Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Vibrational Spectroscopy 58 (2012) 163168

Contents lists available at SciVerse ScienceDirect

Vibrational Spectroscopy
journal homepage: www.elsevier.com/locate/vibspec

Phonon and magnetic properties of nanocrystalline MnWO4 prepared by


hydrothermal method
a,
a

nski

Mirosaw Maczka
, Maciej Ptak a, Adam Pikul a, Leszek Kepi
, Pawe E. Tomaszewski a, Jerzy Hanuza a,b

a
b

Institute of Low Temperature and Structure Research, Polish Academy of Sciences, P.O. Box 1410, 50-950 Wrocaw 2, Poland
Department of Bioorganic Chemistry, Faculty of Industry and Economics, University of Economics, ul. Komandorska 118/120, 53-345 Wrocaw, Poland

a r t i c l e

i n f o

Article history:
Received 26 August 2011
Received in revised form
14 December 2011
Accepted 15 December 2011
Available online 24 December 2011
Keywords:
Infrared
Raman
Magnetization
Specic heat
Multiferroic

a b s t r a c t
Well-crystallized MnWO4 nanoparticles with 20100 nm sizes were synthesized by a mild hydrothermal
crystallization process. X-ray diffraction and TEM results show that particles have tendency to grow in a
shape of nanorods elongated in the c direction. They also show that the particles synthesized from MnCl2
have smaller concentration of defects than those synthesized from MnSO4 H2 O. Raman spectra revealed
weak changes with decreasing particle size due to phonon connement effect and defects. In contrast to
this behavior, IR spectra showed much more pronounced changes. Origin of this behavior is discussed.
Raman and IR spectra showed also presence of additional bands at 923 and 914 cm1 , respectively. These
bands have been attributed to presence of terminal W O bonds at the surface of MnWO4 nanocrystallites.
Specic heat and magnetization studies revealed broad anomalies at 11.011.4 K, which correspond to
the two magnetic transitions found for bulk MnWO4 at 12.5 and 13.5 K.
2011 Elsevier B.V. All rights reserved.

1. Introduction
In recent years, multiferroic magnetoelectric materials have
attracted much attention since a coexistence of magnetic and ferroelectric orders leads to novel physical effects [1]. Multiferroics
are also very attractive materials for a wide range of applications.
For instance, they can be used for fabrication of magnetoelectric sensors and memory chips [2,3]. In recent years, several new
multiferroic materials have been discovered such as DyMnO3
[4], Ni3 V2 O8 [5] and CuFeO2 [6]. Multiferroic properties have
also been discovered in some molybdates and tungstates such as
RbFe(MoO4 )2 [7] and MnWO4 [8].
The crystal structure of MnWO4 is monoclinic, space group
P2/c (#13, C2h 4 ), with the unit cell parameters: a = 4.830(1),
= 91.14 and Z = 2 [9]. It consists of
b = 5.7603(9), c = 4.994(1) A,
edge-sharing WO6 and MnO6 octahedra that form chains along the
c axis (see Fig. 1). The WO6 octahedra within a chain are connected
by a double oxygen bridge with shorter W O1 and longer W O1
respectively [9]. The remaining O2
distance of 1.9101 and 2.1377 A,
Manganese
oxygen atoms form short W O2 distance of 1.7847 A.

Corresponding author. Tel.: +48 713954161; fax: +48 713441029.

E-mail address: m.maczka@int.pan.wroc.pl (M. Maczka).


0924-2031/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.vibspec.2011.12.010

