M.
Węcławik
a,
A.
Gągor
b,
R.
Jakubas
a,
A.
Piecha-Bisiorek
*a,
W.
Medycki
c,
J.
Baran
b,
P.
Zieliński
de and
M.
Gałązka
e
aFaculty of Chemistry, University of Wrocław, F. Joliot-Curie 14, 50-383 Wrocław, Poland. E-mail: anna.piecha@chem.uni.wroc.pl
bW. Trzebiatowski Institute of Low Temperature and Structure Research PAS, P.O. Box 1410, 50-950 Wrocław, Poland
cInstitute of Molecular Physics, Polish Academy of Science, Smoluchowskiego 17, 60-179 Poznań, Poland
dCracow University of Technology, Institute of Physics, ul. Podchorążych 1, 30-084 Kraków, Poland
eInstitute of Nuclear Physics, Polish Academy of Science, Radzikowskiego 152, 31-342 Kraków, Poland
First published on 22nd August 2016
Two hybrid crystals imidazolium iodoantimonate(III) and iodobismuthate(III), (C3H5N2)3[Sb2I9] (ImIA) and (C3H5N2)3[Bi2I9] (ImIB), have been synthesized and characterized in a wide temperature range (100–350 K) by means of X-ray diffraction, dielectric spectroscopy, proton magnetic resonance (1H NMR), FT-IR spectroscopy and optical observations. They undergo two temperature induced solid–solid structural phase transitions. The first one, quasi-continuous (with temperature hysteresis below 1 K), occurs at 324 K in ImIA and 327 K in ImIB, and the second one, clearly of the first order, at 273/278 (cooling/heating) and 291/295 K, in ImIA and ImIB, respectively. Ferroelastic properties are maintained in low-temperature phases. Both materials are isomorphic in the corresponding phases. High temperature phase I has a hexagonal P63/mmc symmetry, and phase II has orthorhombic Cmcm. The crystal architecture is composed of discrete, face-sharing bioctahedra [M2I9]3− (M: Sb, Bi) and imidazolium cations which are highly disordered over phases I and II. The dynamics of the imidazolium cations has a prominent impact on the stability of the particular phases.
In recent years, a new class of photovoltaic organic–inorganic perovskite-like materials have started to gain increasing attention. The chemical formula of these layered hybrid compounds is given by (CH3NH3)[MX3] (M = Sn, Pb and X = Cl, Br, I).16–20 Their general traits are a good charge carrier mobility, high absorption coefficient and tailored electronic parameters. The best characterized and the most promising for applications (solar cells) is (CH3NH3)[PbI3].16 Its serious drawback is, however, a content of toxic lead. In the search for new, less harmful materials, bismuth-based derivatives have been, thus, explored. In particular, methylammonium iodobismuthate(III) (CH3NH3)3[Bi2I9] has been reported as a new potential absorber for photovoltaics.21–24 It crystallizes in the space group P63/mmc of a hexagonal system. Its structure consists of discrete [Bi2I9]3− bioctahedra and highly disordered methylammonium cations. Guanidinium analogs, [C(NH2)3]3[Sb2I9]25 and [C(NH2)3]3[Bi2I9],25 exhibiting a complex sequence of phase transitions, appear to be isomorphic to (CH3NH3)3[Bi2I9]
26 in the high-temperature disordered phases. Since [C(NH2)3]3[Bi2I9] shows structural similarities to its methylammonium analogs one can expect promising properties useful, e.g., in photovoltaics.
Similar halogenoantimonates(III) and halogenobismuthates(III) based on the guanidinium crystallize with three different chemical compositions: R2MX5, R3M2X9 and R3MX6.27–29 The findings concerning (CH3NH3)3[Bi2I9] and its guanidinium analogs have motivated us to extend the study on other R3M2I9 iodide compounds to check how the replacement of the guanidinium or methylammonium cations by imidazolium modifies the structure and ferroicity of the compound.
The aim of the present paper is to give an account of the physicochemical properties of two new ferroic materials containing imidazolium cations and iodobismuthate(III) and iodoantimonate(III) anions, respectively. We present here the results of single-crystal X-ray diffraction, the proton magnetic resonance (spin–lattice relaxation T1) and FT-IR as a function of temperature. We also report on our studies of thermal properties (differential scanning calorimetry (DSC)) of (C3H5N2)3[Sb2I9] (ImIA) and (C3H5N2)3[Bi2I9] (ImIB). The microscopic mechanism of the ferroic phase transitions is discussed.
