Você está na página 1de 10

Journal of

Materials Chemistry C
View Article Online
PAPER View Journal | View Issue

Temperature- and pressure-induced phase


Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

transitions in the niccolite-type formate


Cite this: J. Mater. Chem. C, 2016,
4, 3185 framework of [H3N(CH3)4NH3][Mn2(HCOO)6]†
Mirosław Ma˛czka,*a Anna Ga˛gor,a Nathalia Leal Marinho Costa,b
Waldeci Paraguassu,b Adam Sieradzkic and Adam Pikula

We report the synthesis, crystal structure, thermal, pyroelectric, Raman, infrared and magnetic properties
of [NH3(CH2)4NH3][Mn2(HCOO)6] niccolite. Our results show that this compound crystallizes in a trigonal
%
structure (space group P31c) with dynamically disordered [NH3(CH2)4NH3]2+ cations. It undergoes a
phase transition near Tc = 350 K. The low-temperature structure is polar (space group Cc) and
pyroelectric measurements confirm that it exhibits ferroelectric properties. Detailed analysis of the
structural changes shows that both the spatial arrangement of the [NH3(CH2)4NH3]2+ dipole moments
and distortion of the manganese formate framework contribute to the spontaneous polarization within
the (a, c) plane. Based on Raman and IR data, assignment of the observed modes to the respective
vibrations of atoms is also proposed. Dynamic disorder of organic cations in the high-temperature
phase manifests in the vibrational spectra through very large width of bands corresponding to vibrations
Received 10th February 2016, of the NH3 groups. Ordering of these cations is clearly observed in the spectra through a pronounced
Accepted 4th March 2016 decrease in their bandwidths below the phase transition temperatures. Low-temperature magnetic
DOI: 10.1039/c6tc00611f studies show that this compound is a weak ferromagnet below 9.0 K. We also report high-pressure
Raman scattering studies of this compound, which reveal the presence of two pressure-induced phase
www.rsc.org/MaterialsC transitions between 0.5 and 0.9 GPa and between 1.3 and 1.9 GPa.

Introduction compounds (M = Mg, Zn, Mn, Ni, Co, Fe) were shown to exhibit
ferroelectric order below 254–160 K.1–5 Furthermore, the analogues
Divalent metal formates with general formula [amineH+][M(HCOO)3] with M = Mn, Ni, Co and Fe order magnetically within 30–8 K.3,4
have been extensively studied in recent years due to their interesting Formates with larger cations such as dimethylammonium (DMA+),
structural, magnetic and dielectric properties. For small cations, ethylammonium, formamidinium as well as [NH4][Cd(HCOO)3]
such as NH4+ and HONH3+, these metal organic frameworks crystallize in the perovskite structure,6–9 whereas formates with
(MOFs) crystallize in the chiral structure and [NH4][M(HCOO)3] hydrazinium cations (H2NNH3+) may crystallize in both the
chiral and perovskite structural types.10 Among the family of
perovskite formates, those with DMA+ cations have been most
a
Institute of Low Temperature and Structure Research, Polish Academy of Sciences,
extensively studied due to their multiferroic properties.7,8,11,12 It
Box 1410, 50-950 Wrocław 2, Poland. E-mail: m.maczka@int.pan.wroc.pl;
Fax: +48-713441029; Tel: +48-713954161
is also worth adding that recent studies showed that the properties
b
Faculdade de Fı́sica, Universidade Federal do Pará, 66075-110, Belém, PA, Brazil of perovskite formates may be modified by A-site or B-site doping,
c
Department of Experimental Physics, Wrocław University of Technology, Wybrzeże leading to the diffuse character of phase transitions.13,14
Wyspiańskiego 27, 50-370, Wrocław, Poland A metal formate framework templated by large protonated
† Electronic supplementary information (ESI) available: Tables S1–S6: crystal data
di-, tri- and tetra-amines crystallizes in more complicated
and selected geometrical parameters for the studied compound at different
temperatures, the results of the mode analysis performed in Amplimodes, Raman
topologies.15 We will focus here only on compounds with
and IR wavenumbers together with the proposed assignments and wavenumber protonated di-amines of general formula [amineH22+][M2(HCOO)6].
intercepts at zero pressure (o0) and pressure coefficients (a = do/dP) for the three To the authors’ knowledge, compounds of this composition
phases of bnMn observed in the high pressure experiment. Fig. S1–S8: Le Bail fit were reported only for N,N 0 -dimethylethylenediamine and 1,4-
to the powder XRD pattern, DSC traces, IR and Raman spectra at different
diaminobutane (NH2(CH2)4NH2) and they were shown to crystallize
temperatures, wavenumbers vs. pressure plots of the Raman modes observed in
bnMn crystals for the compression experiment and Raman spectra recorded
in niccolite-type structure.16–18 Only two compounds with 1,4-
during the decompression experiment. CCDC 1446778–1446782. For ESI and diaminobutane are known with M = Mg and Co. X-ray diffraction
crystallographic data in CIF or other electronic format see DOI: 10.1039/c6tc00611f data revealed that [NH3(CH2)4NH3][Mg2(HCOO)6] (bnMg) undergoes

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. C, 2016, 4, 3185--3194 | 3185
View Article Online

Paper Journal of Materials Chemistry C

structural phase transition at 398–412 K from the centrosymmetric High-pressure studies have not yet been reported for any niccolite-
trigonal structure (space group P3% 1c) to the polar monoclinic type formate framework and we will show that in contrast to the
structure (space group Cc).17 Recent studies showed that the reported perovskite-type frameworks, for which pressure enhances
room-temperature crystal structure of bnCo can also be described the ferroelectric polarization,9,25 application of pressure leads to
by the centrosymmetric space group P3% 1c and this compound the suppression of the ferroelectric properties of bnMn.
orders magnetically below 9.9 K.18 However, the presence of
electric order was not reported for this compound and any other
divalent metal formate crystallizing in the niccolite structure. Experimental
The studies of polymorphism in MOFs are very important for Synthesis
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

