CONTROVERSY ON THE FERROELECTRICITY IN METAL–FORMATE FRAMEWORKS

P. Peksa and A. Sieradzki

Department of Experimental Physics, Wrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland
Email: adam.sieradzki@pwr.edu.pl

Received 30 September 2022; accepted 4 October 2022

The metal–organic frameworks (MOFs) crystallizing in a perovskite-like architecture became extremely interesting for scientists due to a variety of applications including memory devices, energy conversion and drug delivery. These compounds are constructed from a metal–oxygen or metal–nitrogen octahedral coordinated by organic ligands. They exhibit various interesting properties due to their hybrid organic–inorganic nature. However, ferroelectric MOFs still remain scarce and the topic of ferroelectricity raises a lot of controversies. In this article, we will discuss the actual state of knowledge of these specific compounds with a focus on ferroelectric properties. We will try to create an order out of the current confusion that followed attributing ferroelectric properties to metal–formate frameworks without a direct proof.

Keywords: ferroelectricity, phase transitions, metal–formate frameworks, pyroelectric physics

1. Introduction

In the world of science, no one questions the notion that the  metal–organic frameworks belong to the  more exciting compounds. The  definition of metal–organic materials encompasses a  very large number of compounds with a wide variety of chemical compositions and structures. A property that unites the whole family is the homogeneity of the size and shape of the pores in this specific type of compound that allows one to selectively absorb only molecules smaller than the  pore size  [1]. The desirable physical and chemical properties of these structures result from the  versatile feature of different organic and inorganic components. An  additional advantage of the  metal–organic frameworks is the possible applications of the hybrid compounds such as efficient solar cells [2], gas storage devices [3, 4] or multiferroic memories [5, 6]. The potential applications of these structures have resulted from their structure consisting of two main elements: metal ion centres (nodes) and organic ligand connected with each other by coordination bonds [7]. The metal centres and various organic ligands are responsible for different optical [8], electric [9], magnetic [10] and gas-absorption related [1113] properties of these materials. We will focus on one of the families of metal–organic frameworks, metal–formate frameworks, that have formate as the organic ligand. The first properties that have been explored in metal–organic frameworks were magnetic ordering  [14]. The  long distance that the  formate ligand occupies can only mediate a weak magnetic coupling, therefore the  metal–formate frameworks only exhibit a  long-range magnetic coupling below 50 K, hence they are less attractive in terms of practical applications [15]. The most interesting, application-wise, feature of these structures is ferroelectricity. Ferroelectricity is a property of certain structures, which have a spontaneous electric polarization, that can be reversed by the applied external electric field. The characteristic feature of these materials is the PE hysteresis loop. It is well known that all ferroelectric materials are pyroelectric, with the additional property that their natural electrical polarization is reversible. The ferroelectricity is the most desirable feature in new materials, which is mostly controlled by the relative sizes of the ions. Several MOF families exhibit structural transitions to phases with ferroelectric properties [5]. One of the most popular subclasses of such hybrid metal–organic materials showing spontaneous polarization is metal–formate frameworks of the  general chemical formula  [A][M(HCOO)3]n, where An+ is a molecular alkylammonium cation, and M2+ is the  metal centre  [16]. The  interest in HCOO as an organic ligand is not without a reason. It is worth mentioning that these anions are derived from the simplest carboxylic acid and keep all its characteristic properties. They can coordinate one, two or three metal ions in different ways, mediating the magnetic interaction between them. As the simplest carboxylic ligand, formate plays an important role in constructing coordinated framework structures with unique magnetic and porous properties. In addition, it may exhibit a polar structure, which means that it participates as an acceptor in the building of hydrogen bonds, while the role of donors most often belongs to the hydroxyl and amine groups of organic cations.

Regarding the issue of ferroelectricity in metal–organic frameworks, the  majority of such hybrid frameworks exhibit structural phase transition due to the  cooperative ordering of these cations followed by the  deformation of the  framework  [17, 18]. In some cases, the transition occurs to a non-centrosymmetric phase with ferroelectric properties [19]. Additionally, most of the structures with paramagnetic transition metal ions exhibit magnetic ordering in the low temperatures. The simultaneous coexistence of the  spontaneous electric and magnetic polarizations makes such materials single phase multiferroics that are highly attractive for application in four-state memory devices. Spontaneous polarization below the Curie temperature and the possibility to change the direction of polarization by reversing the applied electric field are the two basic requirements for ferroelectricity. The most popular and simplest method of demonstrating the  ferroelectric properties is, of course, the  PE hysteresis loop observations. However, the order–disorder phase transition is relatively often considered ferroelectric despite no direct proof. In addition, the dispute over the ferroelectricity of the banana skin highlighted the problem of assigning ferroelectricity and cautioned the  interpretation of electrical polarization loops [20]. In recent years, there has been a worrying increase in scientific articles where ferroelectric properties have been assigned without conclusive evidence [21, 22]. In this article, we will look at the current state of knowledge about ferroelectricity in metal–formate frameworks. In order to determine ferroelectricity in the compounds, it is possible to use several measurements, from pyroelectric measurements, complex dielectric permittivity, to the direct proof, which is the PE hysteresis loop.