The volume of the unit cell (non-primitive) in the Cmcm structure is, however, doubled due to the C-centering. The hexagonal cell is transformed to the orthorhombic one by the matrix [1−10 110 001] and the relationship between the magnitude of the new and old lattice parameters is as follows: . Fig. 2 illustrates the unit cells and the crystal packing in both phases.
The anionic substructure is built of discrete, face-sharing [M2I9]3− bioctahedra with M–I distances and I–M–I angles characteristic of this type of compound.15,30,31 The ‘trans’ effect32,33 observed in halogenoantimonates(III) and halogenobismuthates(III) is preserved in both phases of ImIA and ImIB. The bridging M–Ib bonds are considerably longer than the terminal M–It distances. The difference of M–I distance Δd ∼ 0.35 Å for ImIA and Δd ∼ 0.30 Å for ImIB at room temperature (see Table 1) is conserved in the PT, also the I–M–I angles do not change significantly. Thus, the bioctahedra are stable in both phases and experience only minor distortion reflecting the site symmetry reduction from D3h (its centre of gravity is at c 6(–)m2) in the hexagonal phase to C2v (the centre of gravity is at c m2m).
ImIA | ImIB | ||||||
---|---|---|---|---|---|---|---|
445 K | 298 K | 350 K | 298 K | ||||
Symmetry code(s): i −y + 1, x − y + 1, z; ii −x + y, −x + 1, z; iii −x + y, −x + 2, z; iv −y + 2, x − y + 2, z; v −x + y, y, −z + 1/2; vi −x, y, z; vii −x, y, −z + 1/2; viiix, y, −z + 1/2. | |||||||
Sb1–I1 | 2.884(2) | Sb1–I3 | 2.890(2) | Bi1–I1 | 2.941(2) | Bi1–I3 | 2.950(2) |
Sb1–I1i | 2.884(2) | Sb1–3vi | 2.889(2) | Bi1–I1i | 2.941(2) | Bi1–I3vi | 2.950(2) |
Sb1–I1ii | 2.884(2) | Sb1–I4 | 2.887(2) | Bi1–I1ii | 2.941(2) | Bi1–I4 | 2.951(2) |
Sb1–I2i | 3.236(2) | Bi1–I1 | 3.2442(2) | Bi1–I2i | 3.251(2) | Bi1–I1 | 3.248(2) |
Sb1–I2ii | 3.236(2) | Sb1–I2vii | 3.238(2) | Bi1–I2ii | 3.251(2) | Bi1–I2vii | 3.253(2) |
Sb1–I2 | 3.236(2) | Sb1–I2 | 3.238(2) | Bi1–I2 | 3.251(2) | Bi1–I2 | 3.253(2) |
I–Sb–Icis | 82.80(5)–93.90(6) | I–Sb–Icis | 80.56(6)–93.83(6) | I–Bi–Icis | 82.01(5)–93.78(6) | I–Bi–Icis | 82.04(4)–94.05(5) |
I–Sb–Itrans | 172.24(7) | I–Sb–Itrans | 172.41(6) | I–Bi–Itrans | 171.80(7) | I–Bi–Itrans | 171.90(5) |
The most pronounced structural manifestations of the PT I → II in both compounds concern the spatial arrangement of the [M2I9]3− bioctahedra and imidazolium cations. Similarly to [C(NH2)3]3[M2I9] hybrids25 both ImIA and ImIB exhibit a layered anionic framework which accommodates the cations within the layers. The lack of strong directional hydrogen-bonding interactions in the high-temperature phases, that could stabilize the structure, makes this system very prone to the temperature induced displacements and favors the steric effects in the organization of the structural motifs.
In high temperature phase I, both types of the imidazolium cations (A and B) are disordered. Thus, the A type cations are disordered around the site of the C3v (3m, Wyckoff position f) symmetry. Each cation may occupy one of the three possible positions generated around this site. The B type cations are disordered around the site of the D3h (6(–)m2, Wyckoff position b) site symmetry. Each B cation may occupy one of the three symmetry equivalent positions. It is important to say that the molecular ring planes of the cations of type B are strictly perpendicular to the c hexagonal axis, whereas the molecular planes of the A-type cations are almost, although not completely, parallel to the c hexagonal axis.
