You are on page 1of 6

Journal of Alloys and Compounds 601 (2014) 201206

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Magnetic property study of Gd doped TiO2 nanoparticles


Susmita Paul, Biswajit Choudhury, Amarjyoti Choudhury
Department Of Physics, Tezpur University, Napaam 784028, Tezpur, Assam, India

a r t i c l e

i n f o

Article history:
Received 22 December 2013
Received in revised form 6 February 2014
Accepted 14 February 2014
Available online 22 February 2014
Keywords:
Oxide materials
Solgel processes
Exchange and superexchange
Impurities in semiconductors
Luminescence
Magnetic measurements

a b s t r a c t
In this work we have studied the magnetic properties of solgel synthesized Gd doped TiO2 nanoparticles.
The Gd concentration varying from 0.03 to 0.07 mol. Structural, morphological and compositional analyses have been monitored with X-ray diffraction, transmission electron microscope (TEM), Raman spectroscopy and energy dispersive X-ray (EDX) spectroscopy. XPS spectra establish that Gd ions are in the +3
oxidation state. Photoluminescence intensity enhances at 0.03 and 0.05 mol and then quenches at 0.07.
This is likely due to the formation of emission quenching centers. All the samples exhibit paramagnetism
at room temperature as well as at 10 K. It is observed that due to the shielding of 4f shell of Gd3+ ions by
6s5d shell the direct exchange interaction of these Gd3+ ions with other Gd3+ ions is weak. These noninteracting and localized magnetic spins of Gd3+ induce only paramagnetism. The high magnetization
exhibited by the samples at 10 K is due to minimization of the thermal randomization of the magnetic
spins. Antiferromagnetic interaction persists at 0.03 mol and it gets stronger at 0.07 mol. Antiferromagnetism appears due to strong superexchange interaction of Gd3+ ions via O2 ions.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Diluted magnetic semiconductors (DMS) represent a synergetic
collaboration between charge based semiconductors and spin
based magnetism [13]. Following the success of (Ga, Mn) As,
many DMS materials have been discovered which exhibit room
temperature ferromagnetism (RTMF). In particular, theoretical
studies predicted and experiments veried that some wide band
gap semiconductors such as ZnO and TiO2 doped with transition
metals are effective DMS [4]. Magnetism in oxide based DMS are
often inadequate owing to low doping concentration and poor conductivity, both disfavoring the carrier based mechanism [5]. Furthermore, precipitates of doped magnetic elements and
inadvertent contaminations complicate experimental ndings.
TiO2 is a transparent wide band gap semiconductor. Transition
metal doping has been proposed to introduce magnetism in TiO2
nanoparticles [6]. The magnetic properties of doped TiO2 are
strongly sensitive to preparation condition, morphology and presence of defects. Bhattacharya et al. synthesized Mn doped TiO2
nanoparticles by one pot RAPET technique (reaction under autogenic pressure at elevated temperature) and reported ferromagnetic behavior up to 1%. However, with the increase in dopant
concentration up to 5% and 7% the system showed paramagnetic
behavior [7]. Peng et al. observed ferromagnetism in Mn doped
Corresponding author. Tel.: +91 3712267120; fax: +91 371222345.
E-mail address: ajc@tezu.ernet.in (A. Choudhury).
http://dx.doi.org/10.1016/j.jallcom.2014.02.070
0925-8388/ 2014 Elsevier B.V. All rights reserved.

TiO2 nanoparticles prepared by a simple solgel method [8]. Sellers


and Seebauer reported room temperature ferromagnetism in Mn
doped TiO2 nanopillars prepared by atomic layer deposition method [9].Thus, synthesis condition and morphology has a profound
effect in modulating the magnetic property in TiO2. Presence of defects and annealing condition is also reported to affect the magnetism. Choudhury et al. observed paramagnetism in air annealed co
doped TiO2 nanoparticles while the same material exhibited ferromagnetism on calcination under vacuum. This contrasting nature
of the magnetic property in the same system under two different
annealing conditions is attributed to the formation of high density
of oxygen defects under vacuum annealing [10].
Recently rare-earth doped DMS have caught up great interests
because of their unique uorescence properties and due to the high
emission quantum yields. In transition metal magnetic moment
arises from partially lled outermost 3d electrons, whereas in rare
earth ions magnetic moment arises from the inner 4f incomplete
sub-shell. In transition metal ions the strength of the spinorbital
coupling is much smaller than the crystal eld energy and hence,
the orbital momentum is quenched and only spin moment contributes towards magnetism [11]. In contrast in rare earth ions the 4f
electrons are shielded by the 5d6s electrons and the strength of
spinorbital coupling is much greater than the crystal eld energy
[11]. These 4f electrons undergo indirect exchange interactions via
5d or 6s conduction electrons yielding high magnetic moment per
atoms due to high orbital momentum [12]. Theoretical analysis
have been carried out to explain the ferromagnetism in Gd doped

