4

photocatalyst. The photocatalytic activity of RB (10 mg L-1) on various Sm concentration are

of the order as follows; MnSm2O4 < MnFe2O4 < MnFe1.5Sm0.5O4 < MnFe1Sm1O4 < Mn

Fe0.5Sm1.5O4. The adsorption of dye molecule on the surface of photocatalyst was constant

over time (dark reaction).

The kinetics behaviour of Sm-Mn ferrites in the degradation of RB was explored. To

comprehend the photodegradation kinetics of RB, Langmuir- Hinshelwood model was

applied [64], this model describes rate constant of photodegradation of the RB. It is

observable from Fig. 9b that the chemical kinetic removal of RB fits first order kinetic model.

ln (Ct/C0) = - kt (8)

Where, Ct is concentration of RB in the solution at illumination time t, C0 is the initial

concentration of RB in the solution at t = 0 and k is apparent first order rate constant. The

plot of ln (Ct/C0) as a function of illumination time, which is approximately linear and the

value of k can be obtained from its slope. From Fig. 9b it can be seen that the photocatalytic

activity of MnFe0.5Sm1.5O4 with k value of 0.0147 min-1 is higher than the pure MnFe2O4 (k=

0.0157 min -1) under solar light illumination. Increasing surface area of the MnFe0.5Sm1.5O4

due to the reduction in the particle size can provide greater active surface, leading to the

enhanced photocatalytic efficiency [57].The MnFe0.5Sm1.5O4 photocatalyst showing greater


20

degree of decolourisation on RB, was fixed for further trial. The obtained k values are

0.0728, 0.0980, 0.1823, 0.2038 and 0.0654 min-1 for MnFe2O4, MnFe1.5Sm0.5O4,

MnFe1Sm1O4, MnFe0.5Sm1.5O4 and MnSm2O4 photocatalyst respectively. The substitution of

trivalent cation Sm3+ with x=1.5 is the optimal dosage in MnFe2O4 for the most effeceint

separation of photoinduced e-/h+ pairs. Thus, the photocatalytic performance of MnFe2-

xSmxO4 photocatalysts increased with increase in Sm content initially, nevertheless, above its

optimal limit photocatalytic activity decreased. According to Liang Chun-Hua et al. [58],

when samarium concentration reaches above its optimal limit, surface barrier becomes higher

and the space charge region becomes narrower. Therefore, photo-generated electron-hole pair

recombination becomes easier, leading to the decrease in photocatalytic activity. Hereafter,

the performance of the MnFe0.5Sm1.5O4 in the photocatalytic processes is strongly influenced

by the deformation generated by the Sm content in the structure of MnFe2O4, observed from

XRD and UV-Vis analysis [65].

The degradation kinetics of RB on the various concentrations of MnFe0.5Sm1.5O4

nanophotocatalyst is shown in Fig. 10a. The photocatalytic activity of RB (10 mg L-1) on the

various MnFe0.5Sm1.5O4 concentration are in the order as follows; 0.04 g < 0.06 g < 0.08 g >

0.1 g > 0.12 g. The chemical kinetic removal of RB dye (10 mg L-1) at various concentration

of MnFe0.5Sm1.5O4 nanophotocatalyst were explored (Fig. 10b). The obtained k values are

0.0832, 0.1563, 0.2038, 0.0543 and 0.0242 min-1 for photocatalyst concentration of 0.04,

0.06, 0.08, 0.1 and 0.12 g/100 ml of RB, respectively. Obviously, the concentration of

MnFe0.5Sm1.5O4 nanophotocatalyst influences the rate of photodegrdation of RB. Initially, the

rate of photodegradation increased with increase in concentration of photocatalyst, due to the

large availability of the catalytic sites and thereafter becomes constant for further increase in

concentration. Excessive photocatalyst intensifies the thickness of the layer dispersed in the
21

dye solution. Thus, exposed surface area of the catalyst for light penetration is reduced.

Therefore, the best photocatalytic degradation was achieved for 0.08 g of catalyst

concentration with k value of 0.2038 min-1. This ensures that the rate constant is higher when

photocatalyst concentration increased up to 0.08 g/100 ml of dye solution.