Lowering of temperature gives rise to large deformations of the anionic framework in phase II. The distance between the layers grows, which is well reflected in the increase in the c lattice parameter from 22.285(3) Å at 445 K to 22.372(2) Å at room temperature for ImIA, but the most pronounced changes take place inside the layers. The crystal significantly expands in the bo direction which is equivalent to in the hexagonal phase; the lattice constant bo increases from 16.20 Å to 16.91 Å for ImIA and from 16.09 Å to 16.75 Å for ImIB with decreasing temperature. Both crystals exhibit a negative thermal expansion in the bo direction which is compensated by a huge compression of the structure in the a direction. In ImIA the a lattice parameters decrease from 9.35 Å to 8.92 Å with lowering temperature, which is almost 4.5% of the initial (hexagonal) value. In ImIB the reduction of the a distance is also substantial: from 9.29 to 8.94 Å. Altogether, this gives a negative thermal expansion for ImIB for which the volume of the primitive cell increases with lowering temperature (from 1682 Å3 at 350 K to 1684 Å3 at room temperature). The volume of ImIA does not change within the 3σ limit (σ-standard uncertainty) and is equal to 1688.1(3) at 445 K and 1687.7(3) at room temperature. The observed structural changes are principally associated with a reconstruction of the cationic substructure. In the hexagonal phase there are two inequivalent imidazolium cations A and B. Both are heavily disordered; see Fig. 2(a) and 3(a). Each of them may occupy at least three equivalent positions with the same site occupation factor equal to 1/3. Large voids of 255/256 Å3 for B and 132/138 Å3 for A, for ImIA/ImIB allows for thermally induced rotation of the cations (calculated by Platon34). The diffused electron density around the A and B positions which is reflected in the large displacement parameters of the ring atoms implies almost free in-plane motions of the cations. In the PT the imidazolium B ions change their spatial arrangement by out-of-plane rotations by 90°. In the new position the B ring plane is parallel to the c-axis, while it was perpendicular to the c-axis in the hexagonal phase. Formally, in the orthorhombic phase the number of the orientation states of B ions is reduced from three to one, in fact, in both phases the imidazolium B may freely rotate in-plane as is evidenced by the diffused electron density, see Fig. S1.† Thus, in both phases the number of imidazolium B states formally may be the same. After the transition all the imidazolium cations are directed with their ring planes almost perpendicular to the a-axis, this explains the contraction of the unit cell in this direction, see Fig. 2(b) and 3(b). The cations A exhibit weaker alternation of the spatial arrangement compared to the cations B. The number of orientation states that may be accommodated by A is reduced from 3 to 2 in the orthorhombic phase where the A cations may adopt two positions related by the symmetry plane perpendicular to the a-axis, see Fig. 2(b). The disorder is maintained due to the increase in the volume of the crystal void which is occupied by the A ion. It doubles after the transition from 132/138 to 264/262 Å3, at the same time the space accessible for the ion B is drastically reduced from 255/256 to 128/127 Å3 in ImIA/ImIB. It is worth noting that the electron density of imidazolium A and B in the plane of the rings is still highly diffused in the orthorhombic phase evidencing some in-plane dynamics (rotations and librations) of the cations. In the orthorhombic phase weak N–H⋯I and C–H⋯I hydrogen bonds appear between the [M2I9]3− bioctahedra and imidazolium cations. Table S1† summarises the geometry whereas Fig. S1† illustrates these interactions. The long donor-to-acceptor distances ranging from 3.50(3) to 3.77(3) Å as well as large atomic displacement parameters for both imidazolium and iodide ions suggest that at room temperature both A and B may easily perform in-plane rotations. Additionally, imidazolium A may dynamically switch between the two positions generated by the mirror plane. Probably the last II → III PT stabilizes the crystal structure. Unfortunately, the crystal structure of phase III could not be resolved due to a heavy twinning.