202

S. Paul et al. / Journal of Alloys and Compounds 601 (2014) 201206

GaN [13,14]. Dalpian and Wei reported that the direct coupling between the Gd atoms was antiferromagnetic in nature in Gd doped
GaN [13]. The ferromagnetic phase was stabilized by the electrons
due to the mixing of Gd f with host s states. Liu et al., however, reported that the room temperature ferromagnetism in Gd doped
GaN was due to the interaction of Gd 4f spins via pd coupling
involving holes introduced by the intrinsic defects [14].
In the present study our objective is to investigate the structural, morphological and magnetic properties of TiO2 nanoparticles
due to doping of gadolinium. The structural and morphological
characterizations are done with X-ray diffraction, Raman, HRTEM.
The valence state of Gd in the host matrix is characterized by XPS.
This manuscript describes the magnetism in Gd doped TiO2 nanoparticles. To the best of our knowledge there are no reports on
magnetism in Gd doped TiO2 nanoparticles. Unlike transition metal
doped TiO2 where magnetism is initiated by dopants and oxygen
defects, the similar situation is not encountered in this system.
Although Gd is ferromagnetic, doping of this ion does not initiate
ferromagnetism in TiO2. It is observed that the unpaired spins in
Gd 4f does not undergo direct exchange interaction with nearest
Gd3+ ions nor do they take part in carrier mediated ferromagnetism. We have observed paramagnetism which is due to the presence of localized non-interacting spins present in 4f shell of Gd3+.
2. Experimental methods
2.1. Material synthesis
Gd doped TiO2 nanoparticles with three different nominal concentrations of Gd
(0.03 mol, 0.05 mol and 0.07 mol) were synthesized employing a simple solgel
method. The precursors for dopant and host were taken to be gadolinium nitrate
hexahydrate and titanium iso-propoxide respectively. 5 ml of titanium isopropoxide and 15 ml of 2 propanol were added to a 100 ml conical ask under constant
stirring, followed by the addition of 1 ml of water to initiate the hydrolysis reaction.
To the white sol of titanium isopropoxide, the dopant precursor solution were
added and stirred for 78 h. During such process rst a sol was formed which ultimately transformed into a gel. The gel was then centrifuged in water followed by
ethanol for 4 times. The centrifuged product was dried at 80 C. The resulting product was nally annealed at 450 C to obtain crystalline gadolinium doped anatase
TiO2 nanoparticles.
2.2. Characterization details
The structure of all the samples are determined using RigakuMiniex CD 10041
XRD unit with copper target and k = 0.154 nm at a scanning rate of 1/min and in
the scanning range of 1080. High resolution transmission electron microscope
images for morphology and particle size determination are observed with JEM2100, 200 kV JEOL. The elemental compositions are known from energy dispersive
X-ray analysis (EDX) equipped with a JEOL JSM 6390 LV scanning electron microscope (SEM). Raman spectra of the samples are acquired with Renishaw in via Spectrometer. An Argon-ion laser of 514.5 nm is used as the excitation source. The
oxidation state of the samples are studied with the help of X-ray photoelectron
spectroscopy (XPS) is recorded on KRATOS-AXIS 165 instrument equipped with
dual aluminiummagnesium anodes using the Mg Ka radiation (hm 1253:6 eV)
operated at 5 kV. Photoluminescence (PL) measurements at room temperature
are recorded with Perkin Elmer LS 55 uorescence spectroscopy. Magnetic measurements are carried out in a vibrating sample magnetometer (VSM) using a selection of Varian and Walker electromagnets.