understanding the structure–property relationship.19 Therefore,


MnCl2 (99%, Sigma-Aldrich), methanol (99.8%, Sigma-Aldrich),
MOFs were often studied as a function of temperature. Pressure
1,4-diaminobutane (99%, Sigma-Aldrich) and formic acid (98%,
is another thermodynamics variable but it was much less
Fluka) were commercially available and used without further
explored in the studies of MOFs although high-pressure studies
purification. To obtain single crystals of bnMn, 16 mL of
can be useful in many ways relevant to potential technological
methanol solution containing 5 mmol of 1,4-diaminobutane
applications.20 In this respect, high-pressure studies of chiral
and 20 mmol of formic acid was placed at the bottom of a glass
formates [NH4][Zn(HCOO)3] and [ND4][Zn(DCOO)3] revealed
tube (9 mm inner diameter). To this solution, 16 mL of
that these compounds exhibit negative linear compressibility
methanol solution containing 2 mmol of MnCl2 was gently
and a pressure-induced phase transition into a monoclinic
added. The tube was sealed and kept undisturbed. Light pink
phase at about 1.1–1.35 GPa.21 Pressure-induced phase transitions
crystals were harvested after 3 days. Elemental analysis (C, H,
were also found for azetidinium manganese (AzMn), guanidinium
N) was performed on an Elementar Vario EL CHNS analyzer.
manganese (GuaMn), azetidinium zinc (AzZn), dimethylammonium
Anal. calcd for bnMn (%): C, 25.55; H, 4.29; N, 5.96; found (%):
magnesium (DMMg), dimethylammonium cadmium (DMCd) and
C, 25.47; H, 4.34; N, 5.91. The results of the Le Bail refinement
[NH4][Cd(HCOO)3] formates crystallizing in the perovskite-type
confirm the phase purity of the obtained sample (Fig. S1, ESI†).
architecture.9,22–24 Interestingly, the performed studies suggested
that the polar character of [NH4][Cd(HCOO)3] can be greatly DSC
enhanced by application of pressure.9 Very recent theoretical
studies of guanidinium chromium formate and ethylammonium Heat capacity was measured using a Mettler Toledo DSC-1
manganese formate also revealed that the ferroelectric polarization calorimeter with a high resolution of 0.4 mW. Nitrogen was
of these compounds should enhance by application of compressive used as a purging gas. The weight of the bnMn sample was
strain.25 Among metal formates templated by protonated di-, tri- or 13.3 mg. The heating and cooling rates were 5 K min1. The
tetra-amines, high-pressure data were reported only for erbium excess heat capacity associated with the phase transition in
formate templated by N,N,N0 ,N0 -tetramethylethylenediammonium bnMn was evaluated by subtraction from the data of the baseline
cations.26 representing variation in the absence of the phase transitions.
The most interesting feature of some metal formate frame-
X-ray powder diffraction
works is the coexistence of magnetic and electric order in a
single phase. Single-phase multiferroics have become a hot The powder XRD pattern was obtained for all samples on an
topic in recent years since such materials are promising for X’Pert PRO X-ray diffraction system equipped with a PIXcel
applications in electromagnetic sensors, telecommunication ultrafast line detector, focusing mirror, and Soller slits for
systems, data storage etc.27 However, such materials are very CuKa1 radiation (l = 1.54056 Å).
rare and in the family of divalent metal formate frameworks the
coexistence of magnetic and electric order was reported only for Single crystal X-ray diffraction
the above mentioned chiral and perovskite structures templated XRD data were collected using Xcalibur-Atlas (for low temperatures)
by monoammonium cations.3,7,11 We decided, therefore, to search and Xcalibur-Sapphire1 (for room temperature and high tem-
for novel materials exhibiting the coexistence of magnetic and peratures) diffractometers, both operating in k-geometry,
electric order in the weakly studied niccolite-type frameworks equipped with a two-dimensional CCD detector and a Mo Ka
templated by diammonium cations. [NH3(CH2)4NH3][Mn2(HCOO)6] radiation (0.71073 Å) source. Data were measured in a o-scan
(bnMn) was chosen as a promising candidate since very recent mode with Do = 1.01. The CrysAlis PRO was used for data
reports on Mg analogue indicated a polar structure at room collection and processing [CrysAlis PRO, Agilent Technologies,
temperature.17 We report temperature-dependent X-ray diffraction, Version 1.171.37.35h.]. The structures were solved by direct
Raman, IR and dielectric studies of bnMn in order to understand methods and refined using full-matrix least-squares methods
the origin of electric ordering, i.e., whether it is related to a peculiar using SHELXL2014/7.28 Empirical absorption correction using
ordering of organic cations, as reported for chiral- and perovskite- spherical harmonics, implemented in the SCALE3 ABSPACK
type formates, or it arises due to distortion of the Mn2[COOH]62 scaling algorithm, was applied on all data. The restraints on C–C
framework. Our aim was also to obtain information on the distances (DFIX) were used to fix disordered [NH3(CH2)4NH3]2+
stability of this compound under the application of hydrostatic (bnH22+) cations to be chemically reasonable in the trigonal phase
pressure and the effect of pressure on its ferroelectric properties. at 360 K and in the monoclinic phase at 345 K. Hydrogen atoms

3186 | J. Mater. Chem. C, 2016, 4, 3185--3194 This journal is © The Royal Society of Chemistry 2016
View Article Online

Journal of Materials Chemistry C Paper

were introduced in calculated positions and refined as riding homemade furnace. The spectral resolution of Raman and IR
atoms. Thermal parameters of hydrogen atoms were set to be spectra was 2 cm1.
equal to 1.2 times the thermal parameters of the corresponding
parent atoms. The high temperature was sustained with a hot- High-pressure Raman scattering studies
air flow attachment (Kuma Diffraction, covering the temperature The high-pressure Raman spectra were recorded in back-scattering
range 300–770 K) while an open flow nitrogen cryosystem (Oxford geometry using a microscope attached to a triple-grating spectro-
Cryosystem, covering the temperature range 90–320 K) was used meter Jobin Yvon T64000. The 514.5 nm line of a solid-state ion
for experiments at low temperatures. Table S1, in the ESI,† laser was used as excitation and the spectral resolution was
presents the results of the data collection and refinement together 2 cm1. In order to reach high pressures, a diamond anvil cell
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