2. The PE hysteresis loop measurements

Characteristic of all ferroelectric compounds is the  presence of a  PE hysteresis loop. For such measurements, the sample must be oriented along the polar axis, which is troublesome for many metal–formate frameworks, because the single crystals are too small. In addition, it should be remembered that this technique requires the ability to interpret the obtained results, because often due to the so-called banana effect, ferroelectricity has been incorrectly assigned (Fig. 1) [22].

The most important result is the  successful measurement of the  electric polarization loop during the  measurement of polarization dependent on the  applied external electric field in NH4[Zn(HCOO)3] below the  phase transition temperature (191  K). The  experimentally determined spontaneous polarization (1.03  μC/cm2) agrees with the  theoretical value of spontaneous polarization (0.96  μC/cm2)  [23]. The  magnetic members of this family also exhibit ferromagnetic (NH4[Co(HCOO)3] and NH4[Ni(HCOO)3]) and antiferromagnetic (NH4[Mn(HCOO)3])  [21, 24] properties.

The first reported and the  most thoroughly studied member of the  hybrid metal–formate frameworks is dimethylammonium ([CH3)2NH2], DMA) [22, 2531]. Most of these compounds exhibit a  single structural phase transition that involves a  cooperative ordering of organic cations. Depending on a metal centre, the phase transition temperatures fall between 160–180 K [25, 27, 29, 3234] with an exception of the  Mg compound (270 K) [35]. Several publications revealed indications of the ferroelectric behaviour of the low-temperature phase in some of these hybrid perovskites, though a  proper polarization hysteresis loop was only obtained for [(CH3)2NH2][Co(HCOO)3] [36] and  [(CD3)2ND2][Co(DCOO)3]  [37, 38]. Surprisingly, [(CD3)2ND2][Co(DCOO)3] has a new phase transition above the room temperature at around 319  K  [38], as well as a  low-temperature structural phase transition, in contrast to [(CH3)2NH2] [Co(HCOO)3] with only one temperature phase transition about 155 K [36]. The dielectric properties above 151 K between II and III phases show ‘classical’ dipolar relaxation. Very similar dielectric behaviours were observed in NaNo2, which is a typical order–disorder type ferroelectric [39]. The proper dielectric hysteresis loop with a saturation polarization of 1.04  μC/cm2 is comparable to that of NH4[Zn(HCOO)3]. This hysteresis loop and the SHG measurement and dielectric properties clearly prove that  [(CH3)2NH2][Co(HCOO)3] and  [(CD3)2ND2][Co(DCOO)3] are relaxor-type ferroelectrics. These types of ferroelectrics have received attention, in part because of their excellent electromechanical properties.

img

Fig. 1. (a) Saturated hysteresis loops for ferroelectric compounds, (b) non-saturated hysteresis loops.

3. The structural measurements

Structural measurements make it possible to indicate changes in symmetry and a  non-centrosymmetric crystal structure. The  determination of the  symmetry of high-temperature and low-temperature phases allows for the  determination of potential ferroic properties in accordance with the  Aizu classification. As expected, ferroelectric materials exhibit a  second harmonic generation (SHG) effect that is sensitive to symmetry breaking during temperature changes (only non-centrosymmetric materials display SHG signals). This technology is used in ceramic ferroelectrics, mainly for confirming symmetry breaking and detecting ferroelectric domains [39, 40]. Similarly, the determination of the dipole moment in two possible states in a unit cell proves the ferroelectricity. However, due to the chemical and geometric complexity of the structures in MOF compounds, an accurate determination of symmetry in phases is troublesome and sometimes even impossible.

For the single crystal NH4[Mg(HCOO)3], X-ray diffraction showed that it crystallizes in the  hexagonal space group P6322, with orientation disordered NH4+ ions located inside of frameworks. After cooling down, the  compound undergoes a  phase transition at around 255  K, interestingly to the ferroelectric structure P6[41]. The authors noted that the same structural changes had previously been reported for other compounds in this family. The data presented in the article indicates that in the  ordered phase NH4[Mg(HCOO)3] three nitrogen atoms are heterogeneous on the X-ray timescale, but the  rotation of the  NH4+  does not freeze completely on the phase transition but shows a further slowdown below 255 K on the Raman and IR timescales [41]. This behaviour proves the highly dynamic nature of hydrogen bonds. Authors have noticed that the phase transition temperature increases with a decrease in mass and an increase in the size of metal cations [41]. According to the  authors, the  obtained results indicate a method of modulating the phase transition temperature and ferroelectric properties by changing the cation type in the metal formate structure. In our opinion, based on Raman, IR and structural measurements it is difficult to determine the  ferroelectric properties of a compound. No hysteresis loop or pyroelectric current measurements were demonstrated for the tested compound. The dielectric measurements were compared with the  only compound with the  NH4 cation, for which ferroelectric properties were demonstrated, namely NH4[Zn(HCOO)3] [23, 42, 43], but can ferroelectric properties be determined from the similarity? This proposal raises serious concerns.