3. Results and discussions


3.1. Structural and morphological study
The phase structure is investigated by XRD and the results are
shown in Fig. 1. All the diffraction peaks correspond to the tetragonal anatase phase (JCPDS-782486). A small amount of brookite
phase is detected in between 27 and 31 (JCPDS-761934). The
sizes (d) of the nanocrystal are calculated by using Scherrers
formula

0:9k
bCosh

Fig. 1. X-ray diffraction spectra of pure and Gd doped TiO2 nanoparticles.

where k = wavelength of the X-ray source used; b full width at


half maximum (FWHM), h = Braggs diffraction angle. The crystallite
size is calculated to be 7.8 nm for pristine (0.00 mol) and 5.9 nm,
5.45 nm, 5.42 nm for 0.03, 0.05 and 0.07 mol of Gd respectively.
The width of the diffraction peaks increases with the increase in
doping concentration. The widening of the diffraction peaks indicates systematic decrease in grain size and degradation of the structural quality after doping. Because of the moderately large size
incongruity with Ti4+ (0.68 ), Gd3+ (0.94 ) ions are not expected
to inhabit sites in the lattice of TiO2 [15]. They possibly stay on
the particle surface and on the grain boundaries and produce a
strain in those regions. This result in the enhancement of the repulsive interaction between the Gd3+ ions, prevent the coalescence of
nanocrystallites and thereby inhibits the growth by the formation
of GdOTi bond [16]. It is observed that with increase in Gd doping
concentration the intensity of diffraction peak is increasing. The diffraction peak intensities are affected by the change in electron density due to substitution of foreign (dopant) atoms. Through doping
we change the element that is we change the mean atomic scattering factor. Doping with light atom might results in lowering of diffraction intensity. However doping with a larger Gd3+ results in
more scattering centers and hence, results in increased peak intensity [20]. The diffraction pattern does not show any peak of secondary phases of gadolinium such as Gd2O3. However, slight shifting of
the peaks to lower angle is noticed with the changes in the gadolinium concentration. We have also calculated the d spacing of all the
samples considering (0 0 4) peak. The values are calculated to be
0.23586, 0.23595, 0.23615 and 0.23622 nm for undoped and
(0.03, 0.05, 0.07 mol) Gd doped TiO2 nanoparticle. The change is d
spacing is not prominent inferring that very few percentage of the
total gadolinium impurity ions are actually substituting Ti4+ on
the lattice interior and the others may be located on the surface.
The transmission electron microscope images of the undoped
and Gd doped TiO2 nanoparticles are shown in Fig. 2(ad). As observed from the micrographs, the undoped system comprises of
spherical size particles of 18 nm size as shown in Fig. 2a. Fig 2b
shows the corresponding high resolution image of pure TiO2. The
distinct lattice fringes with inter planar spacing of 3.52 , corresponding to the (1 0 1) plane of anatase TiO2 displays highly crystalline nature of the samples [17]. Fig. 2c shows the TEM image
of 0.03 mol Gd doped TiO2. The average particle size is calculated
to be 12 nm. The interplanar spacing shown as shown in Fig. 2d
is found to be 3.57 . The interplanar spacing of the doped and
undoped TiO2 nanoparticles do not show much variation, further
conrming incorporation of only few gadolinium ions on the TiO2
lattice. It may be mentioned here that we have obtained

S. Paul et al. / Journal of Alloys and Compounds 601 (2014) 201206

203

Fig. 2. TEM images of (a) undoped (c) 0.03 mol Gd doped TiO2; high resolution images showing the lattice fringes of (b) undoped and (d) doped TiO2 nanoparticles (e) EDX
analysis of 0.03 mol Gd doped TiO2 the doped sample.

differences in the size calculated from XRD and TEM. The magnitude of the crystallite and particle size compliments each other
in case of stable, monodisperse, single crystals and hard spheres
and deviation from this criteria leads to variation in measured sizes
[18]. A nanoparticle of polycrystalline aggregates compose of
grains of several sizes. Scherrers formula determines the size of
coherently diffracting grain size. Moreover, the size determined
by Scherrer formula is affected by instrumental broadening, crystalline strain, defects, etc. On the other hand TEM provide direct
evidences of the nanoparticle size, shape. Due to the polycrystalline nature containing several grain sizes the crystallite size is
not generally the same as the particle size [19,20]. The compositional analysis is performed with EDX analysis. The spectrum is
marked with the signals obtained from Ti, O, Gd.
Raman spectroscopy is used for characterizing local structure,
defects, crystallinity, etc. Fig. 3 shows the Raman spectra of pure
and Gd doped TiO2 nanoparticles. Group theory calculation predicts that anatase phase contains six active Raman modes (A1g + 2B1g + 3Eg) [21]. Ohsaka reported six allowed Raman active modes
for anatase TiO2 at 144 cm1 (Eg), 197 cm1 (Eg), 399 cm1 (B1g),
513 cm1 (A1g + B1g) and 639 cm1 [22]. From the Raman spectra