with the crystal description. Tables S2 and S3 (ESI†) show the m-scope DAC HT(S) from Almax easyLab with a diamond of
selected interatomic distances and hydrogen bond parameters. 0.4 mm of culets was used. The sample was loaded into a
100 mm hole drilled in a stainless steel gasket with a thickness
Dielectric and pyroelectric measurements
of 200 mm using an electric discharge machine from Almax
Ambient pressure dielectric measurements of the examined easyLab. The Nujol served as the pressure transmitting media.
sample were carried out using a Novocontrol analyzer. Since the Pressures were measured based on the shifts of the ruby R1 and
obtained single crystals were not big enough to perform single R2 fluorescence lines.
crystal dielectric measurements, pellets made of well-dried
samples were measured instead. The pellets were placed between
two copper, flat electrodes (diameter 6 mm) of the capacitor with a Results and discussion
gap of 0.3 mm. The small signal of amplitude 1 V was applied
DSC
across the sample. The temperature was controlled by the Novo-
Control Quattro system, with use of a nitrogen gas cryostat. The The DSC measurements show the presence of one heat anomaly
measurements were taken every 1 deg over the temperature at 355.2 K upon heating and 349.7 K upon cooling (Fig. S2, ESI†).
range from 300 to 400 K. Temperature stability of the samples The relatively large thermal hysteresis (5.5 K) and symmetric shape
was better than 0.1 K. of the thermal anomaly (Fig. S2 (ESI†) and Fig. 1) suggest a first-
For the pyroelectric measurements, the pellet sample was order character of the phase transition.
poled in an electric field of 1 kV cm1 from 400 to 310 K. After The associated change in enthalpy DH and entropy DS was
removing the poling electric field and releasing space charges estimated to be B2.6 kJ mol1 and B7.4 J mol1 K1, respectively.
for at least 20 minutes, the pyroelectric current was recorded From the Boltzmann equation, DS = R ln(N), where R is the gas
with heating and cooling at a constant rate of 2 K min1. constant and N is the number of sites for the disordered system,
which can be estimated to be 2.4. X-ray diffraction data reported in
Magnetic measurements the next paragraph show the trigonal disorder of the bnH22+
The magnetic properties of a large number of freely oriented cations in the high-temperature phase. The value of N obtained
single crystals of bnMn (about 70 mg in total) were studied by us (2.4) is, therefore, lower than the expected N value (3) for a
using a commercial superconducting quantum interference complete ordering of the bnH22+ cations. It is worth noting that the
device (SQUID) magnetometer equipped with an AC suscepto- former study of the isostructural bnMg compound estimated N to
meter. Magnetization was measured in the temperature range be 3.9, i.e. significantly larger than expected for the trigonal
2–40 K and in the external magnetic field up to 50 kOe. The AC disorder of bnH22+ cations.17 Furthermore, the phase transition
magnetic susceptibility was measured in the zero applied
magnetic field using an alternate probing magnetic field with
an amplitude of 3.73 Oe and frequency varying from 1 Hz to
1.5 kHz. The background coming from a weakly diamagnetic
sample holder (not shown here) was found to be negligible in
comparison to the total signal measured. Therefore its subtraction
was omitted. Also no demagnetization corrections were made to
the data reported here.

Temperature-dependent Raman and IR studies


Temperature-dependent Raman spectra were measured using a
Renishaw InVia Raman spectrometer equipped with a confocal
DM 2500 Leica optical microscope, a thermoelectrically cooled
CCD as a detector, an argon laser operating at 488 nm and
a Linkam cryostat cell. Temperature-dependent IR spectra in
the 5–300 K range were measured using a Biorad 575C FT-IR
spectrometer and a helium-flow Oxford cryostat. Additional Fig. 1 The heat capacity of bnMn measured in a cooling mode. The insets
measurements in the 300–450 K range were done using a show the change in Cp and S related to the phase transition.

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. C, 2016, 4, 3185--3194 | 3187
View Article Online

Paper Journal of Materials Chemistry C

into the disordered phase occurs at higher temperature for phase transition leads to lowering of the crystal symmetry to
bnMg (B412 K in heating mode) compared to bnMn (B355 K the monoclinic with the polar Cc space group. In the new
in heating mode). This behavior shows that similarly as in phase, the site symmetry of all atoms is reduced to C1. The
the case of perovskite-type divalent formates templated by mechanism of the transformation is complex and involves both
protonated amines,29 also for the niccolite-type formate Tc is ordering of the bnH22+ ions and distortion of the Mn2(COOH)62
unusually high for the Mg analogue. framework. The metrics of the primitive monoclinic structure
corresponds to the trigonal unit cell. Due to the C centering the
Single crystal X-ray diffraction volume of the monoclinic cell is doubled compared to the trigonal
BnMn is isomorphic to the recently reported bnMg analogue.17 cell. Fig. S3 (ESI†) presents the temperature evolution of the
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

It crystallizes in a niccolite-type structure, in the P3% 1c space monoclinic lattice parameters during heating. The most affected
group, similar to several newly synthesized heterometallic and by the transition are the a and b directions with Da/a and Db/b B
mixed-valence metal formate framework materials.30 In this 2% whereas the cmono = ctrigonal change only insignificantly.
trigonal structure the divalent manganese ions occupy two non- Nevertheless, close to the transition temperature the monoclinic
equivalent positions of 3.2 (D3) and 3% (C3) symmetries, both are lattice exposes noteworthy changes in the monoclinic structure
octahedrally coordinated by the formate ligands. The Mn1/ in the c direction in quite a wide temperature range (B15 K).
Mn2–O distances are equal to 2.174(3)/2.173(2) Å at 360 K. The second important anomaly refers to the volume, which
Significantly shorter distances of 2.0807(10)/2.0845(11) Å at decreases after the transition to the hexagonal cell manifesting
430 K were reported for bnMg.17 This behavior reflects the small negative thermal expansion of the lattice. Small negative
smaller ionic radius of Mg2+ compared to Mn2+ and thus a thermal expansion was also noticed for isomorphic bnMg.17
significantly smaller unit cell volume of bnMg (853.41(5) Å3 at The observed structure anomalies are most likely associated
410 K)17 compared to bnMn (894.44(6) Å3 at 360 K). The metal with the ordering process and conformational changes of the
centers are connected through formate linkers in the anti–anti bnH22+ ions. The determinant factors of the conformational
mode configuration. The template counterions accommodate preferences of this di-protonated linear polyamine are the
the elongated crystal voids. The bnH22+ ions are dynamically electrostatic forces and steric effects.31 Minimization of the
disordered within three symmetrically equivalent positions repulsive N–H  H–N steric interactions and formation of intra-
around the three-fold axis, see Fig. 2. At 345 K the structural molecular hydrogen bonds with temperature decrease lead to

Fig. 2 (a) Disorder and conformational changes of the bnH22+ ions in the trigonal (360 K) and monoclinic phases (345–90 K). The crosses mark the
positions of additional electron maxima related to the disordered mirror position. (b and c) Arrangement of the template ions in the crystal cavities in the
trigonal and monoclinic phase, respectively. Dashed lines stand for hydrogen bonds.