and tungsten atoms are arranged in alternating sheets parallel to


(1 0 0) [9].
MnWO4 exhibits three phase transitions at TN = 13.5 K,
T2 = 12.7 K and T1 = 6.58 K into antiferromagnetic ordered phases
AF3 , AF2 and AF1 [8,10,11]. The most interesting is the AF2 phase
since in this phase helical magnetic order induces improper ferroelectricity [8,10,11]. Although MnWO4 has been extensively
studied in recent years due to its multiferroic properties, it has also
been shown that MnWO4 is a promising material for humidity sensors [12] and high-gain Raman laser [13]. Moreover, it has attracted
attention due to its electrochemical properties [14].
It is well known that properties of materials are signicantly
modied when size of crystallites is of order of nanometers. Synthesis of nanocrystalline MnWO4 has been reported
in a few papers [13,1519]. It has been shown that nanocrystalline MnWO4 is more suitable for application as humidity
sensors [13]. The temperature dependence of the magnetic
susceptibility, measured for MnWO4 nanowires, revealed presence of a single and broad maximum near 7 K, which was
attributed to appearance of an antiferromagnetic order [18].
However, no features were found, which might correspond to
the three thermodynamic transitions found for bulk MnWO4
[18].
Present paper reports synthesis of nanocrystalline MnWO4
with different size of the crystallites and characterization of thus


M. Maczka
et al. / Vibrational Spectroscopy 58 (2012) 163168

164

ultrasonic agitation. A droplet of suspension was deposited on a


microscope grid covered with carbon lm.
Infrared spectra were measured with a Biorad 575C FT-IR spectrometer in Nujol suspension. Raman spectra were measured using
Bruker FT 100/S spectrometer with the YAG:Nd laser excitation
(1064 nm). Both IR and Raman spectra were recorded with a spectral resolution of 2 cm1 .
Magnetization and specic heat of the nanocrystalline sample
synthesized at 180 C were investigated employing commercial
Quantum Design MPMS and PPMS platforms, respectively. The
magnetization was measured for a powder sample freely placed
in a diamagnetic gel capsule. For calorimetric experiments a
powder sample was used together with small amount of vacuum grease Apiezon N as a sample-mounting medium. While the
magnetization of the sample holder used in magnetic properties
measurement was found to be negligible in comparison to the total
signal measured, the heat capacity of the microcalorimeter with
the grease was carefully subtracted from the total specic heat
measured.
Fig. 1. View of the MnWO4 crystal structure along the a axis.

3. Results and discussion


obtained materials by X-ray powder diffraction (XRD), transmission electron microscopy (TEM), Raman and IR spectroscopy as well
as low temperature magnetization and heat capacity. The main
objective of this work was to obtain information on phonon and
magnetic properties of nanocrystlline MnWO4 , defects and possible structural changes of MnWO4 with decreasing particle size. We
have employed for this purpose Raman and IR spectroscopes since
these techniques may give information on structural changes and
defects in nanomaterials [20,21].

2. Experimental
MnWO4 nanoparticles were synthesized by the hydrothermal
method using chemicals of analytical grade. In synthesis, 4 mmol
of MnSO4 H2 O (SigmaAldrich, 99%), and 4 mmol of Na2 WO4 2H2 O
(Alfa Aesar, 99%) were dissolved in deionized water. All solutions
were mixed together under stirring. The suspension formed in this
way was poured into the Teon reaction vessel, which was put
into a microwave autoclave. Next the introduced preheating at
115 C was run for half an hour. The hydrothermal reaction was
subsequently conducted at isothermal condition for one and half
an hour at different temperatures ranging from 120 to 230 C.
Additionally, two samples were prepared in a different way. First,
the reaction was conducted at 230 C for ve and half an hour
with the above mentioned preheating. Second, the reaction was
conducted at 100 C for 2 h without preheating. After this treatment the autoclave was cooled to room temperature by the water.
Resulting powders were ltered off, washed several times with
deionized water and nally dried in air at 80 C for 20 h. Apart of
the above described syntheses, where MnSO4 H2 O was a source
of manganese ions, one synthesis was performed at 100 C using
MnCl2 (SigmaAldrich, 99%).
XRD patterns were recorded at room temperature by using
XPert PRO powder diffractometer (PANalytical, The Netherlands)
working in the transmission or reection geometry, equipped with

a linear PIXcel detector and using CuK1 radiation ( = 1.54056 A)


in the 2 range from 5 to 90 with a step of 0.026 .
The particle size and morphology of MnWO4 as well as selected
area electron diffraction (SAED) studies were performed using a
200 kV Philips Super-Twin transmission electron microscope (TEM)
providing 0.24 nm resolution. Specimens for TEM were prepared
by grinding the samples in mortar and dispersing in methanol with