it is evident that both undoped and doped TiO2 are in anatase


phase and a small peak in the region of 240 cm1 and 320 cm1
are attributed to the brookite phase [23]. No impression of impurity related modes in the Raman spectra are seen in the doped samples. This is in agreement with the XRD result obtained. It has been
reported that the Eg is mainly caused by symmetric stretching
vibration of OTiO in TiO2, the B1g is caused by symmetric bending vibration of OTiO bond and the A1g is due to the antisymmetric bending vibration of OTiO bond [24]. In undoped TiO2 the Eg
peak appears at 144 cm1, whereas in 0.03 mol Gd doped TiO2 the
peak is shifted to 147 cm1. With the increase in Gd3+ ion concentration up to 0.05 mol and 0.07 mol the Eg peak position at
147 cm1, however, remains the same. In some transition and
other rare earth doped TiO2 nanoparticles it has been reported that
doping generates oxygen vacancies and shifts the peak position to
higher wavenumber [25,26]. But in our case the shifting of the peak
position to higher wavenumber is not occurring gradually at each
dopant concentration. This behavior may be due to the presence
of very few numbers of gadolinium ions substituting Ti4+ in the
TiO2 lattice. A large population of Gd3+ ions remains at the surface
and therefore, no appreciable shifting in the peak position is observed with increase in dopant concentration. The small peak shifting in the doped samples is likely due to the structural disorder
brought about by the substitution of few gadolinium ions in the
TiO2 lattice.

3.2. X-ray photoelectron spectroscopy (XPS) study

Fig. 3. Raman spectra of pure and Gd doped TiO2 nanoparticles.

The XPS spectrum of 0.07 mol Gd doped TiO2 is investigated to


know the oxidation state of the dopant ions. Ti 2p, O 1s and Gd 4d
core level of the highest doped samples is shown in Fig. 4ac. The
core level binding energies of Ti 2p3/2 and Ti 2p1/2 (Fig. 4a) are at
roughly 459.31 eV and 464.90 eV. The variation of 5.59 eV indicates a valence of +4 for Ti in TiO2 [27]. The core level XPS O 1s
spectra are also analysed. Fig. 4b displays the XPS spectrum of O
1s. The spectra species presence of at least two kinds of chemical
states that include crystal lattice oxygen (OL) and chemisorbed
oxygen (OH). The OL peak appears around 530.08 eV and the OH
peak at 531.32 eV [28]. In Fig. 4c, the Gd 4d3/2 peak at 143 eV corresponds to +3 oxidation state of Gd ion [29].

204

S. Paul et al. / Journal of Alloys and Compounds 601 (2014) 201206

Fig. 4. Core level X-ray photoelectron spectrum (XPS) of (a) Ti 2p (b) O 1s (c) Gd d3/2 for 0.07 mol Gd doped TiO2.