The significant solar light driven photocatalytic behaviour of MnFe0.5Sm1.5O4

photocatalyst encouraged us to extend the photocatalytic activity to another few organic

pollutants. Fig. 11a shows the photodegradation rates of methyl orange (MO), methyl red

(MR), methylene blue (MB) and rhodamine B (RhB) under the identical conditions as for

RB. The comparison of their photodegradation yields are summarized in Table 3. The

photodegradation yield of these dyes exhibited analogous results with a minor variation due

to the structure and complexity of the dyes, particularly on the nature and position of

substituents in the aromatic rings; this specifies the efficiency of the MnFe0.5Sm1.5O4

photocatalyst. This result endorses that MnFe0.5Sm1.5O4 photocatalyst is expected to remove

high variety of environmental disrupters in water.

The dye and catalyst loading is important parameters that affect the photocatalytic

activity. These outcomes indicate that the excellent photocatalytic degradation was achieved

at 10 mg L-1 of dye concentration with 0.08 g of photocatalyst. At higher concentration of

dye, adsorption of dye molecule on the surface of catalyst was high and surface was also

saturated, which leads to decrease the photonic efficiency followed by catalytic deactivation

[66]. If catalyst loading is higher, degradation efficiency decreases, due to an unfavourable

light scattering and reduction in the light penetration [67].

The influence of pH on the photocatalytic degradation of RB was studied in the range

of 5.0-9.0 with 10 mg L-1 of initial concentration of dye and 0.08 g of catalyst. The pH of the

solution is greatly related to the surface-charge properties of the photocatalyst and affects the
22

ionic species in the solution. The rate of photodegradation increases with increase in the pH

of the solution up to 8.0. In acidic medium, the anionic dye molecules are in protonated form

and photocatalyst surface also possess positive charge due to the adsorption of H+ ions. So, as

the pH of the solution decreases, the dye molecule and photocatalyst repel each other and the

rate of photodegradation decreases. However, increase in the pH of solution reduces the

repulsion between dye molecule and the photocatalyst. Later, the rate of photodegradation

increases. Above pH 8.0, the rate of photodegradation starts decreasing, due to upturn in

repulsion between photocatalyst (becomes negatively charged) and the dye molecule.

To investigate the photostability and reusability of MnFe0.5Sm1.5O4 photocatalyst

under solar light illumination, the degradation experiment was conducted for five runs with

same sample (Fig.11b.). In each cycle, suspension was separated from solution by external

magnetic field. After five cycles, there was no loss of activity and the degradation efficiency

of RB was about 80%, which makes remarkable photostability. Ferrites as photocatalyst

possess a good magnetic property [68] and easily recoverable from catalytic system in order

to reuse them for long-term degradation process. This class of photocatalyst is expected to

remove high variety of environmental disrupters in water.

4. Conclusion

In this work, co-precipitation method was used to produce a series of samarium substituted

manganese ferrite ceramic nanoparticles, MnFe2-xSmxO4 (x = 0.0, 0.5, 1.0, 1.5, 2.0). The

influence of rare earth (Sm3+) ions on crystal structure led to a formation of cubic spinel

structure with a secondary samarium iron oxide (SmFeO3) phase. The insertion of Sm3+ ion in

the MnFe2O4 spinel matrix shows an increase in lattice parameter with decrease in crystallite

size attributed to larger ionic radii of Sm3+ ion as compared to Fe3+ ion in the B sites. The

platelets like fine granular morphology of pure and substituted MnFe2O4 were observed from
23

FESEM and HRTEM images. The particle size distribution of MnFe0.5Sm1.5O4 nanoparticles

were about 40 nm smaller than the MnFe2O4 as confirmed from HRTEM analysis.The

elemental analysis reveals the presence of the substituted element (Sm). The formation of

cubic spinel structure was confirmed by metal-oxygen bond perceived in FT-IR spectra. The

MnFe2-xSmxO4 calcinated at 973 K shows marked optical behaviour. The composition

MnFe0.5Sm1.5O4 showed a substantial red shift in absorption edge with narrow band gap of

1.64 eV, which proved to be an optimal samarium dosage to achieve highest photocatalytic

degradation of RB. Due to the fact that Sm3+ substitution enhance the solar light harvesting

and inhibit the recombination of electron-hole pairs in MnFe2O4, results in effective solar

light driven photocatalytic activity. The rate of degradation of RB over MnFe2-xSmxO4 rises

with increasing samarium content up to x=1.5 and then decreases. The complete degradation

of RB under direct sunlight illumination was accomplished over MnFe0.5Sm1.5O4

photocatalyst in 50 min with k value of 0.2038 min-1 and the colour removal proves the

photomineralisation of RB. Furthermore, the first-rate photocatalytic degradation of RB was

achieved for 0.08 g of MnFe0.5Sm1.5O4 photocatalyst concentration. This article demonstrate

that the smaller particle size, efficient separation of charge carriers (e-/h+) and highly

magnetic in nature of MnFe0.5Sm1.5O4 nanophotocatalyst greatly assisted for its enhanced

photocatalytic activity in the current study.