3.1. X-ray powder diffraction and TEM


XRD patterns of the nanocrystalline samples show presence of
diffraction peaks characteristic only for crystalline MnWO4 (see Fig.
S1 in supplementary material and ICCD Card No. 13-0434). The particle size can be estimated from Scherrers formula D = K/ cos ,
where D is the mean crystallite size along the [hkl] direction,  is
is broadening
the X-ray wavelength (in our study  = 1.54056 A),
of the diffraction line (in radians),  is the angle of diffraction, and
the Scherrer constant K is conventionally set to 1.0 [22,23]. Fig. S2
shows size of MnWO4 crystallites along a, b and c axes estimated
from broadening of the peaks corresponding to (1 0 0), (0 1 0) and
(0 0 2) planes. This gure indicates that the samples synthesized at
100140 C with use of MnSO4 H2 O contain plate-like crystallites
with the smallest dimension along the a axis, larger along the b axis
and the largest along the c axis (20, 35 and 50 nm for the sample
synthesized at 100 C). When synthesis temperature increases to
160230 C, shape of the crystallites changes to nanorods elongated
in the c direction. When MnCl2 is used instead of MnSO4 H2 O, the
average dimension of the crystallites obtained at 100 C was 24, 25
and 25 nm along a, b and c axis, respectively. This result shows that
the synthesized crystallites have smaller size along the b and c axis,
when compared to the sample synthesized at the same temperature
from MnSO4 H2 O.
TEM micrographs conrm that the MnWO4 sample synthesized
at 100 C from MnSO4 H2 O is composed of plate-like particles with
a length and width of about 3590 and 1437 nm, respectively
(see Fig. S3a). The pattern from selected area electron diffraction
(SAED) and high resolution TEM (HRTEM) images conrm crystalline nature of MnWO4 particles and show that the crystallites
prefer to grow along the [001] direction (see Fig. S3a, inset, and Fig.
S3b and S3c).
When synthesis is performed at 180 C, the MnWO4 crystallites have a shape of nanorods with a diameter of 1530 nm and
length of 4090 nm (see Figs. S3d, S3e and S3f). When the synthesis temperature increases to 230 C, size of MnWO4 nanorods
increases to 2540 nm (diameter) and 40100 nm (length) (see Fig.
S3g). The nanorods have a hexagonal-like shape in the direction
perpendicular to the c axis (see Fig. S3h).
Crystallites synthesized at 100 C from MnCl2 are smaller
(1540 nm) and have better developed faces, when compared to
those obtained from MnSO4 H2 O (see Fig. S4a and S4b). HRTEM


M. Maczka
et al. / Vibrational Spectroscopy 58 (2012) 163168

165

Fig. 2. Raman spectra of the polycrystalline (bulk) and nanocrystalline MnWO4 samples. Arrows indicate the additional Raman band assigned to defects.

images conrm highly crystalline nature of these particles (see Fig.


S4c and S4d, respectively).