3188 | J. Mater. Chem. C, 2016, 4, 3185--3194 This journal is © The Royal Society of Chemistry 2016
View Article Online

Journal of Materials Chemistry C Paper

the lattice distortion and conformational changes of template setting are populated in the 0.7/0.3 ratio. The temperature
ions in bnMn. In the high-temperature phase the counterions decrease reduces the movements. At 298 K this particular
have the eclipsed form with C–C–C–N dihedral angles of 1271 disorder manifests as an additional maximum at the electron
and 1271. The temperature lowering induces rearrangement density map with r B 1.7 e Å3. It is also present at 160 K where
of the ions. In the monoclinic phase the dihedral angles change r B 0.74 e Å3 and finally, disappears at 90 K. Fig. 2(a)
to 741 and 721 at 345 K and 681 and 671 at 90 K indicating illustrates the change of the orientational states for bnH22+
gauche conformation which is, interestingly, less energetically with the temperature lowering. It is worth adding that partial
favored according to the ab initio calculations.31 Simultaneous disorder was also reported for bnMg and the mirror setting
skeletal rearrangements and thermally activated positional population decreased for this compound from 0.75/0.25 at
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

disorder were observed also in other 1,4-diaminobutane hybrids, 390 K to 0.85/0.15 at 290 K and 0.9/0.1 at 240 K.17
e.g. in [NH3(CH2)4NH3][MnCl4].32 Below Tc, the ordering of the bnH2+ ions breaks the symmetry
In the trigonal phase the bnH22+ ion may perform free and induces the spontaneous polarization in the (a, c) plane.
rotations along the c-axis changing the placement among The symmetry mode analysis performed in AMPLIMODES33
3 possible sites. In the monoclinic phase this dynamical disorder shows, however, that non-trivial contribution to the induction
is significantly reduced. The cation is stabilized through the of the spontaneous polarization has also a polar distortion of
N–H  O hydrogen bonds of medium strength that involve all the Mn2[COOH]62 framework. Thus, the ferroelectric phase
amine hydrogen atoms and formate oxygens. The donor to transition has both order–disorder and displacive contributions.
acceptor distances range from 2.823(8) Å to 3.138(8) Å, the The details concerning the atomic displacements between P3% 1c
donor to hydrogen to acceptor angles range from 1571 to 1771 and Cc phases are given in ESI,† in Table S4.
at 345 K and do not change considerably with temperature
decrease except that at room temperature and below additional Dielectric and pyroelectric measurements
bifurcated contact appears (Table S3, ESI†). The corresponding Fig. 3 shows the temperature dependence of the complex
hydrogen bond parameters for bnMg just below Tc (at 390 K) dielectric permittivity for bnMn. The real part of the dielectric
were 2.847(2)–3.220(2) Å and 158.0–174.81.17 Slightly larger permittivity shows a step-like change at about 350 K (Fig. 3(a)).
donor to acceptor distances for bnMg suggest slightly weaker Nearly discontinuous-like anomaly in the dielectric permittivity
hydrogen bonds for this compound. X-ray diffraction data also suggests a first-order type phase transition. Furthermore, the
show that the hydrogen bond parameters of bnMg exhibit more real part of the dielectric permittivity is almost frequency
significant changes upon cooling compared to bnMn. As a result, independent at frequencies higher than 1 KHz and some
the donor to acceptor distances at low temperatures become deviation can be noticed only on the low temperature side.
slightly shorter for bnMg (2.815(13)–3.1147(12) Å at 100 K)17 The loss tangent also shows a distinct anomaly at the phase
than bnMn (2.817(3)–3.160(4) Å at 90 K). transition temperature (Fig. 3(b)). It can be noticed that at low
Despite the stabilization of –CH2–NH3+ terminal groups the frequencies (below 2 kHz) the real part of the dielectric permittivity
central ethylene is prone to temperature-induced rotations and the loss tangent exhibit a significant increase on the high
within the additional mirror position, see Fig. 2(a). At 345 K, temperature side. Such behavior may indicate some conductivity
very close to the phase transition, the main and the mirror process, which is activated at higher temperatures and which

Fig. 3 Temperature dependence of complex dielectric permittivity: (a) real part and (b) loss tangent. The measurements were taken with increasing
frequency.

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. C, 2016, 4, 3185--3194 | 3189
View Article Online