Acknowledgement

One of the authors, S. K. Rashmi expresses their gratitude for the University Grant

commission (UGC), New Delhi for providing RGNF (JRF-RGNF-2015-17-SC-KAR-12376)

and SAIF, IIT Bombay for facilitating HRTEM images. The author also acknowledges IISc

(Bangalore) for FESEM and other spectral data of the samples.

Reference
24

[1] P.A. Carneiro, R.F. Pupa Nogueira, M.V.B. Zanoni, Homogeneous photodegradation

of C.I. Reactive Blue 4 using a photo-Fenton process under artificial and solar

irradiation, Dyes Pigments. 74 (2007) 127.

[2] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide, Prog.

Solid State Chem. 32 (2004) 33.

[3] E. Yu. Yuzik-Klimova, N. V. Kuchkina, S. A. Sorokina, D. G. Morgan, B. Boris, L.

Zh. Nikoshvili, N. A. Lyubimova, V. G. Matveeva, E. M. Sulman, B. D. Stein, W. E.

Mahmoud,eA. A. Al-Ghamdi, A. Kostopoulou, A. Lappas, Z. B. Shifrina, L. M.

Bronstein, Magnetically recoverable catalysts based on polyphenylenepyridyl

dendrons and dendrimers, RSC Adv. 4 (2014) 23271-23280.

[4] Y.A. Kabachi, A.S. Golub, S.Y. Kochev, N.D. Lenenko, S.S Abramchuk, M.Y.

Antipin, P.M. Valetsky, B.D. Stein, W.E. Mahmoud, A.A Al-Ghamdi, L.M.

Bronstein, Multifunctional nanohybrids by self-assembly of monodisperse iron oxide

nanoparticles and nanolamellar MoS2 plates, Chem. Mater. 25 (2013) 2434-2440.

[5] S.H. Gage, B.D. Stein, L.Z. Nikoshvili, V.G. Matveeva, M.G. Sulman, E.M. Sulman,

D.G. Morgan, E.Y. Yuzik-Klimova, W.E. Mahmoud, L.M. Bronstein,

Functionalization of monodisperse iron oxide NPs and their properties as

magnetically recoverable catalysts, Langmuir 29 (2012) 466-473.

[6] W.E. Mahmoud, F. Al-Hazmi, F. Al-Noaiser, A.A. Al-Ghamdi, L.M. Bronstein, A

facile method to syntheses monodisperse γ-Fe2O3 nanocubes with high magnetic

anisotropy density, Superlattices Microstruct. 68 (2014) 1-5.

[7] R. Easterday, O. Sanchez-Felix, Y. Losovyj, M. Pink, B. D. Stein, D. G. Morgan, M.

Rakitin, V.Yu. Doluda, M. G. Sulman, W. E. Mahmoud, A.A. Al-Ghamdi, L.M.

Bronstein, Design of ruthenium/iron oxide nanoparticle mixtures for hydrogenation of

nitrobenzene, Catal. Sci. Tech. 5 (2015) 1902-1910.


25

[8] R. Easterday, C. Leonard, O. Sanchez-Felix, Y. Losovyj, M. Pink, B.D. Stein, D.G.

Morgan, N. A. Lyubimova, L.Z. Nikoshvili, E.M. Sulman, W.E. Mahmoud, A.A. Al-

Ghamdi, L.M. Bronstein, Fabrication of magnetically recoverable catalysts based on

mixture of Pd and iron oxide nanoparticles for hydrogenation of alkyne alcohols, ACS

Appl. Mater. Interfaces 6 (2014) 21652-21660.

[9] A. Goyal, S. Bansal, S. Singhal, Facile reduction of nitrophenols: Comparative

catalytic efficiency of MFe2O4 (M = Ni, Cu, Zn) nano ferrites, Int. J. Hydrogen

Energy 39 (2014) 4895-4908.

[10] X. Li, Y. Hou, Q. Zhao, L. Wang, A general, one-step and template-free synthesis of

sphere-like zinc ferrite nanostructures with enhanced photocatalytic activity for dye

degradation, J. Colloid Interface Sci. 358 (2011) 102–108.