3.3. PL measurements
PL spectra are obtained by exciting the doped and undoped TiO2
at 320 nm. Fig. 5 shows the emission spectra of pure and Gd doped
TiO2 nanoparticles. The spectra are deconvoluted into ve emission
peaks with the help of Gaussian tting (r2 = 0.998). All the samples
exhibit a UV emission peak at 398 nm, violet emission peak at
429 nm, two blue emission peak at 457 nm and 488 nm and one
green emission peak at 537 nm. The UV emission peak is the band
edge emission of host TiO2 arising from X1b transition to C3 [30].
The 429 nm peak can be attributed to self trapped excitons (STE)
originated by interactions of conduction band electrons localized
on the Ti 3d orbital with holes present in the O 2p orbital of TiO2
[31]. The peaks at 457 nm and 537 nm are due to the oxygen vacancy related defect centers present below the conduction band.
The intrinsic defects in wide band gap semiconductors are expected to introduce deep energy levels inside the band gap
[31,32]. The signature of peak at 488 nm is due to charge transfer
2
transition from Ti3+ to oxygen anion in a TiO6 complex associated
with oxygen vacancy [33]. The emission spectrum does not contain
any Gd3+ related emission peaks and the emission spectra of doped
and undoped TiO2 are alike. It is the intensity of the emission peaks
which are affected on doping. Photoexcitation of Gd doped TiO2 at
320 nm results in above band gap excitation of electrons to the
upper lying band states in conduction band. These electrons rst
jump to the lower of the conduction band in non-radiative way.
From the lowest of the conduction band edge the electrons either
jump directly to the valence band or initially jump to defects or
to intermediate Gd states and then emit light. Non-existence of
Gd emission peak indicates that the Gd 4f levels are possibly at
the same height as that of oxygen defect states. The signicant contribution of Gd is seen in the intensity of the emission peaks. The
emission intensity gradually increases on incorporation of
0.03 mol and 0.05 mol Gd and then there is sudden quenching of
the emission intensity on adding 0.07 mol Gd. The emission intensity of 0.07 mol Gd doped TiO2 is even less than that of pure TiO2.
At lower dopant concentration the interactions between ions are
too weak to have an effect on the energy levels of each dopant
ion. But at higher dopant concentration the distance between ions
and defects decreases. Decrease in the distance increases the interactions between the ions. The emission energy is transferred from
one ion to another ions or defects and the energy is dissipated nonradiatively resulting in the quenching of emission intensity. We
have calculated the average shortest distance (known as the critical distance) between ions at which the emission energy transfer
occurs, using the equation [34].


RC 2

3V
4pxc N

Fig. 5. (a) Room temperature photoluminescence spectra of undoped and Gd doped


TiO2 nanoparticles. (b) Deconvolution of the emission spectra into several emission
peaks by Gaussian tting.

N = Number of Z ions in the unit cell, Rc = the critical transfer distance. Putting the values the critical distance (Rc) for 0.07 mol is calculated to be 0.9731 nm.
3.4. Magnetic measurements
Fig. 6 shows the MH curves of 0.03, 0.05 and 0.07 mol Gd
doped TiO2 nanoparticle in the magnetic eld range of 15 kOe.
The curves show the typical behavior of paramagnetism. There is
a negligible increase in the magnetization with increase in Gd up
to 0.05 mol. The magnetization of 0.07 mol is less than that of
0.05 mol. The observed paramagnetism might be due to the presence of isolated Gd3+ ions, whereas reduction in the magnetization
is most likely due to the presence of antiferromagnetic interaction
in the system. While incorporating Gd into TiO2, room temperature
ferromagnetism with high magnetization was expected in the system. The notion behind this expectation is that Gd is ferromagnetic
in nature with persistence of ferromagnetic ordering up to 289 K

1=3
2

V = volume of the unit cell = a2c = 0.135 nm3, where a = 0.3785 nm;
c = 0.9424 nm. xc = critical concentration of the dopant (= 0.07),

Fig. 6. Room temperature MH curves of Gd doped TiO2 nanoparticles.