4. The pyroelectric current measurements

The pyroelectric measurement as a  function of temperature is an alternative method. It is a direct measurement of the  polarization derivative over temperature. Due to the averaging of the polarization vector component for polycrystalline samples, this method does not allow for a precise determination of the value of spontaneous polarization, but it gives a picture of changes in this value as a function of temperature.

In 2017, it was reported that the  compound [(CH3)2NH2][Mn(HCOO)3] exhibits a first-order ferroelectric phase transition with a  high polarization, induced by the order–disorder transition of hydrogen bonds [44]. The authors found the ferroelectricity of the compound based on a pyroelectric measurement. It has been suggested that the sharp first-order ferroelectric transition is related to the hydrogen bond ordering process, therefore single crystals  [(CH3)2NH2][Mn(HCOO)3] show a high pyroelectric coefficient (5.16 · 10–2 C/m2K) and a  high thermal expansion coefficient [4448]. In our opinion, the measurement of pyroelectric current alone cannot provide sufficient evidence for the existence of the ferroelectric transition. However, it should be remembered that earlier Sanchez-Andujar and others showed the structure for [(CH3)2NH2][Mn(HCOO)3] and assigned a  monoclinic eccentric space group Cc, which could suggest ferroelectricity [47]. It would seem that the  demonstration of the  change in symmetry and non-centrosymmetricity and, additionally, a large coefficient of pyroelectric current are a sufficient proof of ferroelectric properties. It is worth noting that the  measurements of  [(CH3)2NH2] [Mn(HCOO)3] in the form of mesoscale particles do not show a detectable pyroelectric current [49]. Surface polarization effects limit spontaneous polarization and this is a consequence of a relatively low electrical polarization in this formate family. A similar behaviour can be observed for the analogue [(CH3)3NH2](Mg(HCOO)3) [50].

Let us look at other structures with the (CH3)2NH2 cation. Another compound for which pyroelectric measurements have been made is [(CH3)2NH2][Fe(HCOO)3]  [51]. The  strong dependence of dielectric permittivity on the  frequency indicates the relaxor nature of the electrical order. The  pyroelectric current was measured  – a  small, sharp anomaly is observed at the phase transition temperature. The  resulting electrical polarization of 18  μC/m[51] suggests a  weak ferroelectricity. It is worth noting that currently all measurements of the  second harmonic generation (SHG) and pyroelectric current measurements in DMA+ analogue structures indicate only a pyroelectric, but not a ferroelectric behaviour. Assigning ferroelectric properties to these structures is, in our opinion, improper and it still hampers the  widely discussed potential applications of compounds with this cation in ferroelectric devices. Additionally, we should distinguish proper and improper ferroelectricity [52, 53]. For the  improper ferroelectrics, it will not be possible to measure the hysteresis loop. An interesting case is the  compounds with the  general formula  [CH3NH2NH2][M(HCOO)3] for which it was not possible to obtain direct evidence in the form of a hysteresis loop, but the measurements made allowed one to determine the ferroelectricity [54, 55]. These compounds show the  two structural phase transitions. The first phase transition temperature is practically independent of the type of divalent metal and is observed in 310–327 K from the nonpolar phase R-3c to the polar R3c phase. It is a second type of phase transition associated with the  partial ordering of methylhydrazinium cations. The second, low-temperature phase transition indicates the  first-order character and is related to the  complete ordering of methylhydrazinium cations and deformation of the  metal–formate frameworks. The pyroelectric measurements show the  ferroelectric nature of the  phase transition at room temperature (Fig. 2). Based on the analysis of dielectric measurements, the  improper ferroelectricity was indicated.

img

Fig. 2. Pyroelectric current as a function of temperature after poling  [CH3NH2NH2][M(HCOO)3] from 350 to 150  K with ±2  kV/cm, during heating with the temperature rate of 1 K/min.

Let us return to the already mentioned analogue with the (CH3)2NH2 cation – for the [(CH3)2NH2)] [Co(HCOO)3] and  [(CD3)2ND2)][Co(DCOO)3] compounds it was possible to show a hysteresis loop. The static pyroelectric method, dynamic dielectric measurements and high-pressure Raman spectra were used to study the  ferroelectric order–disorder transition in [(CD3)2ND2)][Co(DCOO)3] [38]. The two phase transitions were noted, at 1.6 and 6.3 GPa. Near 1.6 GPa, there was some correlation between the tendency of the jumping energy of nitrogen atoms in DMA+ cations and the tendency to order due to dipole–dipole interactions; this competition caused a slow decrease in Tc, in contrast to the sharp decrease observed with the KDP-type ferroelectric. The applied pressure suppresses the electrical polarization that is accompanied by a  shift in Tc towards lower temperature. The  ferroelectricity decays at 1.6 GPa, and this is related with the  distortion of the  anionic structure  [38]. It can be seen that the application of pressure has a large influence on the dynamic behaviour of dimethylamine cations and affects the  ferroelectric properties. These studies show the  feasibility of fine-tuning the  ferroelectric properties in MOFs under high pressure.