3.2. Raman and IR spectra


The monoclinic C2h 4 unit cell of MnWO4 comprises 12
atoms, which have 36 zone-center degrees of freedom described
by irreducible representation of the factor group C2h as
8Ag + 8Au + 10Bg + 10Bu . Three of these modes, Au + 2Bu , are
acoustic modes. The optic modes can be subdivided into
2Ag + Au + 4Bg + 2Bu translational modes of the Mn and W atoms
and 6Ag + 6Au + 6Bg + 6Bu modes involving oxygen atoms. Among
these modes the Ag and Bg modes are Raman-active whereas the
Au and Bu modes are IR-active. This analysis shows that one expects
to observe 15 and 18 modes in IR and Raman spectra of MnWO4 ,
respectively.
Raman and IR spectra of the obtained nanocrystalline powders
as well as bulk MnWO4 are presented in Figs. 2 and 3. Recent
results of lattice dynamics calculations as well as polarized IR
and Raman scattering studies of MnWO4 , Mn0.85 Co0.15 WO4 and
Mn0.97 Fe0.03 WO4 crystals provided detailed information on origin
of different Raman and IR bands [24,25]. Here we recall a few most
important conclusions. First, the Raman-active Ag modes of bulk
MnWO4 are observed at 884, 698, 545, 398, 326, 258, 206 and
129 cm1 whereas the Bg modes at 774, 674, 511, 357, 294, 272, 178,
166, 160 and 90 cm1 . Second, the transverse optical (TO) and longitudinal optical (LO) IR-active modes of Au symmetry are observed
at 861 (897), 656 (753), 495 (536), 420 (428), 348 (373), 307 (308)
and 182 (192) cm1 (the values in parentheses correspond to LO
modes). The corresponding IR-active modes of Bu symmetry are
observed at 795 (901), 577 (739), 458 (472), 324 (382), 280 (281),
248 (273), 209 (220) and 147 (153) cm1 . Third, the Raman and IR
bands in the 770890 cm1 range can be assigned to symmetric and
asymmetric stretching modes of the short W O2 bonds. Fourth,
the bands in the 600700 cm1 range can be assigned to stretching
modes of the double oxygen bridge involving W O1 bonds. Fifths,
translational motions of W and Mn atoms contribute strongly to the
modes observed below 290 cm1 and the remaining modes in the

Fig. 3. Mid-IR (a) and far-IR (b) spectra of the polycrystalline (bulk) and nanocrystalline MnWO4 samples.

166

M. Maczka
et al. / Vibrational Spectroscopy 58 (2012) 163168

550290 cm1 range can be assigned to the WO6 bending modes


[25].
Figs. 2 and 3 show that the spectra of MnWO4 synthesized from
MnSO4 H2 O exhibit signicant changes with decreasing crystallite size. First, width of the observed bands markedly increases.
For instance, the bandwidth of the most intense Raman band at
884 cm1 is 9.3 and 17.8 cm1 for the polycrystalline sample and
the sample synthesized at 100 C, respectively. Second, Raman and
IR bands exhibit wavenumber decrease or increase. These changes
are weak for the Raman bands and the most pronounced downshifts (about 34 cm1 ) are observed for the 774, 698, 544 and
510 cm1 bands. The IR bands show much more pronounced shifts
with decreasing particle size (see Fig. 3). For instance, the 185, 513
and 614 cm1 IR bands exhibit downshift of 7, 10 and 21 cm1 ,
respectively, when going from the polycrystalline sample to the
sample synthesized at 100 C. Interestingly, the stretching band
exhibits very pronounced upshift from 818 cm1 for the polycrystalline sample to 832 cm1 for the sample synthesized at 100 C.
Third, a number of bands exhibit signicant intensity decrease with
decreasing particle size. This behavior is especially pronounced for
the 458 and 145 cm1 IR bands. Fourth, the spectra of nanocrystalline MnWO4 show presence of an additional Raman (IR) band at
about 923 cm1 (914 cm1 ) not predicted for bulk MnWO4 . Intensity of these bands increases strongly with decreasing synthesis
temperature (see Figs. 2 and 3).
Raman spectrum of MnWO4 synthesized at 100 C from MnCl2
shows signicant differences in comparison with the Raman spectrum of MnWO4 synthesized from MnSO4 H2 O. First, the bands that
were observed at 328 and 128 cm1 for the sample prepared from
MnSO4 H2 O shift to 324 and 127 cm1 . Second, many Raman bands
become signicantly narrower. For instance, width of the 884, 397,
324328, 202 and 127128 cm1 bands decreases from 17.8, 13.9,
21.5, 19.3 and 8.3 cm1 for the sample prepared from MnSO4 H2 O
to 12.7, 9.3, 16.7, 11.4 and 5.6 cm1 for the sample prepared from
MnCl2 . Third, no additional Raman band near 923 cm1 is observed
for the sample prepared from MnCl2 . Comparison of the corresponding IR spectra shows that the additional band near 914 cm1
is also present for the sample prepared from MnCl2 but its intensity
is signicantly lower.
The question that arises is what are the reasons of the observed
changes. Let us at rst discuss origin of the additional Raman
and IR bands at 923 and 914 cm1 . Intensity of these bands
increases signicantly with decreasing synthesis temperature. This
result suggests that they may correspond to unreacted precursor. According to literature data, when MnSO4 H2 O (or MnCl2 )
and Na2 WO4 2H2 O water solutions are mixed together, hydrated
manganese tungstate is formed with water molecules intercalated
between the (0 1 0) planes of MnWO4 [26]. Our Raman spectra indicate that the additional band cannot be assigned to the hydrated
manganese tungstate impurity phase since it is observed at much
higher wavenumber (923 cm1 ) than the most intense Raman band
of hydrated MnWO4 (906 cm1 [26]). Another possibility is that
as a result of the reaction, NaBi(WO4 )2 scheelite impurity phase
was formed. However, according to the literature data the most
intense Raman band of this scheelite phase should be observed near
909 cm1 [27], not 923 cm1 . Moreover, presence of the scheelite
impurity would not explain presence of the 914 cm1 IR band. The
third possibility is that that these bands correspond to amorphous
or nanocrystalline impurity containing tungsten oxide. This type of
impurity can probably be excluded since amorphous and nanocrystalline WO3 should show a strong Raman band near 950 cm1
[28,29]. We suppose, therefore, that these additional bands appear
due to presence of surface defects, not an impurity phase. Former studies of tungsten oxide-based compounds showed that such
materials often have Raman and IR bands in the 900960 cm1