[11] X. Zuo, A. Yang, S. Yoon, J. A. Christodoulides, V. G. Harris, C. Vittoria, Large

induced magnetic anisotropy in manganese spinel ferrite films, J. Appl. Phys., 87

(2005) 152505.

[12] P.Z. Guo, G.L. Zhang, J.Q. Yu, H.L. Li, X. S. Zhao, Controlled synthesis, magnetic

and photocatalytic properties of hollow spheres and colloidal nanocrystal clusters of

manganese ferrite, Colloids Surf. A Physicochem. Eng. Asp .395 (2012) 168-174.

[13] X. Liu, Z. Zhang, W. Shi, Y. Zhang, S. An, L. Zhang, Adsorbing Properties of

Magnetic Nanoparticles Mn-Ferrites on Removal of Congo Red from Aqueous

Solution, J. Dispersion Sci. Technol. 36 (2015) 462–470.

[14] S. Bhukal, S. Mor, S. Bansal, J. Singh, S. Singhal, Influence of Cd2+ ions on the

structural, electrical, optical and magnetic properties of Co–Zn nanoferrites prepared

by sol gel auto combustion method, J. Mol. Struct. 1071 (2014) 95–102.

[15] C. Singh, S. Jauhar, V. Kumar, J. Singh, S. Singhal, Synthesis of zinc substituted

cobalt ferrites via reverse micelle technique involving in situ template formation: A
26

study on their structural, magnetic, optical and catalytic properties, Mater. Chem.

Phys. 156 (2015) 188-197.

[16] V. T. Vader, Photocatalytic performance of fine particles of Cr doped magnesium

ferrites prepared by sol–gel combustion route, J. Mater. Sci. Mater. Electron. 26

(2015) 66–71.

[17] T.K. Pathak, N. H. Vasoya, Thillai Sivakumar Natarajan, K.B.Modi, Rajesh J.Tayade,

Photocatalytic Degradation of Aqueous Nitrobenzene Solution Using Nanocrystalline

Mg-Mn Ferrites, Mater. Sci. Forum 764 (2013) 116-129.

[18] P.P. Hankare, A. V. Jadhav, R.P. Patil, K. M. Garadkar, I. S. Mulla, R. Sasikala,

Photocatalytic degradation of rose bengal in visible light with Cr substituted MnFe2O4

ferrospinel, Arch. Phys. Res., 3 (2012) 269-276.

[19] L. Zhao, H. Yang, L. Yu, Y. Cui, X. Zhao, S. Feng, Study on magnetic properties of

nanocrystalline La-, Nd-, or Gd-substituted Ni–Mn ferrite at low temperatures, J.

Magn. Magn. Mater. 305 (2006) 91-94.

[20] S.E. Jacobo, S. Duhalde, H.R. Bertorello, Rare earth influence on the structural and

magnetic properties of NiZn ferrites, J. Magn. Magn. Mater. 272 (2004) 2253-2254.

[21] Z. Peng, X. Fu, H. Ge, Z. Fu, C. Wang, L. Qi, H. Miao, Effect of Pr3+ doping on

magnetic and dielectric properties of Ni–Zn ferrites by “one-step synthesis”, J. Magn.

Magn. Mater. 323 (2011) 2513-2518.

[22] C.B. Kolekar, P.N. Kamble, A.S. Vaingankar, X-ray, far IR characterization and

susceptibility study of Gd3+ substituted copper-cadmium ferrites, Ind. J. Phys. A 68

(1994) 529.

[23] K.N. Harish, H.S. Bhojya Naik, P.N. Prashanth kumar, R. Viswanath, Optical and

Photocatalytic Properties of Solar Light Active Nd-Substituted Ni Ferrite Catalysts:

For Environmental Protection, ACS Sustainable Chem. Eng. 1 (2013) 1143-1153.


27

[24] S. Thankachan, M. Kurian, D.S. Nair, S. Xavier, E.M. Mohammed, Effect of rare

earth doping on structural, magnetic, electrical properties of magnesium ferrite and its

catalytic activity, IJIRSET. 3 (2014) 529-537.

[25] J. Gao, Y. Zhao, W. Yang, J. Tian, F. Guan, Y. Ma, Y. Wang, Preparation of

samarium oxide nanoparticles and its catalytic activity on the esterification, J. Mater.