S. Paul et al. / Journal of Alloys and Compounds 601 (2014) 201206

[35]. The electronic conguration of Gd3+ is [Xe] 6s25d14f7 with 7


unpaired electrons in the 4f shell. These unpaired 4f electrons
polarize the 6s and 5d valence electrons resulting in a magnetic
moment of 7.63 lB/Gd3+ [36]. However, because of the shielding
of this 4f shell by 6s5d valence shell electrons, the interaction of
these unpaired spins with the outermost ligands or other Gd3+ is
expected to be weak. These non-interacting and localized magnetic
spins of Gd3+ induce only paramagnetic moment [35,36]. Paramagnetism at room temperature has been reported in pure Gd2O3 as
well as in Gd doped oxide semiconductors such as TiO2, SnO2,
and ZnO [3540]. Opera et al. in his work on Gd doped ZnO reported ferromagnetism and attributed it to the intra-ion 4f5d exchange interaction followed by inter-ion 5d5d coupling mediated
by charge carriers [41]. In Gd doped ZnO lms, Ney et al. reported
paramagnetic behavior of the samples with magnetic moment of
3
7.63 lB =Gd with a small lattice distortion, the results were in
agreement with the rst-principles density-functional theory calculations. They detected presence of secondary oxide phases of
Gd2O3 or other antiferromagnetic phase of Gd3+ contributing to
the reduction in the paramagnetic behavior [42]. Adhikari et al.
in Gd doped SnO2 observed absence of ferromagnetism due to surface spin effect and enhanced GdOGd interactions [39]. Magnetic properties of rare earth doped (RE = Nd, Sm, Gd, Tb, Er and
Dy) CeO2 have been investigated by Dimri et al. [43]. The origin
of magnetism in this sample is reported due to the presence of oxygen vacancies created due to the rare-earth dopants [43]. Wang
et al. studied the magnetic property of pure and Gd doped HfO2
and reported very weak ferromagnetism in the sample, attributing
it to either impure target materials or signal from the substrates
[44]. For further conrmation of the nature of magnetism in the
system, we carried out temperature dependent magnetic measurements (MT) for 0.03 mol and 0.07 mol Gd doped TiO2. Fig. 7ab
shows the susceptibility (v) vs. temperature (T) curves in the temperature range of 10300 K at an applied eld of 200 Oe. The zero
eld cooling (ZFC) and eld cooling (FC) curves are completely
overlapped up to 300 K and the nature of the curves follow the CurieWeiss behavior of paramagnetism. The rise in magnetization at
low temperature is because of the decrease in the thermal randomization of uctuating magnetic spins present in isolated Gd3+ ions.
The temperature dependent susceptibility (v) curves were tted
with modied CurieWeiss law-

v v0

C
T h

where v0 is diamagnetic susceptibility, C is Curie constant and h is


CurieWeiss temperature. For 3 mol% Gd doped TiO2 the Curie
Weiss tting gives the following results, v0 = 8.37  106,
C = 0.0018, h = () 0.2044. For 0.07 mol Gd doped TiO2 the tted values for v0 , C and h are 2.96  106, 0.0019 and () 0.3025

205

Fig. 8. MH curves of 0.03 mol and 0.07 mol Gd doped TiO2 measured at 10 K.

respectively. The small negative value of CurieWeiss temperature


indicates presence of antiferromagnetic interaction in the system.
As compared to 0.03 mol Gd, the value of h is higher at 0.07 mol.
The large negative value of h conrms large antiferromagnetic interaction among dopant ions at higher doping level. Therefore, from
the results of MH and MT it is anticipated that both paramagnetic
and antiferromagnetic phases are present in the system. During discussion on XRD and Raman it is mentioned that with increase in
doping concentration, a large number of Gd3+ resides on the surface.
Doping at higher level decreases the separation among Gd3+ ions.
During discussion on PL it was stated that the distance between
Gd3+ ions is 0.9731 nm when Gd3+ doping level is 0.07 mol. These
Gd3+ ions at this distance apart may undergo superexchange interaction with each other via O2 ions. This results in antiparallel
alignment of the magnetic spins of Gd 4f shell present in the nearest
neighbor. The antiferromagnetic interaction is even stronger at
0.07 mol and that is why the CurieWeiss temperature is more negative at 0.07 mol than at 0.03 mol. We also carried out magnetic
measurements of 0.03 and 0.07 mol Gd doped TiO2 at 10 K (Fig. 8)
and at magnetic eld of 45 kOe. It is observed that even at 10 K
and at such a high applied eld the nanoparticles of Gd doped
TiO2 do not show any ferromagnetic ordering. The magnetic curves
are nearly linear with magnetic eld exhibiting the characteristic of
paramagnetism. However, unlike the linear MH curves of the samples measured at room temperature, the magnetic curves at 10 K
are trying to get the shape of hysteresis. But the magnetization is
not saturated even at 45 KOe. The slight bent in the MH curve is
likely due to the minimization of the thermal randomization of
the magnetic spins at 10 K. It if further noticed that the magnetization at 0.07 mol is slightly higher than at 0.03 mol. Decrease in the
thermal randomization and contribution of large numbers of

Fig. 7. Susceptibility (v) vs. temperature (T) curves (a) 0.03 mol and (b) 0.07 mol doped TiO2 nanoparticles.