range due to presence of terminal W O groups at the surface


[2830]. We suppose, therefore, that the additional Raman and
IR bands appear due to presence of terminal W O bonds at the
surface of MnWO4 nanocrystallites. It is worth noting that Tong
et al. also attributed the additional IR and Raman bands to surface
defects and suggested that a surface disordered layer of 1.8 nm
thickness is formed [19]. Interestingly, our present results show
that the additional Raman band is not observed for the MnWO4
crystallites synthesized from MnCl2 , even though the size of the
crystallites is smaller than that observed for the sample synthesized from MnSO4 H2 O. However, TEM images show that the later
crystallites have not well developed faces and are strongly defected.
Therefore presence of the additional Raman band for this sample
supports our conclusion that it is related to surface defects. It is also
worth noting that although Raman spectrum of the sample synthesized from MnCl2 does not show presence of the additional band at
923 cm1 , the IR spectrum of this sample shows an additional band
at 914 cm1 . This result indicates that there is no clear correlation
between the Raman band at 923 cm1 and the IR band at 914 cm1 .
We suppose, therefore, that these bands arise due to different type
of defects.
As mentioned above, Raman bands show some changes with
decreasing particle size. It is well known that phonon properties
of materials are signicantly modied when size of crystallites is
of order of nanometers due to a few factors such as creation of
defects, distribution of crystallite size, changes of the lattice parameters (strain), phonon connement effect and variation in phonon
relaxation time [21,31]. These factors give rise to broadening and
shifts of bands. Moreover, the bands usually become asymmetric
due to phonon connement effect. The observed broadening of
Raman bands of our MnWO4 samples can be understood in terms of
phonon connement effect and creation of defects in these nanosized MnWO4 samples. Larger broadening of many bands for the
sample synthesized at 100 C from MnSO4 H2 O, when compared
to the sample synthesized from MnCl2 , indicates that the crystallites in the former sample are more defected. As far as Raman
shifts are concerned, phonon connement may lead to both negative and positive shift of a Raman band, depending on shape of
the phonon dispersion curves [20,21,31]. Although there are no
reports on phonon dispersion for MnWO4 , such data are available for isostructural CdWO4 [32]. These data show that nearly
all Raman active modes exhibit weak dispersion near the Brillouin
zone center and therefore weak and negative shift of Raman bands
with decreasing particle size is expected for majority of modes. The
observed by us weak shifts of the Raman bands can be, therefore,
attributed mainly to phonon connement effect.
In contrast to the Raman spectra, the IR spectra show much
more pronounced changes with decreasing particle size. First,
large shifts towards lower wavenumbers are observed for some
bands. For instance, the 614 cm1 band for the polycrystalline
sample shifts to 593 cm1 for the sample synthesized at 100 C.
Second, the 818 cm1 band of the polycrystalline sample shifts
towards higher wavenumbers. For instance, this band is observed
at 832 cm1 for the sample synthesized at 100 C. Similarly as discussed above for the Raman-active modes, phonon connement
effect is expected to lead to weak negative shifts also for the IRactive modes. It is, therefore, unlikely that this effect is signicant
enough to cause such large IR bands shifts. The observed shifts cannot also be attributed to strain, defects or some structural changes
since in such case similar changes should be observed also in the
Raman spectra. We suppose, therefore, that such large shifts of IR
bands can be attributed, at least partially, to polar character of
the observed modes because shift of a corresponding IR band is
clearly correlated with magnitude of LOTO splitting, i.e. the larger
the splitting the more pronounced changes in bands position are
observed. It is worth noting that it has also been reported in many