Chem. Phys. 77 (2002) 65–69.

[26] T.J. Shinde, A.B. Gadkari, P.N. Vasambekar, Effect of Nd3+ substitution on structural

and electrical properties of nanocrystalline zinc ferrite, J. Magn. Magn. Mater. 322

(2010) 2777- 2781.

[27] Y.M.Z. Ahmed, E.M.M. Ewair, Z.I. Zaki, In situ synthesis of high density magnetic

ferrite spinel (MgFe2O4) compacts using a mixture of conventional raw materials and

waste iron oxide, J. Alloys Compund. 489 (2010) 269-274.

[28] M.G. Naseri, E.B. Saion, H.A. Ahangar, M. Hashim, A.H. Shaari, Synthesis and

characterization of manganese ferrite nanoparticles by thermal treatment method, J.

Magn. Magn. Mater. 323 (2011) 1745-1749.

[29] P. Hu, H.B. Yang, D.A. Pan, H. Wang, J.J. Tian, S.G. Zhang, X.F. Wang, A. A.

Volinsky, Heat treatment effects on microstructure and magnetic properties of Mn–Zn

ferrite powders, J. Magn. Magn. Mater. 322 (2010) 173–177.

[30] C.H. Dong, G.X. Wang, L. Shi, D.W. Gou, C.J. Jiang, D.S. Xue, Investigation of the

thermal stability of Mn ferrite particles synthesized by a modified co-precipitation

method, Sci. China Phys. Mech. Astron. 56 (2013) 568-572.

[31] B.D. Cullity, Elements of X-ray Diffraction, Second ed., NewYork, (1978) 46.

[32] K.J. Standely, Oxide Magnetic Materials, 2nd edition, Clarendon Press, Oxford, 1972.
28

[33] M.M. Rashad, R.M. Mohamed, H. El-Shall, Magnetic properties of nanocrystalline

Sm-substituted CoFe2O4 synthesized by citrate precursor method, J. Mater. Process.

Technol. 198 (2008) 139- 146.

[34] E. Melagiriyappa, M. Veena, A. Somashekarappa, G.J. Shankaramurthy, H.S.

Jayanna, Dielectric behavior and ac electrical conductivity in samarium substituted

Mg–Ni ferrites, Indian J. Phys. 88 (2014) 795-80.

[35] W.V. Aulock, Handbook of microwaves ferrites Materials., Newyork, London, 1965.

[36] A.A Sattar, A.M. Samy, R.S. El-Ezza, A.E. Eatch, Effect of Rare Earth Substitution

on Magnetic and Electrical Properties of Mn–Zn Ferrites, Phys. Status Solidi A. 193

(2002) 86-93.

[37] J. Jiang, L.C. Li, F. Xu, Y.L. Xie, Preparation and magnetic properties of Zn–Cu–Cr–

Sm ferrite via a rheological phase reaction method, Mater. Sci. Eng. B 137

(2007)166–169.

[38] L.J. Zhao, H. Yang, L.X. Yu, Y.M. Cui, Effects of Gd2O3 on structure and magnetic

properties of Ni-Mn ferrite, J. Mater. Sci. 41(2006)3083–3087.

[39] A. Loganathan, K. Kumar, Effects on structural, optical, and magnetic properties of

pure and Sr-substituted MgFe2O4 nanoparticles at different calcination temperatures,

Appl Nanosci, 6 (2016) 629-639.

[40] S. Thankachan, B.P. Jacob, S. Xavier, E.M. Mohammed, Effect of samarium

substitution on structural and magnetic properties of magnesium ferrite nanoparticles,

J. Magn. Magn. Mater. 348 (2013) 140-145.

[41] N. Lwin, R. Othman, S. Sreekantan, M.A. Fauzi, Study on the structural and

electromagnetic properties of Tm-substituted Mg–Mn ferrites by a solution

combustion method, J. Magn. Magn. Mater. 385 (2015) 433–440.


29

[42] M.A. Khan, M.J. ur Rehman, K. Mahmood, I. Ali, M.N. Akhtar, G. Murtaza, I.

Shakir, M.F. Warsi, Impacts of Tb substitution at cobalt site on structural,

morphological and magnetic properties of cobalt ferrites synthesized via double

sintering method, Ceram. Int. 41 (2015) 2286-2293.