![]() | (1) |
![]() | ||
Fig. 4 Temperature dependence of the complex dielectric permittivity, (a) real (ε′) and (b) imaginary (ε′′) part, recorded during the cooling cycle for the polycrystalline sample of ImIB. |
![]() | ||
Fig. 5 Temperature dependence of the real part of the complex dielectric permittivity recorded for the polycrystalline sample of ImIA. |
In turn, the dielectric anomalies in ImIA are well visible around two PT points. At 278 K the PT is accompanied by a step-wise change in ε′ with the dielectric increment of the order of 8–12 units. The PT at 324 K reveals a clear anomaly of ε′. Nevertheless, the dielectric relaxation processes were not found in ImIA (more in ESI, Part 2†).
As discussed in ref. 25 and in references cited therein a variant can be obtained from another by a mirror reflection “lost” in the phase transition (see Fig. 6). The domain wall along such a plane is a W wall. The domain wall which is perpendicular to the W wall in the limit of infinitesimal deformation strain is classified as a W′ wall. The angles between the domain walls change for finite strains. The angles between the W and W′ domain walls separating the same variants or a selected variant from the reaming two are:23
![]() | (2a) |
![]() | (2b) |
![]() | (2c) |
![]() | ||
Fig. 6 Three variants VI, VII and VIII of phase Cmcm as a result of symmetry reduction P63/mmc. Thick lines and thin dashed lines correspond to W and W′ wall orientation without orthorhombic strain. |
Using the data for the orthorhombic phase for ImIA at 298 K one easily gets:
(W′I–II,WI–II) ≅ 85.5°, | (3a) |
(WI–II,WI–III) ≅ 57.7°, | (3b) |
(W′I–II,W′III–I) ≅ 55.6°, | (3c) |
(W′I–II,WI–II) ≅ 86.1°, | (4a) |
(WI–II,WI–III) ≅ 58.0°, | (4b) |
(W′I–II,W′III–I) ≅ 56.2°. | (4c) |
![]() | ||
Fig. 7 Ferroelastic domain structure of (a) ImIB and (b) ImIA in the orthorhombic phase along the c-axis. |
The optical observations at lower temperatures (see Fig. S5†) provide an insight into the structure of the phase III of ImIB despite the lack of crystallographic data. The PT II → III turns out to be ferroelastic, although it cannot be stated if phase III is a proper or improper ferroelastic with respect to phase II. A comparison of angles implies that a stress-free orientation of domain walls in phase III is perpendicular to the longest edge of the plate visible in Fig. 7a and Fig. S5.† At the same time the plane halves the angle 56° that should correspond to the (W′I–II, W′I–III) angle in phase II. Such being the case, the longest edge of the plate would be parallel to the plane (a, c) of the orthorhombic phase and the stress-free domain boundary of orientational variants of phase III would be parallel to the plane (b, c). This is in accordance with Fig. 2 and 3 showing that the plane (0 1 0), i.e. the one with the longest edge, in phase is in fact the closest packed. Such planes are usually the preferable facets of the crystal as deduced from the Wulff construction.34
![]() | ||
Fig. 8 1H spin–lattice relaxation time T1versus temperature for ImIB (red circles) and ImIA (blue circles) at 25 MHz and theoretical predictions (solid lines) obtained as a sum of three relaxation processes (eqn (5) and (6)). |
The complex crystal structures of both compounds and their temperature evolution result in a complex spin–lattice relaxation process. It is shaped by the spatial disorder as well as by the fact that the imidazolium cations may adopt different motion modes: (i) small-angle librations, (ii) pseudo-C5 in-plane rotations, (iii) rotations around axes which are in-plane or (iv) tumbling. The three relaxation minima occurring in phase III result from the dynamics of three structurally and dynamically non-equivalent imidazolium cations. It is known that for the dipole–dipole couplings of protons with neighboring nuclei, the most important is the dipole–dipole interaction between protons of the same imidazolium ring. In fact, protons of the imidazolium cation may relax in different relaxation pathways, which are summed up as effective relaxation times (eqn (5)):
![]() | (5) |
![]() | (6) |
At the present stage we are not able to describe exactly the movements of the relaxing imidazolium cations in both the studied compounds in phase III. We may suppose that the different types of the small-angle librations of the imidazolium rings in-plane and out-of plane appear as it is postulated from X-ray diffraction. Noteworthy is that all the parameters with relatively small values of correlation times (typical of ring cations) listed in Table S2† represent a fast motion mode of imidazolium. At low temperatures, when the influence of quadrupole interactions is usually relevant, both the fitting lines have no deviations.43 It means that in ImIA and ImIB the quadrupolar interaction may appear, if at all, at much lower temperatures than these reached in the present study. On the other hand, at higher temperatures in both crystals the fitting lines definitely deviate from experimental points just before III → II PT temperatures. The values of the measured relaxation times significantly higher than those predicted by the theory may be related to the critical effects of upcoming rapid structural reconstruction at the PT. As a result, the deviation of both the observed slopes of the temperature dependencies of longitudinal relaxation times T1 drops to Ea within 1.7–2.0 kcal mol−1. The second moment of the 1H NMR line as a function of temperature (95–340 K) displays only minor anomalies around PTs in both compounds (see Fig. S6†).