5. Dielectric spectroscopy

One of the extremely useful electrical techniques is broadband dielectric spectroscopy. Measurements of complex dielectric permittivity as a function of both temperature and frequencies are an often used method for identifying phase transitions. In a phase transition where the electrical polarization changes, the dielectric permittivity values must also change. In the case of ferroelectric materials, the Curie–Weiss law must be met when switching between the paraelectric and the ferroelectric phase. Measurements of the dielectric response as a function of frequency are extremely useful tools in identifying the mechanisms of ferroelectric phase transitions. This technique allows for observation of relaxation processes and connecting them with structural changes of the  compound, consequently determining the  factors leading to the  loss of ferroelectricity.

Initially, it was suggested that the  DMA-M materials indicate anti-ferroelectric properties, but this assumption  –  based only on the  shape of the  temperature dependence of the  dielectric constant  –  turned out to be incorrect  [56]. Ferroelectric phase transition is usually accompanied by a  change in the  symmetry of lattice constants. In addition, studies of pyroelectric current and thermal expansion may provide evidence for ferroelectricity in the studied compounds. The  ferroelectric order below Tc was confirmed by measurements of the second harmonic generation (SHG) and the  existence of an electrical hysteresis loop for  [(CH3)2NH2)] [Co(HCOO)3] [36]. The estimated spontaneous polarization value for this material is 1.02  μC/cm2, which is approximately six times the value obtained with the  prototype hybrid ferroelectric (0.2  μC/cm2)  [15] of Rochelle salt (sodium potassium tartrate). Additionally, the  frequency form of the  dielectric permittivity indicated the  Debye relaxation. The  determined values of activation energy (Ea  =  464  J/mol) and relaxation time (τ = 1.02 · 10–15s) resemble analogous values obtained for traditional ferroelectric crystals showing order–disorder phase transitions, such as NaNO2 [39, 57].

Another interesting metal–formate framework is a  structure with the  general formula  [C2H5NH3][M(HCOO)3]. The  results of experimental studies and theoretical calculations are presented for the  compound  [C2H5NH3] [Na0.5Fe0.5(HCOO)3] [58, 59]. Based on the results of calorimetric tests, the  phase transition temperature (360  K) of the  second type from the high-temperature phase transition to the low-temperature phase was determined. Above this temperature, the dynamic disorder of the [C2H5NH3]+ cations does not cause a significant dispersion in the  dielectric spectra. The  dependence of the  dielectric permittivity as a  function of temperature showed strong changes in the vicinity of the phase transition temperature, which may be characteristic of ferroelectric compounds. The  results of calculations based on the  first principles density functional theory (DFT) clearly confirmed that the  low-temperature phase is ferroelectric. The  conclusions were also based on structural studies showing that non-centrosymmetric and spontaneous polarization in the ferroelectric phase results from the ordering of ethylammonium cations. The estimated value of the electric polarization in  [C2H5NH3][Na0.5Fe0.5(HCOO)3] is 0.8  μC/cm2 and is about four times the  value obtained for the  prototype hybrid ferroelectric (0.2  μC/cm2) of Rochelle salt (sodium tartrate potassium)  [58, 59]. It is worth recalling here that ferroelectric MOF crystals are still rare at room temperature.

[AmineH+] Metal ions MII Symmetry Activation energy Ea, eV Temperature of phase transitions T0, K Physical properties
NH4+

Cu2+

Zn2+

Mg2+

Mn2+

Co2+

Ni2+

Fe2+

Orthorhombic/hexagonal [6]

0.76 [60]

0.17 [43]

0.28 [41]

0.09 [61]

0.094 [61]

225; 360 [60]

192 [61]

255 [61]

254 [61]

191 [24]

199 [61]

212 [61]

Multiferroicity [61]

Ferroelectricity [23]

(CH3)2NH2+

Mn2+

Co2+

Mg2+

Zn2+

Fe2+

Ni2+

Rhombohedral/monoclinic [27]

~1.15 [62]

0.28 [36]

0.30 [35]

0.28 [63]

187 [48, 64]

165 [36]

270 [35]

156 [63]

160 [27]

180 [65]

Multiferroicity [36]

Dielectricity [35]

Magnetism [65]

CH3NH2NH2+

Mn2+

Mg2+

Fe2+

Zn2+

Trigonal [54] 0.76 [66, 67]

220; 310 [54]

235; 327 [54]

181; 310 [54]

166; 322 [67]

Ferroelectricity [58]
NH2-CH+-NH2 Mn2+ Orthorombic [68] 0.46 [68] 280; 334 [68] Magnetism [68]
CH3C(NH2)2+ Mn2+ Orthorombic [69] 0.6; 1.1 [69] 304 [69] Ferromagnetism [69]
(CH2)3NH2+

Cu2+

Mg2+

Mn2+

Zn2+

Monoclinic [70]

Orthorombic [70]

285 [70]

288 [71]

273 [70]

255; 300 [70]

Dielectricity [70]
CH3CH2NH3+

Cu2+

Mg2+

Orthorombic [72] 0.39 [73]

357 [72]

374; 426 [73]