Paper Journal of Materials Chemistry C


Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

Fig. 4 The pyroelectric current as a function of temperature for poly-


Fig. 5 Temperature dependence of magnetization M of bnMn measured
crystalline bnMn in heating and cooling modes. The sharp pyroelectric
in zero-field-cooling (ZFC) and field-cooling (FC) regimes in the external
peaks indicate a ferroelectric transition. The inset shows the estimated
magnetic field H. Left inset: M vs. H measured at constant temperature
electric polarization obtained by integrating the pyroelectric current with
T upon increasing and decreasing field (closed and open symbols, respectively).
time as a function of temperature.
Right inset: Real (w 0 ) and imaginary (w00 ) parts of AC magnetic susceptibility wAC
measured as a function of temperature in the zero applied magnetic field for
different frequencies fAC of the probing magnetic field. Solid lines serve as a
can be most likely attributed to the presence of some amount of guide to the eye, and the arrows mark the ordering temperature Tm.
water absorbed by the studied pellet. The dielectric properties of
bnMn differ significantly from those reported for the bnMg
analogue,17 i.e., the dielectric permittivity values of bnMn are an bifurcation as well as very weak magnetic hysteresis (hardly
order of magnitude smaller. Furthermore, whereas the Mg visible in the field dependence of the magnetization M(H)
analogue exhibited pronounced dielectric relaxation attributed plotted in the left inset to Fig. 5) shows that the ferromagnetism
to the relaxor-like behavior of the sample, frequency dependence is very weak in the compound studied.
of the tangent loss of bnMn shows the weak relaxor nature of the In order to clarify the nature of magnetic ordering in bnMn
electric ordering. we measured its AC magnetic susceptibility wAC in the zero
In order to clarify the ferroelectric nature of the phase transition, magnetic field and in the ZFC regime. As can be inferred from
we have performed the pyroelectric current measurements. As the right inset to Fig. 5 a distinct cusp visible in the real part w 0
seen in Fig. 4, a distinct pyroelectric peak is observed at the of wAC at Tm is associated with featureless and temperature
phase transition temperature Tc = 350 K. The estimated electric independent imaginary part w00 . Moreover, there is no significant
polarization value (inset in Fig. 4) is 5.8 mC m2, indicating difference between the wAC curves probed by an alternate
good ferroelectric properties. It is worth adding here that in the magnetic field of different frequencies fAC (for the sake of clarity
family of metal formate frameworks pyroelectric current was only the data collected at two terminal frequencies are plotted in
successfully measured only for perovskite-type [(CH3)2NH2]- Fig. 5). Such behavior points out clearly at antiferromagnetic
[Mn(HCOO)3].11 Ferroelectric properties were also experimen- character of the ordering at 0 Oe.
tally confirmed by observation of the dielectric hysteresis loops Taking into account the above findings one can conclude
for [NH4][M(HCOO)3] chiral formates.3 Thus bnMn is the third- that bnMn is most probably a weak ferromagnet, i.e. an anti-
type of metal formate templated by protonated amine for which ferromagnetically ordered compound with a small canting of
ferroelectric properties are proved. In contrast to the previously the spins in the underlying antiferromagnetic lattice. The origin
reported compounds templated by monovalent cations (NH4+ of the ferromagnetism in such systems is a small ferromagnetic
or DMA+), which crystallize in the chiral and perovskite architecture, component of the magnetic moments, which is produced
respectively, bnMn is templated by diammonium cation (bnH22+) perpendicular to the spin-axis of the antiferromagnet by an
and it crystallizes in the niccolite architecture. antisymmetric exchange or single-ion anisotropy.34,35 The magnetic
behavior of bnMn is similar to that observed in other metal–
Magnetic properties organic frameworks with Mn, e.g. [(CH3)2NH2][Mn(HCOO)3]
and [NH4][Mn(HCOO)3].3,11
Temperature dependence of the magnetization M(T) of bnMn
measured in an external magnetic field H of 100 Oe is displayed
in the main panel of Fig. 5. A distinct anomaly visible in the Temperature-dependent Raman and IR studies
M(T) curve at low temperatures indicates that the compound Temperature-dependent Raman and IR spectra are presented in
orders magnetically below Tm = 9.0(1) K. Bifurcation of M(T) Fig. 6 and Fig. S4–S6 (ESI†). Table S5 (ESI†) lists the wavenumbers
appearing in the ordered region upon changing the cooling of the observed modes together with the proposed assignment.
regime of the sample suggests the ferromagnetic character of Assignment is proposed by comparison of the bnMn spectra with
the phase transition at Tm. However, a small magnitude of that the spectra of other formates,8,23,36 and experimental as well as

3190 | J. Mater. Chem. C, 2016, 4, 3185--3194 This journal is © The Royal Society of Chemistry 2016
View Article Online

Journal of Materials Chemistry C Paper

theoretical studies of neutral bn and related 1,3-diaminopropane note that some of the IR and Raman bands of the high-
molecules, bnH22+ cations in the C–H and N–H stretching temperature phase disappear in the low-temperature phase or
regions, [NH3(CH2)4NH3]BiCl5 and [NH3(CH2)4NH3]BiCl5H2O exhibit very pronounced changes in intensities and shifts. This
compounds.37–42 behavior is especially pronounced in the 850–1200 cm1 region
Fig. 6 and Fig. S4 (ESI†) show that the Raman spectra remain (see Fig. 6b and Fig. S6c, ESI†), where most bands can be
qualitatively the same in the 360–380 K range. The spectrum attributed to vibrations of the NH3+ groups and the skeleton.
measured at 350 K is, however, significantly different. Firstly, This behavior can be attributed to the change in the conformation
the bands in the 2800–3050 cm1 range are narrower. As a of the bnH22+ ion during the phase transition. Indeed, former
result, the band at 2862 cm1 visible only as a weak shoulder at studies of polyamines and sodium oleate showed that the n(CC)
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

360 K, is clearly observed as a well-resolved band at 350 K. bands are very sensitive to the conformational changes.43,44 For
Secondly, some bands exhibit a sudden change in intensity. For instance, the antisymmetric n(CC) Raman band of sodium oleate
instance, at 360 K the 871 cm1 Raman band is much weaker observed at 1063 cm1 for the all-trans conformer shifted to
than the 790 cm1 band but at 350 K the intensity of this band 1086 cm1 for the gauche conformer.43 The observed shift of
significantly increases (see Fig. 6b). Thirdly, the band at the n(CC) Raman bands of bnMn from 998 cm1 for the high-
1025 cm1 disappears. Clear narrowing of bands and a sudden temperature phase (at 380 K) to 1005 cm1 for the low-
change in intensities are also observed in the IR spectra (Fig. S5 temperature polymorph (at 350 K) is consistent with the change
and S6, ESI†). For instance, the full width at half maximum from all-trans conformation in the high-temperature phase
(FWHM) of the 441 cm1 IR band decreases from 49 cm1 at to gauche conformation in the monoclinic phase. Finally, it is
440 K to 19.8 cm1 at 340 K. It is worth noting that the most worth noting that the IR spectra change very weakly when
significant narrowing of bands is observed for the bands temperature is lowered from 120 K to 5 K (Fig. S5 and S6, ESI†).
related to the bnH22+ cation, for instance the ds(NH3) band This behavior indicates that bnMn does not undergo any
observed at 1516 cm1 (see Fig. S6a, ESI†). This behavior is temperature-induced phase transition below 120 K.
consistent with the postulated order–disorder mechanism of
the phase transition. Our data show, however, that many bands Pressure-dependent Raman scattering studies
exhibit further pronounced narrowing upon cooling the sample The pressure-dependent Raman spectra are presented in Fig. 7.
below 340 K. This behavior indicates that the bnH22+ cations Fig. S7 (ESI†) shows pressure dependence of the wavenumbers,
have some degree of reorientational motions in the low- which can be well described using a linear function o(P) = o0 +
temperature phase and a complete freezing of these motions aP. Table S6 (ESI†) summarizes the values of wavenumber
occurs below 140 K. Low-temperature data also reveal that the intercepts at zero pressure (o0) and pressure coefficients (a =
phase transition leads to splitting of many bands related to do/dP), obtained from fitting of the experimental data by linear
HCOO ions (see Fig. 6 and Fig. S6, and Table S5, ESI†). This functions.
behavior shows that ordering of cations is associated also with When pressure increases, the spectra exhibit weak changes
distortion of the manganese formate framework. It is interesting to up to 0.5 GPa. At 0.9 GPa, the intensity of the 877 cm1 band