The most spectacular changes which occur due to the (III → II) PT are assigned to the vibrations of the ring (R). Fig. 9 shows the evolution of the τ(R) modes in ImIB between 600 and 630 cm−1 with temperature. At the lowest temperature (11 K) seven components are observed at 625, 623, 617, 614, 611, 608 and 606 cm−1. Four of them (625, 611, 606 and 623 cm−1) disappear over phase III at ca. 100 K. With further heating the triplet structure of the τ(R) mode evolves continuously into a doublet. At the PT temperature (295 K) one of the components disappears (608 cm−1) whereas the mode at 614 cm−1 shifts step-wisely by ca. 1 cm−1 towards the higher values.
![]() | ||
Fig. 9 (a) Infrared spectra between 595 and 630 cm−1 of the ImIB as a function of temperature, (b) plot of the τ(R) mode frequencies of ImIB as a function of temperature. |
Fig. 10 shows the region of the deformation ring mode δ(R) (i.e. ν(R) β(CH) β(NH)) (1570–1590 cm−1) as a function of temperature. On cooling from 320 K to 297 K a single band is visible. Below II → III PT a gradual splitting takes place and finally, at 250 K, a new component appears. It should be noticed that the sign of the temperature coefficient (dν/dT) of a central band (1578 cm−1 at 11 K) is changed at ca. 200 K.
![]() | ||
Fig. 10 (a) Infrared spectra between 1570–1590 cm−1 of the ImIB as functions of temperature, (b) plot of the deformation band: δ(R) mode frequencies of ImIB as a function of temperature. |
On the other hand, GIA/GIB and ImIA/ImIB display large similarities both from the structural and dynamical point of view. The PT I → II in all the cases belong to the proper ferroelastic species of Aizu: 6/mmmFmmm.47 In all the crystals the PTs are driven by a mixed order–disorder and displacive mechanism. The latter component is related to significant relative displacement of the anionic and cationic substructures. Nevertheless, the character of cation arrangement is substantially different in both subclasses. One can observe considerable changes in the mutual positions of the A and B guanidinium cations through the PT. The distance between the centers of gravity of both of them diminishes along the b-axis, which is accompanied by a leaning out of the B cations toward the c-axis (up and down). This effect deepens in the subsequent PT (II → III). The ring plane of the cation A in each phase is practically unchanged with regard to its surroundings (see Fig. 10 in ref. 25).