Magnetism [72]
NH2NH3+

Co2+

Fe2+

Mg2+

Mn2+

Zn2+

Orthorombic/hexagonal [74]

1.03 [74]

0.17; 0.11;

0.03 [75]

380 [74]

347; 336 [76]

348 [74]

360; 298 [75]

367 [74]

Multiferroicity [74]
C3N2H5+ Mn2+ Monoclinic [77] 0.83 [77] 438 [77, 78] Ferroelectricity [78]

6. Conclusions

Searching for ferroelectric properties in metal–formate frameworks is an activity based mainly on empirical research. Despite the many experiments carried out on formate compounds of the MOF type, the topic of ferroelectricity still raises a lot of controversies. The dynamics of molecules and built-in cations can induce complex structural phase transitions that reflect the details of the interactions that occur between the  various components in the  structure, When the built-in organic cation and the framework molecule fit together well, creating a long-range bond, ferroelectricity will eventually occur. A key issue for ultra-fine ferroelectric materials is the depolarization effect. It is known that the ferroelectricity of conventional ferroelectric thin films is usually suppressed because the layers are thinner than the critical thickness due to the action of the depolarizing field caused by uncompensated charges in the presence of metal electrodes. Despite the technological advances made, research and the potential applications of ferroelectric metal–formates are still quite limited due to a lack of understanding of ferroelectricity, which requires extensive, complementary research. It is worth noting that ferroelectricity is an interdisciplinary topic on the border of chemistry, physics, materials science, electronics and crystallography. An important issue seems to be the theoretical calculations, such as DFT, that can unequivocally confirm or contradict the ferroelectric property of the compound. There has been a  tendency in recent years to publish information about the  ferroelectric properties of compounds without a direct proof. Only a set of a few experimental techniques and common results provide certainty about the ferroelectric properties of a given material. Of course, the hysteresis loop is the most desirable result, but also NMR, EPR or dielectric spectroscopy measurements can be a window to show the ferroelectric properties of the compound.

References

[1] M. Mączka, T. Almeida da Silva, W. Paraguassu, and K. Pereira Da Silva, Raman scattering studies of pressure-induced phase transitions in perovskite formates  [(CH3)2NH2][Mg(HCOO)3] and  [(CH3)2NH2][Cd(HCOO)3], Spectrochim. Acta A 156, 112–117 (2016).

[2] S. Kitagawa, R. Kitaura, and S.I. Noro, Functional porous coordination polymers, Angew. Chem. Int. Ed. 43(18), 2334–2375 (2004).

[3] Y. He, W. Zhou, G. Qian, and B. Chen, Methane storage in metal–organic frameworks, Chem. Soc. Rev. 43(16), 5657–5678 (2014).

[4] L.J. Murray, M. Dinca, and J.R. Long, Hydrogen storage in metal–organic frameworks, Chem. Soc. Rev. 38, 1294–1314 (2009).

[5] M. Guo, H.L. Cai, and R.G. Xiong, Ferroelectric metal organic framework (MOF), Inorg. Chem. Commun. 13(12), 1590–1598 (2010).

[6] L. Jiao, J.Y.R. Seow, W.S. Skinner, Z.U. Wang, and H.L. Jiang, Metal–organic frameworks: Structures and functional applications, Mater. Today 27, 43–68 (2019).

[7] R.J.  Kuppler, D.J.  Timmons, Q.R.  Fang, J.R.  Li, T.A.  Makal, M.D.  Young, D.  Yuan, D.  Zhao, W.  Zhuang, and H.C.  Zhou, Potential applications of metal–organic frameworks, Coord. Chem. Rev. 253(23), 3042–3066 (2009).

[8] M.D.  Allendorf, C.A.  Bauer, R.K.  Bhakta, and R.J.T. Houk, Luminescent metal–organic frameworks, Chem. Soc. Rev. 38(5), 1330–1352 (2009).

[9] P. Ramaswamy, N.E. Wong, and G.K.H. Shimizu, MOFs as proton conductors – challenges and opportunities, Chem. Soc. Rev. 43(16), 5913–5932 (2014).

[10] M.  Kurmoo, Magnetic metal–organic frameworks, Chem. Soc. Rev. 38, 1353–1379 (2009).

[11] C.  Xu, R.  Fang, R.  Luque, L.  Chen, and Y.  Li, Functional metal–organic frameworks for catalytic applications, Coord. Chem. Rev. 388, 268–292 (2019).

[12] Y.S. Wei, M. Zhang, R. Zou, and Q. Xu, Metal–organic framework-based catalysts with single metal sites, Chem. Rev. 120(21), 12089–12174 (2020).

[13] J.  Lee, O.K.  Farha, J.  Roberts, K.A.  Scheidt, S.T.  Nguyen, and J.T.  Hupp, Metal–organic framework materials as catalysts, Chem. Soc. Rev. 38, 1450–1459 (2009).

[14] X.Y. Wang, Z.M. Wang, and S. Gao, Constructing magnetic molecular solids by employing three-atom ligands as bridges, Chem. Commun. 3, 281–294 (2008).