[43] S.M. Masoudpanah, S.A. Seyyed Ebrahimi, M. Derakhshani, S.M. Mirkazemi,

Structure and magnetic properties of La substituted ZnFe2O4 nanoparticles

synthesized by sol–gel autocombustion method, J. Magn. Magn. Mater. 370 (2014)

122–126.

[44] L.Y. Zhang, G.H. Zheng, Z.X. Dai, Structural, magnetic, and photoluminescence of

MnFe2O4:xEu3+ nanostructures, J Mater Sci: Mater Electron 27 (2016) 8138–8145.

[45] M. Hemeda, Structural and Magnetic Properties of Co_{0.6}Zn_{0.4}Mn_{x}Fe_{2

- x}O_{4}, Turk. J. Phys. 28 (2004)121–132.

[46] M.G. Naseri, M.K. Halimah, A. Dehzangi, A. Kamalianfar, E.B. Saion, B.Y. Majlis,

A comprehensive overview on the structure and comparison of magnetic properties of

nanocrystalline synthesized by a thermal treatment method, J. Phys. Chem. Solids 75

(2014) 315-327.

[47] K.K. Bamzai, G. Kour, B. Kaur, M. Arora, R.P. Pant, Infrared spectroscopic and

electron paramagnetic resonance studies on Dy substituted magnesium ferrite, J.

Magn. Magn. Mater. 345 (2013) 255–260.

[48] A. Sutka, M. Millers, M. Vanags, U. Joost, M. Maiorov, V. Kisand, I. Juhnevica,

Comparison of photocatalytic activity for different co-precipitated spinel ferrites, Res.

Chem. Intermed. 41 (2015) 9439-9449.

[49] J.W. Lekse, M.K. Underwood, J.P. Lewis, C. Matranga, Synthesis, Characterization,

Electronic Structure, and Photocatalytic Behaviour of CuGaO2 and CuGa1–xFexO2 (x

= 0.05, 0.10, 0.15, 0.20) Delafossites, J. Phys. Chem. C 116 (2012) 1865–1872.
30

[50] H. Lv, L. Ma, P. Zeng, D. Ke, T. Peng, Synthesis of floriated ZnFe2O4 with porous

nanorod structures and its photocatalytic hydrogen production under visible light, J.

Mater. Chem. 20 (2010) 3665-3672.

[51] P.S. Khiew, N.M. Huang, S. Radiman, M.S. Ahmad, Synthesis and characterization of

conducting polyaniline-coated cadmium sulphide nanocomposites in reverse

microemulsion, Mater. Lett. 58 (2004) 516-521.

[52] B.D. Viezbicke, S. Patel, B.E. Davis, D.P. Birnie, Evaluation of the Tauc method for

optical absorption edge determination: ZnO thin films as a model system, Phys. Status

Solidi 252 (2015) 1700-1710.

[53] P.H. Borse, J.S. Jang, S.J. Hongand, J.S. Lee, J.H. Jung, T.E. Hong, C.W. Ahn, E.D.

Jeong, K.S. Hong, J.H. Yoon, H.G. Kim, Photocatalytic Hydrogen Generation from

Water-methanol Mixtures Using Nanocrystalline ZnFe2O4 under Visible Light

Irradiation, J. Kor. Phys. Soc. 55 (2009) 1472.

[54] M. Ishimaru, Y.I. Hirotsu, I.V. Afanasyev-Charkin, K.E. Sickafus, Atomistic

structures of metastable and amorphous phases in ion-irradiated magnesium aluminate

spinel, J. Phys.:Condens. Matter 14 (2002) 1237-1247.

[55] K. Fujihara, S. Izumi, T. Ohno, M. Matsumura, Time-resolved photoluminescence of

particulate TiO2 photocatalysts suspended in aqueous solutions, J.Photochem.

Photobiol., A 132 (2000) 99–104.

[56] S. Makishim, H. Yamamoto, T. Tomotsu , S. Shionoya, Luminescence Spectra of

Sm3+ in BaTiO3 Host Lattice, J. Phys. Soc. Jpn., 20 (1965) 2147-2151.

[57] E. Moreira, L. A. Fraga, M. H. Mendonca, O. C. Monteiro, Synthesis, optical, and

photocatalytic properties of a new visible-light-active ZnFe2O4–TiO2 nanocomposite

material, J Nanopart. Res. 14 (2012) 937.


31

[58] C.H. Liang, F.B. Li, C.S. Liu, J.L. Lu, X.G. Wang, The enhancement of adsorption

and photocatalytic activity of rare earth ions doped TiO2 for the degradation of

Orange I, Dyes Pigments 76 (2008) 477-484.