206

S. Paul et al. / Journal of Alloys and Compounds 601 (2014) 201206

localized non-interacting magnetic spins of Gd3+ may be the reason


for the higher magnetization in 0.07 mol.
4. Conclusion
In summary, we have prepared Gd doped TiO2 nanoparticles by
solgel method with different concentration of Gd. X-ray diffraction and Raman studies reveal that Gd doping has little inuence
on the internal structure of TiO2. Photoluminescence intensity
quenches when TiO2 is loaded with 0.07 mol. This is likely due to
the formation of emission quenching centers. All the doped samples exhibit paramagnetism at room temperature. Due to the localized and non-interacting nature Gd3+ ions could not undergo
strong ferromagnetic exchange interaction with the nearest ions
and thus, induce only paramagnetism. The magnetization carried
out at 10 K tries to get the shape of hysteresis and attain very high
magnetization. This is expected due to the reduction of thermal
randomization of magnetic spins so that large numbers of localized
non-interacting magnetic spins of Gd3+ contribute towards paramagnetism. The samples exhibit negative CurieWeiss temperature and thus, it is believed that the samples exhibit
antiferromagnetic interaction. Presence of antiferromagnetic
ordering is because of the superexchange interaction of nearest
Gd3+ ions via O2 ions on the surface.
Acknowledgements
We are thankful to, Sophisticated Analytical Instrument Facility
(SAIF), NEHU, Shillong for helping us in providing the HRTEM
images, IIT Kanpur for helping us in providing the magnetic measurement results and IICT Hyderabad for XPS datas. We are also
thankful to UGC for nancial support to the Project F.No. 42-785/
2013 (SR).
References
[1] G.Z. Xing, J.B. Yi, J.G. Tao, T. Liu, L.M. Wong, Z. Zhang, G.P. Li, S.J. Wang, J. Ding,
T.C. Sum, C.H.A. Huan, T. Wu, Adv. Mater. 20 (2008) 35213527.
[2] S.A. Wolf, D.D. Awschalom, R.A. Buhrman, J.M. Daughton, S. von Molnar, M.L.
Roukes, A.Y. Chtchelkanova, D.M. Treger, Science 294 (2001) 14881495.
[3] I. Zutic, J. Fabian, S.D. Sarma, Rev. Mod. Phys. 76 (2004) 323410.
[4] Y. Matsumoto, M. Murakami, T. Shono, T. Hasegawa, T. Fukumra, M. Kawasaki,
P. Ahmet, T. Chikyow, S. Koshihara, H. Koinuma, Science 291 (2001) 854856.
[5] R. Janisch, P. Gopal, N.A. Spaldin, J. Phys. -Condens. Mater. 17 (2005) R657
R689.
[6] D. Chu, Y.P. Zeng, D. Jiang, Y. Masuda, Sci. Adv. Mater. 1 (2009) 227229.
[7] S. Bhattacharyya, A. Pucci, D. Zitoun, A. Gedanken, Nanotechnology 19 (2008)
495711 (8pp).
[8] D. Peng, L.F. Min, Z.C. Cang, Z.W. Wu, Z. Huan, C.L. Gang, Z.L. Gui, Chin. Phys. B
19 (2010) 118102 (6pp).