[15] W. Li, Z. Wang, F. Deschler, S. Gao, R.H. Friend, and A.K.  Cheetham, Chemically diverse and multifunctional hybrid organic–inorganic perovskites, Nat. Rev. Mater. 2, 1–18 (2017).

[16] P.  Jain, V.  Ramachandran, R.J.  Clark, D.Z.  Hai, B.H.  Toby, N.S.  Dalal, H.W.  Kroto, and A.K. Cheetham, Multiferroic behavior associated with an order–disorder hydrogen bonding transition in metal–organic frameworks (MOFs) with the perovskite ABX3 architecture, J. Am. Chem. Soc. 131(38), 13625–13627 (2009).

[17] K. Asadi and M.A. van der Veen, Ferroelectricity in metal–organic frameworks: characterization and mechanisms, Eur. J. Inorg. Chem. 2016(27), 4332–4344 (2016).

[18] W.  Li, Z.  Zhang, E.G.  Bithell, A.S.  Batsanov, P.T.  Barton, P.J.  Saines, P.  Jain, C.J.  Howard, M.A.  Carpenter, and A.K.  Cheetham, Ferroelasticity in a metal–organic framework perovskite; towards a  new class of multiferroics, Acta Mater. 61(13), 4928–4938 (2013).

[19] T.  Hang, W.  Zhang, H.Y.  Ye, and R.G.  Xiong, Metal–organic complex ferroelectrics, Chem. Soc. Rev. 40, 3577–3598 (2011).

[20] J.F.  Scott, Ferroelectrics go bananas, J. Phys. Condens. Matter. 20(2), 021001 (2008).

[21] G.C.  Xu, W.  Zhang, X.M.  Ma, Y.H.  Chen, L.  Zhang, H.L.  Cai, Z.M.  Wang, R.G.  Xiong, and S.  Gao, Coexistence of magnetic and electric orderings in the metal–formate frameworks of  [NH4][M(HCOO)3], J. Am. Chem. Soc. 133(38), 14948–14951 (2011).

[22] N. Abhyankar, J.J. Kweon, M. Orio, S. Bertaina, M.  Lee, E.S.  Choi, R.  Fu, and N.S.  Dalal, Understanding ferroelectricity in the  Pb-free perovskite-like metal–organic framework  [(CH3)2NH2]Zn(HCOO)3: Dielectric, 2D NMR, and theoretical studies, J. Phys. Chem. C 121(11), 6314–6322 (2017).

[23] G.C.  Xu, X.M.  Ma, L.  Zhang, Z.M.  Wang, and S.  Gao, Disorder–order ferroelectric transition in the  metal formate framework of  [NH4] [Zn(HCOO)3], J. Am. Chem. Soc. 132(28), 9588–9590 (2010).

[24] Z.  Wang, B.  Zhang, K.  Inoue, H.  Fujiwara, T.  Otsuka, H. Kobayashi, and M. Kurmoo, Occurrence of a rare 49 · 66 structural topology, chirality, and weak ferromagnetism in the [NH4] [MII(HCOO)3] (M=Mn, Co, Ni) frameworks, Inorg. Chem. 46(2), 437–445 (2007).

[25] Z. Zhang, H. Tang, D. Cheng, J. Zhang, Y. Chen, X. Shen, and H. Yu, Strain coupling and dynamic relaxation in multiferroic metal–organic framework [(CH3)2NH2][Mn(HCOO)3] with perovskite structure, Results Phys. 12, 2183–2188 (2019).

[26] M.  Mączka, A.  Gągor, K.  Hermanowicz, A. Sieradzki, L. Macalik, and A. Pikul, Structural, magnetic and phonon properties of Cr(III)-doped perovskite metal formate framework  [(CH3)2NH2][Mn(HCOO)3], J. Solid State Chem. 237, 150–158 (2016).

[27] M. Maçzka, M. Ptak, and L. Macalik, Infrared and Raman studies of phase transitions in metal–organic frameworks of  [(CH3)2NH2][M(HCOO)3] with M=Zn, Fe, Vib. Spectrosc. 71, 98–104 (2014).

[28] B. Pato-Doldán, M. Sánchez-Andújar, L.C. Gómez-Aguirre, S.  Yáñez-Vilar, J.  López-Beceiro, C. Gracia-Fernández, A.A. Haghighirad, F. Ritter, S.  Castro-García, and M.A.  Senaris-Rodriguez, Near room temperature dielectric transition in the perovskite formate framework [(CH3)2NH2] [Mg(HCOO)3], Phys. Chem. Chem. Phys. 14(24), 8498–8501 (2012).

[29] K.D.  Hughey, A.J.  Clune, M.O.  Yokosuk, J.  Li, N.  Abhyankar, X.  Ding, N.S.  Dalal, H.  Xiang, D.  Smirnov, J.  Singleton, and J.L.  Musfeldt, Structure-property relations in multiferroic  [(CH3)2NH2]M(HCOO)3 (M  =  Mn, Co, Ni), Inorg. Chem. 57(18), 11569–11577 (2018).