Fig. 6 Detail of the Raman spectra corresponding to the spectral ranges 1310–1400, 750–1100 and 50–500 cm1 at different temperatures in
heating run.

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. C, 2016, 4, 3185--3194 | 3191
View Article Online

Paper Journal of Materials Chemistry C


Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

Fig. 7 Pressure-dependent Raman spectra of bnMn recorded during the compression experiment.

strongly decreases and a new band is clearly observed at when pressure reaches 1.9 GPa. A strong intensity increase can
1038 cm1 (see Fig. 7). Furthermore, the bands in the 1350– also be noticed for the 1368 cm1 band. Fourthly, some modes
1420 and 100–250 cm1 regions exhibit large broadening. exhibit a weak but discontinuous shift. For instance, the 878,
These changes indicate that bnMn undergoes a structural phase 1005 and 1072 cm1 bands shift towards higher wavenumbers
transition between 0.5 and 0.9 GPa. It is worth noting that the by nearly 5, 6 and 8 cm1, respectively. These changes indicate
observed changes in the Raman spectra are very similar to those that bnMn exhibits a second pressure-induced phase transition
observed at the temperature-induced phase transition near between 1.3 and 1.9 GPa. A significant and abrupt shift of bands,
350 K. It is, therefore, very likely that pressure induced transition which can be attributed to vibrations of the bnH22+ cations,
into the trigonal disordered phase observed at ambient pressure indicates that bnMn undergoes a first-order phase transition that
above 350 K. Since the phase transition has a first-order nature, the leads to significant strengthening of the interactions between the
change in Tc upon applying pressure can be described by the bnH22+ cations and the manganese formate framework. The
Claussius–Clapeyron equation: appearance of additional bands suggests the lowering of the crystal
symmetry. The observed shifts of bands corresponding to the
@Tc Tc ðDV=VÞ
¼ Vmolar (1) formate ions are weak and this result indicates that the transition
@p DHmolar
leads to weak distortion of the manganese formate framework.
where Tc is the transition temperature, DV/V is the relative volume When pressure increases beyond 1.9 GPa, Raman bands become
change at Tc, Vmolar is the molar unit cell volume and DHmolar broader and less intense and the spectra become very weak above
represents the change in the molar enthalpy. The negative 3.4 GPa. This behavior suggests that bnMn starts to amorphize at
pressure dependence of Tc means that a negative change of high pressures. It is worth noting that amorphization, irreversible or
the volume takes place at the phase transition. Indeed, our reversible, was observed for many MOFs.20 For porous MOFs,
X-ray diffraction data show that DV/V is negative (about 0.002 amorphization is sometimes observed at very low pressures. For
according to Fig. S3, ESI†). Taking Tc as 350 K, DHmolar = instance, the Sc2BDC3 MOF (BDC = 1,4-benzenedicarboxylate)
2.6 kJ mol1 and Vmolar = 0.0002704 m3 mol1 we can estimate exhibits reversible amorphization already at 0.14 GPa.20,45
qTc/qp to be about 73 K GPa1. Thus one would expect to Raman spectra of bnMn during the decompression are presented
observe the ferroelectric phase transition at about 0.7 GPa. As in Fig. S8 (ESI†). The observed changes in the Raman spectra are
can be noticed, the estimated value is in excellent agreement similar to those observed upon compression indicating reversibility
with the high-pressure experiment showing the phase transition of the amorphization-like behavior and phase transitions.
between 0.5 and 0.9 GPa.
Upon further increase of pressure, significant changes occur Conclusions
in the spectra when pressure changes from 1.3 to 1.9 GPa.
Firstly, new bands appear at 215, 295 and 988 cm1. Secondly, We report synthesis and characterization of bnMn niccolite.
the 787 and 793 cm1 bands merge into one band. Thirdly, the DSC and detailed temperature-dependent X-ray diffraction,
intensity of the 1005 cm1 band (at 1.3 GPa) strongly increases dielectric, Raman and IR data show that this compound

3192 | J. Mater. Chem. C, 2016, 4, 3185--3194 This journal is © The Royal Society of Chemistry 2016
View Article Online

Journal of Materials Chemistry C Paper

undergoes a first-order phase transition from the P3% 1c to Cc 7 (a) P. Jain, V. Ramachandran, R. J. Clark, H. D. Zhou,
phase near 350 K. Pyroelectric current shows a distinct peak B. H. Toby, N. S. Dalal, H. W. Kroto and A. K. J. Cheetham,
near 350 K proving that this formate has ferroelectric properties J. Am. Chem. Soc., 2009, 131, 13625–13627; (b) M. Sánchez-
below 350 K. X-ray diffraction and vibrational data also reveal Andújar, S. Presedo, S. Yáñez-Vilar, S. Castro-Garcia,
that the phase transition is associated with the change of the J. Shamir and M. A. Señaris-Rodriguez, Inorg. Chem., 2010,
bnH22+ conformation from all trans in the high-temperature 49, 1510–1516.
phase to gauche in the low-temperature phase. Detailed structural 8 M. Ma ˛czka, A. Ga ˛gor, B. Macalik, A. Pikul, M. Ptak and
analysis also shows that both organic cations and distortion of the J. Hanuza, Inorg. Chem., 2014, 53, 457–467.
manganese formate framework contribute to the spontaneous 9 L. C. Gómez-Aguirre, B. Pato Dolán, A. Stroppa, S. Yáñez-
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