[59] C. Singh, S. Jauhar, V. Kumar, J. Singh, S. Singhal, Synthesis of zinc substituted

cobalt ferrites via reverse micelle technique involving in situ template formation: A

study on their structural, magnetic, optical and catalytic properties, Mater. Chem.

Phys. 156 (2015) 188-197.

[60] S. Horikoshi, A. Saitou, H. Hidaka, N. Serpone, Environmental Remediation by an

Integrated Microwave/UV Illumination Method. V. Thermal and Non thermal Effects

of Microwave Radiation on the Photocatalyst and on the Photodegradation of

Rhodamine-B under UV/Vis Radiation, Environ. Sci. Technol. 37 (2003) 5813-5822.

[61] Z. Chen, D. Li, W. Zhang, Y. Shao, T. Chen, M. Sun, X. Fu, Photocatalytic

Degradation of Dyes by ZnIn2S4 Microspheres under Visible Light Irradiation, J.

Phys. Chem.C 113 (2009) 4433-4440.

[62] S. Horikoski, A. Tokunaga, H. Hidaka, N. Serpone, Environmental remediation by an

integrated microwave/UV illumination method: VII. Thermal/non-thermal effects in

the microwave-assisted photocatalyzed mineralization of bisphenol-A, J. Photochem.

Photobiol., A 162 (2004) 33-40.

[63] S. Kaneco, M.A. Rahman, T. Suzuki, H. Katsumata, K. Ohta, Optimization of solar

photocatalytic degradation conditions of bisphenol A in water using titanium dioxide,

J. Photochem. Photobiol., A 163 (2004) 419-424.

[64] C. Zhao, M. Pelaez, X. Duan, H. Deng, K. O’Shea, D.Fatta-Kassinos, Role of pH on

photolytic and photocatalytic degradation of antibiotic oxytetracycline in aqueous

solution under visible/solar light: Kinetics and mechanism studies, Appl. Catal., B

134 (2013) 83-92.


32

[65] S. Feraru, A.I. Borhan, P. Samoila, C. Mita, S. Cucu-Man, A.R. Iordan , M.N.

Palamaru, Development of visible-light-driven Ca2Fe1−xSmxBiO6 double perovskites

for decomposition of Rhodamine 6G dye, J. Photochem. Photobiol., A 307 (2015) 1-

8.

[66] H. Chun, W. Yizhong, T. Hongxiao, Destruction of phenol aqueous solution by

photocatalysis or direct photolysis, Chemosphere 41 (2000) 1205-1209.

[67] J. Arana, J.M. Nieto, J.H. Melian, J.D. Rodriguez, O.G. Diaz, J.P. Pena, O. Bergasa,

C. Alvarez, J. Mendez, Photocatalytic degradation of formaldehyde containing

wastewater from veterinarian laboratories, Chemosphere 55 (2004) 893-904.

[68] A.I. Borhan, A.R. Iordan, M.N. Palamaru, Correlation between structural, magnetic

and electrical properties of nanocrystalline Al3+ substituted zinc ferrite, Mater. Res.

Bull. 48 (2013) 2549-2556.


33

Fig. 1.
34

Fig. 2.
35

Fig. 3.
36

Fig. 4.
37

Fig. 5.

Fig. 6.
38

Fig. 7.
39

Fig. 8.
40

CB Reduction
e-
O2
e-
Substitution
MnFe0.5 Sm1.5O4 O2• 

1.64 eV

H2O
+ +
h h VB
OH• Degraded product
Oxidation
Rose Bengal

Scheme 1
41

Fig. 9.

Fig. 10.
42

Fig. 11.

FIGURE CAPTIONS

Fig. 1. (a) Powder X-ray diffraction pattern of MnFe2O4 calcinated at different temperature

(673, 873, 973 and 1373 K). (b) MnFe2-xSmxO4 (x = 0.0, 0.5, 1.0, 1.5, 2.0) nanoparticles at

973 K. (c) MnFe0.5Sm1.5O4 nanoparticle obtained at 673, 873 and 973 K.

Fig. 2. FESEM micrographs of MnFe2-xSmxO4 nanoparticles calcinated at 973 K; (a & b)

MnFe2O4, (c & d) MnFe0.5Sm1.5O4.

Fig. 3. EDS spectra of MnFe2-xSmxO4 nanoparticles calcinated at 973 K; (a) x = 0, (b) x =

0.5, (c) x = 1.0 and (d) x=1.5 samples.