[30] T.  Besara, P.  Jain, N.S.  Dalal, P.L.  Kuhns, A.P.  Reyes, H.W.  Kroto, and A.K.  Cheetham, Mechanism of the order–disorder phase transition, and glassy behavior in the  metal–organic framework [(CH3)2NH2]Zn(HCOO)3, Proc. Natl. Acad. Sci. 108(17), 6828–6832 (2011).

[31] M. Šimėnas, L. Macalik, K. Aidas, V. Kalendra, D.  Klose, G.  Jeschke, M.  Maçzka, G.  Völkel, J.J. Banys, and A. Pöppl, Pulse EPR and ENDOR study of manganese doped  [(CH3)2NH2] [Zn(HCOO)3] hybrid perovskite framework, J. Phys. Chem. C 121(48), 27225–27232 (2017).

[32] A.  Clune, N.  Harms, K.R.  O’Neal, K.  Hughey, K.A. Smith, D. Obeysekera, J. Haddock, N.S. Dalal, J. Yang, Z. Liu, and J.L. Musfeldt, Developing the pressure-temperature-magnetic field phase diagram of multiferroic [(CH3)2NH2]Mn(HCOO)3, Inorg. Chem. 59(14), 10083–10090 (2020).

[33] M.  Šimėnas, M.  Ptak, A.H.  Khan, L.  Dagys, V. Balevičius, M. Bertmer, G. Völkel, M. Maczka, A.  Pöppl, and J.  Banys, Spectroscopic study of [(CH3)2NH2][Zn(HCOO)3] hybrid perovskite containing different nitrogen isotopes, J. Phys. Chem. C 122(18), 10284–10292 (2018).

[34] R.  Scatena, R.D.  Johnson, P.  Manuel, and P.  Macchi, Formate-mediated magnetic superexchange in the  model hybrid perovskite [(CH3)2NH2]Cu(HCOO)3, J. Mater. Chem. C 8(37), 12840–12847 (2020).

[35] T. Asaji, S. Yoshitake, Y. Ito, and H. Fujimori, Phase transition and cationic motion in the perovskite formate framework [(CH3)2NH2][Mg(HCOO)3], J. Mol. Struct. 1076, 719–723 (2014).

[36] R. Yadav, D. Swain, H.L. Bhat, and S. Elizabeth, Order–disorder phase transition and multiferroic behaviour in a  metal organic framework compound (CH3)2NH2Co(HCOO)3, J. Appl. Phys. 119, 064103 (2016).

[37] P.  Jain, A.  Stroppa, D.  Nabok, A.  Marino, A. Rubano, D. Paparo, M. Matsubara, H. Nakotte, M. Fiebig, S. Picozzi, E.S. Choi, A.K. Cheetham, C.  Draxl, N.S.  Dalal, and V.S.  Zapf, Switchable electric polarization and ferroelectric domains in a  metal–organic-framework, Npj Quantum Mater. 1, 1–6 (2016).

[38] D.W. Fu, W. Zhang, H.L. Cai, Y. Zhang, J.Z. Ge, R.G.  Xiong, S.D.  Huang, and T.  Nakamura, A multiferroic perdeutero metal–organic framework, Angew. Chem. Int. Ed. 50(50), 11947–11951 (2011).

[39] I. Hatta, Experimental study on dielectric relaxation in NaNO2, J. Phys. Soc. Japan 24, 1043–1053 (1968).

[40] G.  Bator and R.  Jakubas, Dielectric dispersion in ferroelectrics  [NH2(CH3)2]3Sb2Cl9 and [NH2(CH3)2]3Sb2Br9, Phys. Status Solidi 147(2), 591–600 (1995).

[41] M.  Mączka, A.  Pietraszko, B.  Macalik, and K. Hermanowicz, Structure, phonon properties, and order–disorder transition in the  metal formate framework of [NH4][Mg(HCOO)3], Inorg. Chem. 53(2), 787–794 (2014).

[42] H.T.  Nguyen, M.T.  Chau, T.B.T.  Phan, A.Y. Milinskiy, and S.V. Baryshnikov, Phase transition and ferroelectricity of composites based on ferroelectric metal–organic framework of [NH4] [Zn(HCOO)3], Ferroelectr. Lett. Sect. 49(1–3), 22–29 (2022).

[43] M. Maczka, P. Kadlubański, P.T.C. Freire, B. Macalik, W.  Paraguassu, K.  Hermanowicz, and J.  Hanuza, Temperature- and pressure-induced phase transitions in the metal formate framework of [ND4][Zn(DCOO)3] and [NH4][Zn(HCOO)3], Inorg. Chem. 53(18), 9615–9624 (2014).

[44] Y.  Ma, J.  Cong, Y.  Chai, L.  Yan, D.  Shang, and Y.  Sun, Large pyroelectric and thermal expansion coefficients in the [(CH3)2NH2]Mn(HCOO)3 metal–organic framework, Appl. Phys. Lett. 111, 042901 (2017).