polarization. Another interesting observation is that this com- Vilar, L. Bayarjargal, B. Winkler, S. Castro-Garcia, J. Mira,
pound orders magnetically below 9.0 K. M. Sánchez-Andújar and M. A. Señaris-Rodriguez, Inorg.
In addition to the temperature-dependent studies, we also Chem., 2015, 54, 2109–2116.
performed high-pressure Raman studies of this compound. 10 (a) S. Chen, R. Shang, K.-L. Hu, Z.-M. Wang and S. Gao, Inorg.
The obtained data show strong evidence for two pressure- Chem. Front., 2014, 1, 83–98; (b) G. Kieslich, S. Kumagai, K. T.
induced phase transition near 0.5–0.9 and 1.3–1.9 GPa. The Butler, T. Okamura, C. H. Hendon, S. Sun, M. Yamashita,
first transition leads most likely to transformation of the crystal A. Walsh and A. C. Cheetham, Chem. Commun., 2015, 51,
into the trigonal paraelectric phase, observed at ambient pressure 15538–15541.
above 350 K. Thus pressure leads to suppression of its ferro- 11 W. Wang, L.-Q. Yan, J.-Z. Cong, Y.-L. Zhao, F. Wang,
electric properties. The second transition is most likely associated S.-P. Shen, T. Zhou, D. Zhang, S.-G. Wang, X.-F. Han and
with some reorientation and/or conformational change of the Y. Sun, Sci. Rep., 2013, 3, 2024.
bnH22+ ion, which leads to strengthening of the hydrogen bonds 12 (a) M. Sánchez-Andújar, S. Presedo, S. Yáñez-Vilar, S. Castro-
and lowering of crystal symmetry. We also observe partial Garcia, J. Shamir and M. A. Señaris-Rodriguez, Inorg. Chem.,
amorphization of the sample when pressure increases beyond 2010, 49, 1510–1516; (b) M. Sánchez-Andújar, L. C. Gómez-
3 GPa. This process is, however, fully reversible when pressure Aguirre, P. Pato-Dolán, S. Yáñez-Vilar, R. Artiaga, A. L. Llamaz-
decreases from 4.2 GPa to ambient pressure. Saiz, R. S. Manna, F. Schnelle, M. Lang, F. Ritter, A. A.
Haghighirad and M. A. Señarı́s-Rodrı́guez, CrystEngComm,
2014, 16, 3558–3566; (c) M. Kosa and D. T. Major, CrystEng-
Acknowledgements Comm, 2015, 17, 295–298.
13 S. Chen, R. Shang, B.-W. Wang, Z.-M. Wang and S. Gao,
The authors acknowledge the Brazilian National Research
Angew. Chem., Int. Ed., 2015, 54, 11093–11096.
Council (CNPq) for a fellowship and grant 401849/2013-9.
14 M. Ma ˛czka, A. Sieradzki, B. Bondzior, P. Dereń, J. Hanuza
and K. Hermanowicz, J. Mater. Chem. C, 2015, 3, 9337–9345.
References 15 Z. Wang, K. Hu, S. Gao and H. Kobayashi, Adv. Mater., 2010,
22, 1526–1533.
1 R. Shang, S. Chen, Z. M. Wang and S. Gao, in Metal-Organic 16 (a) Z. Wang, X. Zhang, S. R. Batten, M. Kurmoo and S. Gao,
Framework Materials, ed. R. L. MacGillivray and C. M. Lukehart, Inorg. Chem., 2007, 46, 8439–8441; (b) M.-Y. Li, M. Kurmoo,
John Wiley & Sons Ltd., 2014, pp. 221–238. Z.-M. Wang and S. Gao, Chem. – Asian J., 2011, 6, 3084–3096.
2 (a) Z. Wang, B. Zhang, K. Inoue, H. Fujiwara, T. Otsuka, 17 R. Shang, G.-C. Xu, Z.-M. Wang and S. Gao, Chem. – Eur. J.,
H. Kobayashi and M. Kurmoo, Inorg. Chem., 2007, 46, 2014, 20, 1146–1158.
437–445; (b) M. Ma ˛czka, A. Pietraszko, B. Macalik and 18 R. Shang, S. Chen, K.-L. Hu, Z.-C. Jiang, B.-W. Wang,
K. Hermanowicz, Inorg. Chem., 2014, 53, 787–794; (c) M. M. Kurmoo, Z.-M. Wang and S. Gao, APL Mater., 2014,
Ma ˛czka, K. Szymborska-Małek, A. Ciupa and J. Hanuza, Vib. 2, 124104.
Spectrosc., 2015, 77, 17–24; (d) J. Xu, B. E. G. Lucier, 19 D. Aulakh, J. R. Varghese and M. Wriedt, Inorg. Chem., 2015,
R. Sinelnikov, V. V. Terskikh, V. N. Staroverov and 54, 8679–8684.
Y. Huang, Chem. – Eur. J., 2015, 21, 14348–14361. 20 S. C. McKellar and S. A. Moggach, Acta Crystallogr., Sect. B:
3 G. C. Xu, W. Zhang, X. M. Ma, Y. H. Hen, L. Zhang, H. L. Cai, Struct. Sci., Cryst. Eng. Mater., 2015, 71, 587–607.
Z. M. Wang, R. G. Xiong and S. J. Gao, J. Am. Chem. Soc., 21 (a) W. Li, M. R. Probert, M. Kosa, T. D. Bennett,
2011, 133, 14948–14951. A. Thirumurugan, R. P. Burwood, M. Parinello, J. A. K. Howard
4 J. M. M. Lawler, P. Manuel, A. L. Thompson and P. J. Saines, and A. K. Cheetham, J. Am. Chem. Soc., 2012, 134, 11940–11943;
Dalton Trans., 2015, 44, 11613–11620. (b) M. Ma ˛czka, P. Kadłubański, P. T. C. Freire, B. Macalik,
5 B. Liu, R. Shang, K.-L. Hu, Z.-M. Wang and S. Gao, Inorg. W. Paraguassu, K. Hermanowicz and J. Hanuza, Inorg. Chem.,
Chem., 2012, 51, 13363–13372. 2014, 53, 9615–9624.
6 (a) Z. Wang, B. Zhang, T. Otsuka, K. Inoue, H. Kobayashi and 22 W. Li, A. Thirumurugan, P. T. Barton, Z. Lin, S. Henke,
M. Kurmoo, Dalton Trans., 2004, 2209–2216; (b) M. Ma ˛czka, H. H. M. Yeung, M. T. Wharmby, E. G. Bithell, J. A. K.
A. Ciupa, A. Ga ˛gor, A. Sieradzki, A. Pikul, B. Macalik and Howard and A. K. Cheetham, J. Am. Chem. Soc., 2014, 136,
M. Drozd, Inorg. Chem., 2014, 53, 5260–5268. 7801–7804.