[9] M.C.K. Sellers, E.G. Seebauer, Mater. Lett. 114 (2014) 4447.
[10] B. Choudhury, A. Choudhury, A.K.M. Maidul Islam, P. Alagarsamy, M.
Mukherjee, J. Magn. Magn. Mater. 323 (2011) 440446.
[11] http://www.uni-stuttgart.de/gkmr/lectures/lectures_WS_0203/
crystaleld.PDF.
[12] G.M. Dalpian, S.-H. Wei, Phys. Rev. B 72 (2005) 115201115206.
[13] S. Dhar, O. Brandt, M. Ramsteiner, V.F. Sapega, K.H. Ploog, Phys. Rev. Lett. 94
(2005). 037205037205.
[14] L. Liu, P.Y. Yu, Z. Ma, S.S. Mao, Phys. Rev. Lett. 100 (2008) 127203127207.
[15] N.D. Abazovic, M.B. Radoicic, T.D. Savic, I.A. Jankovic, M.I. Comor, Dig. J.
Nanomater Biosci. 8 (2013) 871875.
[16] Q. Chen, T. Lu, M. Xu, C. Meng, Y. Hu, K. Sun, I. Shlimak, Appl. Phys. Lett. 98
(2011) 073103 (3pp).
[17] L. Nagaveni, M.S. Hegde, N. Ravishankar, G.N. Subbanna, G. Madras, Langmuir
20 (2004) 29002907.
[18] M. Baalousha, Y.J. Nam, P.A. Cole, B. Gaiser, T.F. Fernandes, J.A. Hriljac, M.M.A.
Jepson, V. Stone, C.R. Tyler, J.R. Lead, Environ. Toxicol. Chem. 31 (2012) 983
993.
[19] B. Choudhury, B. Borah, A. Choudhury, Mater. Sci. Eng. B Adv. 178 (2013)
239247.
[20] P.K. Sharma, R.K. Dutta, R.J. Choudhary, A.C. Pandey, Cryst. Eng. Commun. 15
(2013) 44384447.
[21] M. Popa, L. Diamendescu, F. Vasiliu, C.M. Teodorescu, V. Cosoveanu, M. Baia, M.
Feder, L. Baia, V. Danciu, J. Mater. Sci. 44 (2009) 358364.
[22] T. Ohsaka, F. Izumi, Y. Fujikim, J. Raman spectrosc. 7 (1978) 321324.
[23] M.P. Moret, R. Zallen, D.P. Vijay, S.B. Desu, Thin Solid Films 366 (2000)
810.
[24] F. Tian, Y. Zhang, J. Zhang, C. Pan, J. Phys. Chem. C 116 (2012) 75157519.
[25] S. Paul, P. Chetri, A. Choudhury, J. Alloys Comp. 583 (2014) 578586.
[26] M. Pal, U. Pal, J.M. Gracia, Y. Jimnez, F.P. Rodrguez, Nanoscale Res. Lett. 7
(2012) 112.
[27] C. Rath, P. Mohanty, A.C. Pandey, N.C. Mishra, J. Phys. D: Appl. Phys. 42 (2009)
205101 (6pp).
[28] L. Jhing, B. Xin, F. Yuan, L.P. Xue, B. Wang, H. Fu, J. Phys. Chem. B 110 (2006)
1786017865.
[29] D. Raiser, J.P. Deville, J. Electron Spectrosc. Relat. Phenom. 57 (1991) 9197.
[30] N. Serpone, D. Lawless, R. Khairutdinov, J. Phys. Chem. 99 (1995) 16646
16654.
[31] W.Y. Wu, Y.M. Chang, J.M. Ting, Cryst. Growth Des. 10 (2010) 16461651.
[32] N. Serpone, J. Phys. Chem. B 110 (2006) 2428724293.
[33] J.C. Yu, J. Yu, W. Ho, Z. Jiang, L. Zhang, Chem. Matter. 14 (2002) 38083816.
[34] D. Wang, Q. Yin, Y. Li, M. Wang, J. Lumin. 97 (2002) 16.
[35] H.T. Wong, H.L.W. Chan, J.H. Hao, Appl. Phys. Lett. 95 (2009) 022512.
[36] M. Petersen, J. Hafner, M. Marsman, J. Phys.: Condens. Matter. 18 (2006) 7021
7043.
[37] L.G. Jacobsohn, B.L. Bennett, R.E. Muenchausen, S.C. Tornga, J.D. Thompson, O.
Ugurlu, D.W. Cooke, A.L. Sharma, J. Appl. Phys. 103 (2008) 104303.
[38] H.S. Hafez, M. Saif, J.T. McLeskey Jr, M.S.A. Abdel-Mottaleb, I.S. Yahia, T. Story,
W. Knoff, Int. J. Photoenergy 2009 (2009) 240402.
[39] R. Adhikari, A.K. Das, D. Karmakar, J. Ghatak, J. Magn. Magn. Mater. 322 (2010)
36313637.
[40] S. Ghosh, G.G. Khan, K. Mandal, ACS Appl. Mater. Interfaces 4 (2012) 2048
2056.
[41] O. Oprea, O.R. Vasile, G. Voicu, L. Craciun, E. Andronescu, Dig. J. Nanomater
Biosci. 7 (2012) 17571766.
[42] V. Ney, S. Ye, T. Kammermeier, K. Ollefs, F. Wilhelm, A. Rogalev, S. Lebegue, A.L.
da Rosa, A. Ney, Phys. Rev. B 85 (2012) 235203235207.
[43] M.C. Dimri, H. Khanduri, H. Kooskora, J. Subbi, I. Heinmaa, A. Mere, J. Krustok,
R. Stern, Phys. Status Solidi A 209 (2012) 353358.
[44] W. Wang, Y. Hong, M. Yu, B. Rout, Gary A. Glass, J. Tang, J. App. Phys. 99 (2006)
(3pp).

You might also like