M. Maczka
et al. / Vibrational Spectroscopy 58 (2012) 163168

publications that shape and maximum intensity of an IR transmission band can be very dependent upon the size and shape of the
specimen due to the long-range Coulomb forces [3336]. As a result,
in ionic crystals the maximum absorption of an IR band may be
located between TO and LO [33,37]. For cubic crystals, maximum
absorption can be calculated from the known LOTO splitting for
different shapes of nanocrystals [36]. For instance, long rodlike particle should absorb near TO for polarization parallel to the rod axis
and between TO and LO for polarization perpendicular to the rod
axis [36]. In case of monoclinic MnWO4 , behavior of the IR bands
is expected to be more complicated. Furthermore, literature data
showed that if particle aggregations are formed, the dipoledipole
interactions between particles might give rise to an additional
broadening and shift of an IR band [38]. Closer inspection of our IR
spectra show that for the sample synthesized at 230 C all modes
absorb slightly above TO, except of four modes that exhibit large
shifts towards LO value, i.e. from 795, 656, 495 and 348 cm1 (TO
values) to 828, 700, 512 and 360 cm1 , respectively. When synthesis temperature decreases, nearly all Au and Bu modes shift closer
to TO values. The only exception is the Bu stretching mode, which
shifts slightly closer to LO value, i.e. from 828 cm1 for the sample synthesized at 230 C to 832 cm1 for the sample synthesized
at 100 C. TEM images show that decrease of the synthesis temperature leads not only to smaller size of the crystallites but also
to some change in their shape. Change of shape should contribute
to the observed shifts of IR bands but this effect should be similar
for the modes polarized in the same direction. Our former polarized IR studies of Mn0.85 Co0.15 WO4 and Mn0.97 Fe0.03 WO4 crystals
showed that the mentioned above Bu mode, which exhibits positive shift with decreasing particle size, is polarized along the a axis
[25]. However, the remaining Bu modes polarized along the a axis
at 795, 458, 359, 209 and 141 cm1 (TO values) exhibit negative
shifts. Therefore, change of the MnWO4 particle shape alone cannot explain the observed behavior of the IR bands. We conclude,
therefore, that the observed shifts of IR bands are related to many
factors such as change of particle shape, concentration of defects,
degree of particles aggregation and phonon connement effect.
3.3. Magnetic susceptibility and heat capacity
Fig. 4a presents temperature dependence of the magnetization
M(T) of the MnWO4 sample synthesized at 180 C. This sample
was chosen for magnetic studies since it is composed of wellcrystallized particles and shows relatively small distribution of
particle sizes. The distinct maximum at 11.0 K is an evidence of a
magnetic ordering that occurs in the compound. Rise of the magnetization below 6 K can be most likely attributed to presence of some
paramagnetic impurities. The antiferromagnetic ordering manifests itself also in the temperature variation of the specic heat
Cp (T) as a distinct anomaly located at 11.4 K (see Fig. 4b). No anomalies are observed near the expected phase transition at 6.58 K. The
absolute values of the magnetization are close to those reported
for bulk MnWO4 [10]. However, only one broad anomaly appears,
instead of two expected transitions at TN and T2 . This result can
be attributed to small size of the crystallites and size distribution
of nanocrystallites. Our results also show that the magnetic phase
transition appears at slightly lower temperature for the nanocrystalline sample (11.011.4 K) when compared to the bulk material
(T2 12.5 K and TN 13.5 K [10]). It is worth to add that former magnetic susceptibility studies of MnWO4 nanowires with an average
diameter of 200 nm could not detect any anomalies, which might
correspond to the three thermodynamic transitions found for bulk
MnWO4 [18]. Instead, a broad anomaly was observed near 7.5 K,
which was attributed to structural disorder in the sample, such
as the presence of small-crystalline particulate regions [18]. Our
MnWO4 nanoparticles are highly crystallized and in this case the