In the case of the imidazolium analogs the observed structural changes through the I → II and II → III PT are essentially assigned to the reconstruction of the cationic substructure. During the PT imidazolium cations type B change their spatial arrangement by the out of plane rotation equal to 90°. This reorientation is an essential order–disorder contribution to the I → II PT. The X-ray diffraction patterns at room temperature indicate that the electron density of imidazolium A and B in the plane of the rings is still highly diffused in phase II evidencing the in-plane dynamics (rotations and librations) of the cations. This, together with the calorimetric and 1H NMR results, allows us to propose the mechanism of II → III PT in both compounds. The entropy transition (ΔStr) for II → III PT in ImIA and ImIB is of the order of 32–33 J mol−1 K−1 (3·Rln 4) which indicates huge changes in the dynamical state of cations. The PT II → III is, most probably, accompanied by a freezing of the motion of cations within the ring plane, which is distinctly reflected on the T1vs. 1/T curve, as a rapid jump. Below II → III PT a small-angle libration motion is admitted that is confirmed by the dielectric relaxation process observed in ImIB at low temperatures. The 1H NMR studies suggest the presence of three non-equivalent imidazolium cations in phase (III). The structure of the phase III of ImIA and ImIB could not be resolved by XRD. In spite of the huge thermal effect the samples do not lose their integrity in the PT II → III. The microscopy image of the ImIB (Fig. S5†) suggests a ferroelastic character of the transformation with a stress-free orientation of domain walls perpendicular to the long edge of the sample. This orientation apparently bisects an angle of the stress-free domain walls in phase II. Therefore one can expect that the transformation shear plane lies in the (b,c) crystallographic plane of phase II. Such a shear can be engendered by an ordering of the A imidazolium cations if they all selected the left (or right in the other domain) position out of two shown in Fig. 3(b).
The crystals were grown by a slow evaporation of a concentrated HI solution containing the 3:
2 ratio of C3N2H4 and SbI3 (ImIA) or BiI3 (ImIB). The salts obtained were twice re-crystallized and their composition was verified by an elemental analysis: ImIB C: 8.25% (theor. 8.34%), N: 6.77% (theor. 6.49%), H: 1.55% (theor. 1.17%); ImIA C: 8.89% (theor. 9.08%), N: 7.35% (theor. 7.06%), H: 1.55% (theor. 1.27%). The single-crystals were grown from an aqueous solution at constant room temperature.
(C3H5N2)3[Sb2I9] (ImIA) | (C3H5N2)3[Bi2I9] (ImIB) | |||
---|---|---|---|---|
Crystal data | ||||
Chemical formula | C9H15I9N6Sb2 | C9H15Bi2I9N6 | ||
M r | 1592.77 | 1592.89 | 1767.33 | 1767.33 |
Crystal system, space group | Hexagonal, P63/mmc | Orthorhombic, Cmcm | Hexagonal, P63/mmc | Orthorhombic, Cmcm |
Temperature (K) | 445 | 298 | 350 | 300 |
a, b, c (Å) | 9.3526(9), 9.3526(9), 22.285(3) | 8.9215(7), 16.9118(10), 22.3716(15) | 9.2914(9), 9.2914(9), 22.503(3) | 8.9409(14), 16.749(2), 22.494(3) |
α, β, γ (°) | 90, 90, 120 | 90, 90, 90 | 90, 90, 120 | 90, 90, 90 |
V (Å3) | 1688.1(3) | 3375.4(4) | 1682.4(3) | 3368.5(8) |
Z | 2 | 4 | 2 | 4 |
μ (mm−1) | 9.84 | 18.71 | ||
Crystal size (mm) | 0.20 × 0.08 × 0.01 | 0.13 × 0.05 × 0.05 | ||
Data collection | ||||
T min, Tmax | 0.291, 1.000 | 0.401, 1.000 | ||
No. of measured, independent and observed [I > 2σ(I)] reflections | 15![]() |
18![]() |
13![]() |
17![]() |
R int | 0.130 | 0.074 | 0.044 | 0.060 |
Θ max (°) | 25.7 | 25.7 | 23.5 | 25.7 |
(sin![]() |
0.610 | 0.609 | 0.562 | 0.610 |
Refinement | ||||
R[F2 > 2σ(F2)], wR(F2), S | 0.065, 0.234, 0.95 | 0.065, 0.233, 1.05 | 0.041, 0.101, 0.88 | 0.047, 0.113, 0.85 |
No. of reflections | 652 | 1751 | 520 | 1747 |
No. of parameters | 27 | 48 | 27 | 46 |
No. of restraints | 16 | 13 | 18 | 13 |
Δρmax, Δρmin (e Å−3) | 0.73, −0.59 | 1.06, −0.55 | 0.59, −0.33 | 0.72, −0.48 |
Footnote |
† Electronic supplementary information (ESI) available: Crystal characterization, dielectric, vibrational and optical properties as well as proton magnetic resonance studies (1H NMR). CCDC 1494297–1494300. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6qi00260a |
This journal is © the Partner Organisations 2016 |