This journal is © The Royal Society of Chemistry 2016 J. Mater. Chem. C, 2016, 4, 3185--3194 | 3193
View Article Online

Paper Journal of Materials Chemistry C

23 M. Ma ˛czka, T. Almeida da Silva, W. Paraguassu, M. Ptak and 33 D. Orobengoa, C. Capillas, M. I. Aroyo and J. M. Perez-Mato,
K. Hermanowicz, Inorg. Chem., 2014, 53, 12650–12657. J. Appl. Crystallogr., 2009, 42, 820–833.
24 M. Ma ˛czka, T. Almeida da Silva, W. Paraguassu and K. Pereira 34 I. Dzyaloshinsky, J. Phys. Chem. Solids, 1958, 4, 241–255.
da Silva, Spectrochim. Acta, Part A, 2016, 156, 112–117. 35 T. Moriya, Phys. Rev., 1960, 120, 91–97.
25 S. Ghosh, D. Di Sante and A. Stroppa, J. Phys. Chem. Lett., 36 (a) M. Maczka, J. Hanuza and A. A. Kaminskii, J. Raman
2015, 6, 4553–4559. Spectrosc., 2006, 37, 1257–1264; (b) A. L. Magalhaes, S. R. R. S.
26 E. C. Spencer, M. S. R. N. Kiran, W. Li, U. Ramamurty, Mandail and M. J. Ramos, Theor. Chem. Acc., 2000, 105, 68–76.
N. L. Ross and A. K. Cheetham, Angew. Chem., Int. Ed., 2014, 37 E. Kasap and S. Öscelik, J. Inclusion Phenom. Mol. Recognit.
53, 5583–5586. Chem., 1997, 28, 259–267.
Published on 04 March 2016. Downloaded by Federal University of Ceará on 14/11/2016 13:22:43.

27 (a) K. F. Wang, J. M. Liu and Z. F. Ren, Adv. Phys., 2009, 58, 38 A. M. Amorim da Costa, M. P. M. Marques and L. A. E. Batista
321–448; (b) N. A. Hill, J. Phys. Chem. B, 2000, 104, 6694–6709. de Carvalho, J. Raman Spectrosc., 2003, 34, 357–366.
28 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 39 M. P. M. Marques, L. A. E. Batista de Carvalho and
2008, 64, 112–122. J. Tomkinson, J. Phys. Chem. A, 2002, 106, 2473–2482.
29 B. Pato Dolán, M. Sánchez-Andújar, L. C. Gómez-Aguirre, 40 A. L. M. Batista de Carvalho, S. M. Fiuza, J. Tomkinson,
S. Yáñez-Vilar, J. Lopez-Beceiro, C. Gracia-Fernandez, A. A. L. A. E. Batista de Carvalho and M. P. M. Marques, Spectrosc.
Haghighirad, F. Ritter, S. Castro-Garcia and M. A. Señaris- Int. J., 2012, 27, 403–413.
Rodriguez, Phys. Chem. Chem. Phys., 2012, 14, 8498–8501. 41 H. Jeghnou, A. Ouasri, A. Rhandour, M. C. Dhamelincourt,
30 (a) J.-P. Zhao, B.-W. Hu, F. Lloret, J. Tao, Q. Yang, X.-F. Zhang P. Dhamelincourt, A. Mazzah and P. Roussel, J. Raman
and X.-H. Bu, Inorg. Chem., 2010, 49, 10390–10399; Spectrosc., 2005, 36, 1023–1028.
(b) A. Ciupa, M. Ma ˛czka, A. Ga˛gor, A. Sieradzki, J. Trzmiel, 42 A. Rhandour, A. Ouasri, P. Roussel and A. Mazzah, J. Mol.
A. Pikul and M. Ptak, Dalton Trans., 2015, 44, 8846–8854; Struct., 2011, 990, 95–101.
(c) A. Ciupa, M. Ma ˛czka, A. Ga˛gor, A. Pikul and M. Ptak, 43 P. T. T. Wong and H. H. Mantsch, J. Phys. Chem., 1983, 87,
Dalton Trans., 2015, 44, 13234–13241; (d) M. Ma ˛czka, 2436–2443.
A. Ciupa, A. Ga ˛gor, A. Sieradzki, A. Pikul and M. Ptak, 44 J. P. Anastassopoulou, M. Berjot, J. Marx, C. M. Paleos,
J. Mater. Chem. C, 2016, 4, 1186–1193. T. Theophanides and A. J. P. Alix, J. Mol. Struct., 1997, 415,
31 M. P. M. Marques and L. A. E. Batista de Carvalho, COST 917: 225–237.
Biogenically active amines in food, 2000, vol. IV, pp. 122–129. 45 A. J. Graham, A. M. Banu, T. Düren, A. Greenaway, S. C.
32 K. Tichy, J. Benes, R. Kind and H. Arend, Acta Crystallogr., McKellar, J. P. S. Mowat, K. Ward, P. A. Wright and S. A.
Sect. B: Struct. Crystallogr. Cryst. Chem., 1980, 36, 1355–1367. Moggach, J. Am. Chem. Soc., 2014, 136, 8606–8613.

3194 | J. Mater. Chem. C, 2016, 4, 3185--3194 This journal is © The Royal Society of Chemistry 2016

Você também pode gostar