167

Fig. 4. Low-temperature magnetization M (a) and specic heat Cp (b) of the


nanocrystalline MnWO4 sample prepared at 180 C as a function of temperature
T, measured in external magnetic eld B.

magnetic ordering occurs at only slightly lower temperature than


for bulk MnWO4 .
4. Conclusions
MnWO4 nanoparticles with 20100 nm sizes were synthesized by a mild hydrothermal crystallization process. Particles
synthesized at 160230 C have form of nanorods along the c crystallographic axis. Our results show that crystallites synthesized
from MnCl2 have smaller concentration of defects than those synthesized from MnSO4 H2 O. These defects contribute to broadening
of Raman bands and give rise to an additional Raman (IR) band
at 923 (914) cm1 , which can be attributed to presence of terminal W O bonds at the surface of MnWO4 nanocrystallites. Raman
bands revealed weak wavenumber shifts with decreasing particle size due to phonon connement and defects. In contrast to
this behavior, IR bands show much more pronounced wavenumber
shifts, which can be attributed to polar character of the observed
modes.
Specic heat and magnetization studies revealed that magnetic
order appears at low temperatures also for the MnWO4 nanocrystals. However, the transition temperature, at which magnetic order
occurs, is lowered by about 2 K in comparison with bulk MnWO4 .
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.vibspec.2011.12.010.
References
[1]
[2]
[3]
[4]

D. Khomskii, Physics 2 (2009) 20.


M. Fiebig, J. Phys. D: Appl. Phys. 38 (2005) R123.
Y. Tokura, J. Magn. Magn. Mater. 310 (2007) 1145.
T. Kimura, G. Lawes, T. Goto, Y. Tokura, A.P. Ramirez, Phys. Rev. B 71 (2005)
224425.