Fig. 4. TEM images of MnFe0.5Sm1.5O4 nanoparticles calcinated at 973 K with different

magnification (a-b), HRTEM images and SAED pattern (c, d).


43

Fig. 5. The FTIR absorption spectra of MnFe2O4 (a) and MnFe0.5Sm1.5O4 (b) ferrite

nanoparticles calcinated at 973 K

Fig. 6. (a) Optical absorption spectra and (b) optical band gap from plots of (αhν) 2 vs. hν to

α=0 of MnFe2-xSmxO4 photocatalyst calcinated at 973 K.

Fig. 7. Room-temperature photoluminescence spectra of MnFe2O4 and MnFe0.5Sm1.5O4

annealed at 973 K.

Fig. 8. Time dependence photodegradation of RB using MnFe2-xSmxO4 photocatalyst under

visible light illumination (a-e), degradation percentage with concentration of samarium (f).

Scheme. 1. Photocatalytic mechanism involved in the degradation of RB molecule.

Fig. 9. (a) Degradation of RB (10 mg/L) with different MnFe2-xSmxO4 photocatalyst under

direct sun light illumination: (A) x=0.0, (B) x=0.5, (C) x =1.0 (D) x =1.5, (E) x =2.0; (b)

Photodegradation kinetics of RB intended with different MnFe2-xSmxO4 photocatalyst.

Fig. 10. (a) Effect of MnFe0.5Sm1.5O4 photocatalyst concentration on degradation of RB (10

mg/L) under direct sun light illumination (b) Photodegradation kinetics of RB intended for

MnFe0.5Sm1.5O4 photocatalyst under different concentration.

Fig. 11. (a) Comparison of percent degradation of different dyes (10.0 mg L−1) with

MnFe0.5Sm1.5O4 photocatalyst. (b) Photocatalytic degradation of RB in five cycles using

MnFe0.5Sm1.5O4 photocatalyst.
44

Table 1

MnFe2O4 nanoparticles Crystallite size (nm) Lattice parameter (Å)

673 K 39.55 8.349

873 K 45.03 8.356

973 K 59.75 8.423

1373 K 62.99 8.545

Table 1 Average crystallite size and lattice parameter of MnFe2O4 nanoparticles.

Table 2

x A-site B-site rA (Å) rB (Å) d(Å) aexp (Å) ath (Å) D (nm) u (Å)

0.0 Mn0.75Fe0.25 Mn0.25Fe1.75 0.672 0.652 2.519 8.356 8.325 59.75 0.3865

0.5 Mn0.75Fe0.25 Mn0.25Sm0.5Fe1.25 0.672 0.731 2.702 8.964 8.536 58.48 0.3783

1.0 Mn0.8Fe0.2 Mn0.2Sm1Fe0.8 0.673 0.809 2.718 9.017 8.747 48.21 0.3778

1.5 Mn0.8Fe0.2 Mn0.2Sm1.5Fe0.3 0.673 0.920 2.713 9.001 9.044 43.93 0.3767

2.0 Mn1 Sm2 0.68 0.964 2.746 9.108 9.164 38.79 0.3786

Table 2 The rA, rB, d, aexp, ath, D and u values of MnFe2-xSmxO4 nanoparticles.
45

Table 3

Organic Dye Illumination light Molecular structure Photodegradation


yield (%)
Rose Bengal (RB) Solar 97.8

Methyl Orange (MO) Solar 95.4

Methyl Red (MR) Solar 96.9

Methylene Blue (MB) Solar 97.1

Rhodamine B (RhB) Solar 94.2

Table 3 Summary of photodegradation yields for different dyes with MnFe0.5Sm1.5O4

photocatalyst under solar light illumination


46

Graphical Abstract
47

Highlights

1. MnSmxFe2-xO4 nanoparticles were synthesised by co-precipitation method.

2. The insertion of Sm3+ ion in the MnFe2O4 spinel matrix shows an increase in lattice

parameter.

3. The calcination temperature is sensitive knob for tuning the optical properties of the

MnFe2-xSmxO4 photocatalyst.

4. The MnFe2-xSmxO4 calcinated at 973 K shows enhanced optical behaviour.

5. The composition MnFe0.5Sm1.5O4 showed substantial red shift in absorption edge with

narrow band gap of 1.64 eV.

S-ar putea să vă placă și