168

M. Maczka
et al. / Vibrational Spectroscopy 58 (2012) 163168

[5] G. Lawes, A.B. Harris, T. Kimura, N. Rogado, R.J. Cava, A. Aharony, O. EntinWohlman, T. Yildrim, M. Kenzelmann, C. Broholm, A.P. Ramirez, Phys. Rev. Lett.
95 (2005) 087205.
[6] T. Kimura, J.C. Lashley, A.P. Ramirez, Phys. Rev. B 73 (2006) 220401.
[7] M. Kanzelmann, G. Lawes, A.B. Harris, G. Gasparovic, C. Broholm, A.P. Ramirez,
G.A. Jorge, M. Jaime, S. Park, Q. Huang, A.Y. Shapiro, L.A. Demianets, Phys. Rev.
Lett. 98 (2007) 267205.
[8] O. Heyer, N. Hollmann, S. Jodlauk, L. Bohaty, P. Becker, J.A. Mydosh, T. Lorenz,
D. Khomskii, J. Phys. Condens. Matter 18 (2006) L471.
[9] J. Macavei, H.Z. Schulz, Kristallogr. 207 (1993) 193.
[10] A.H. Arkenbout, T.T.M. Palstra, T. Siegrist, T. Kimura, Phys. Rev. B 74 (2006)
184431.
[11] K. Taniguchi, N. Abe, T. Takenobu, Y. Iwasa, T. Arima, Phys. Rev. Lett. 97 (2006)
097203.
[12] W. Qu, J.U. Meyer, Meas. Sci. Technol. 8 (1997) 593.
[13] P. Becker, L. Bohaty, H.J. Eichler, H. Rhee, A.A. Kaminskii, Laser Phys. Lett. 4
(2007) 884.
[14] L. Zhang, C. Lu, Y. Wang, Y. Cheng, Mater. Chem. Phys. 103 (2007) 433.
[15] S. Thongtem, S. Wannapop, A. Phuruangrat, T. Thongtem, Mater. Lett. 63 (2009)
833.
[16] S.J. Chen, X.T. Chen, Z. Xue, J.H. Zhou, J. Li, J.M. Hong, X.Z. You, J. Mater. Chem.
13 (2003) 1132.
[17] Y. Xing, S. Song, J. Feng, Y. Lei, M. Li, H. Zhang, Solid State Sci. 10 (2008) 1299.
[18] H. Zhou, Y. Yiu, M.C. Aronson, S.S. Wong, J. Solid State Chem. 181 (2008)
1539.
[19] W. Tong, L. Li, W. Hu, T. Yan, X. Guan, G. Li, J. Phys. Chem. C 114 (2010)
15298.
[20] P. Colomban, Spectroscopy Europe 15 (2003) June 8.

[21] J.E. Spanier, R.D. Robinson, F. Zhang, S.W. Chan, I.P. Herman, Phys. Rev. B 64
(2001) 245407.
[22] S. Chattopadhyay, P. Ayyub, V.R. Palkar, M. Multani, Phys. Rev. B 52 (1995)
13177.
[23] J.M. Amig, F.J. Serrano, M.A. Kojdecki, J. Bastida, V. Esteve, M.M. Revents, F.
Marti, J. Eur. Ceram. Soc. 25 (1995) 1479.
[24] M.N. Iliev, M.M. Gospodinov, A.P. Litvinchuk, Phys. Rev. B 80 (2009) 212302.
[25] M. Maczka, M. Ptak, K. Hermanowicz, A. Majchrowski, A. Pikul, J. Hanuza, Phys.
Rev. B 83 (2011) 174439.
[26] B. Ingham, C.V. Chong, J.L. Tallon, Soft condensed matter: new research, in:
K.I. Dillon (Ed.), OrganicInorganic Layered Hybrid Materials, Nova Science
Publishers, 2007 (Chapter 6).
[27] J. Hanuza, M. Maczka, J.H. Van der Maas, J. Mol. Struct. 348 (1995) 349.
[28] A. Kuzmin, M. Kalendarev, A. Kursitis, J. Purans, J. Non-Cryst. Solids 353 (2007)
1840.
[29] Y. Djaoued, S. Priya, S. Balaji, J. Non-Cryst. Solids 354 (2008) 673.
[30] M.F. Daniel, B. Desbat, J.C. Lassegues, B. Gerand, M. Figlarz, J. Solid State Chem.
67 (1987) 235.
[31] A.K. Arora, M. Rajalakshmi, T.R. Ravindran, V. Sivasubramanian, J. Raman Spectrosc. 38 (2007) 604.
[32] M. Saito, J. Suda, T. Sato, J. Spectrosc. Soc. Jpn. 51 (2002) 65.
[33] I.I. Shaganov, T.S. Perova, V.A. Melnikov, S.A. Dyakov, K. Berwick, J. Phys. Chem.
C 114 (2010) 16071.
[34] Q. Zhang, Z. Zhang, Appl. Phys. A 91 (2008) 631.
[35] G. Gouadec, Ph. Colomban, Prog. Cryst. Growth Charact. Mater. 53 (2007) 1.
[36] J.T. Luxon, D.J. Montgomery, R. Summitt, Phys. Rev. 188 (1969) 1345.
[37] K. Genzel, T.P. Martin, Phys. Status Solidi B 51 (1972) 91.
[38] M. Abdulkhadar, B. Thomas, Nanostruct. Mater. 5 (1995) 289.

You might also like