[45] P.J.  Baker, T.  Lancaster, I.  Franke, W.  Hayes, S.J. Blundell, F.L. Pratt, P. Jain, Z.M. Wang, and M. Kurmoo, Muon spin relaxation investigation of magnetic ordering in the hybrid organic–inorganic perovskites  [(CH3)2NH2]M(HCOO)3 (M=Ni, Co, Mn, Cu), Phys. Rev. B 82, 012407 (2010).

[46] M. Sánchez-Andújar, S. Presedo, S. Yáñez-Vilar, S.  Castro-García, J.  Shamir, and M.A.  Señarís-Rodríguez, Characterization of the order–disorder dielectric transition in the  hybrid organic–inorganic perovskite-like formate Mn(HCOO)3 [(CH3)2NH2], Inorg. Chem. 49(4), 1510–1516 (2010).

[47] M. Sánchez-Andújar, L.C. Gómez-Aguirre, B. Pato Doldán, S. Yáñez-Vilar, R. Artiaga, A.L. Llamas-Saiz, R.S. Manna, F. Schnelle, M. Lang, F. Ritter, A.A. Haghighirad, and M.A. Señarís-Rodríguez, First-order structural transition in the  multiferroic perovskite-like formate  [(CH3)2NH2] [Mn(HCOO)3], CrystEngComm, 16(17), 3558–3566 (2014).

[48] A.V.  Chitnis, H.  Bhatt, M.  Mączka, M.N.  Deo, and N.  Garg, Remarkable resilience of the  formate cage in a multiferroic metal organic framework material: dimethyl ammonium manganese formate (DMAMnF), Dalt. Trans. 47(37), 12993–13005 (2018).

[49] N.  Abhyankar, M.  Lee, M.  Foley, E.S.  Choi, G. Strouse, H.W. Kroto, and N.S. Dalal, Efficient synthesis and tailoring of magnetic and dielectric properties of Pb-free perovskite-like ABX3 metal-organic frameworks, Phys. Status Solidi 10(8), 600–605 (2016).

[50] S.A.  Locicero, C.M.  Averback, U.  Shumnyk, E.S.  Choi, and D.R.  Talham, Particle size effects on the  order–disorder phase transition in  [(CH3)2NH2]Mg(HCOO)3, J. Phys. Chem. C 124(38), 21113–21122 (2020).

[51] Y.  Tian, A.  Stroppa, Y.  Chai, L.  Yan, S.  Wang, P. Barone, S. Picozzi, and Y. Sun, Cross coupling between electric and magnetic orders in a multiferroic metal-organic framework, Sci. Rep. 4, 1–5 (2014).

[52] E. Bousquet, M. Dawber, N. Stucki, C. Lichtensteiger, P. Hermet, S. Gariglio, J.M. Triscone, and P. Ghosez, Improper ferroelectricity in perovskite oxide artificial superlattices, Nature 452, 732–736 (2008).

[53] A.P. Levanyuk and D.G. Sannikov, Improper ferroelectrics, Sov. Phys. Uspekhi 17, 276–282 (1974).

[54] M. Mączka, A. Gągor, M. Ptak, W. Paraguassu, T.A. Da Silva, A. Sieradzki, and A. Pikul, Phase transitions and coexistence of magnetic and electric orders in the methylhydrazinium metal formate frameworks, Chem. Mater. 29(5), 2264–2275 (2017).

[55] M. Šimėnas, A. Ibenskas, A. Stroppa, A. Gągor, M.  Mączka, J.R.  Banys, and E.E.  Tornau, Simulation of structural phase transitions in perovskite methylhydrazinium metal–formate frameworks: coupled Ising and Potts models, J. Phys. Chem. C 123(32), 19912–19919 (2019).

[56] P.  Jain, N.S.  Dalal, B.H.  Toby, H.W.  Kroto, and A.K. Cheetham, Order–disorder antiferroelectric phase transition in a  hybrid inorganic–organic framework with the  perovskite architecture, J.Am. Chem. Soc. 130(32), 10450–10451 (2008).

[57] S.  Sawada, S.  Nomura, and Y.  Asao, Dielectric properties of ferroelectric NaNO2, J. Phys. Soc. Japan 16(11), 2207–2212 (1961).

[58] M.  Ptak, M.  Mączka, A.  Gągor, A.  Sieradzki, A.  Stroppa, D.  Di Sante, J.M.  Perez-Mato, and L.  Macalik, Experimental and theoretical studies of structural phase transition in a novel polar perovskite-like  [C2H5NH3][Na0.5Fe0.5(HCOO)3] formate, Dalt. Trans. 45(6), 2574–2583 (2016).

[59] A.  Sieradzki, S.  Pawlus, S.N.  Tripathy, A.  Gągor, M.  Ptak, M.  Paluch, and M.  Mączka, Dielectric relaxation and anhydrous proton conduction in  [C2H5NH3][Na0.5Fe0.5(HCOO)3] metal–organic frameworks, Dalt. Trans. 46(11), 3681–3687 (2017).

[60] P.  Peksa, J.K.  Zarȩba, M.  Ptak, M.  Mączka, A. Gągor, S. Pawlus, and A. Sieradzki, Revisiting a  perovskite-like copper-formate framework NH4[Cu(HCOO)3]: order–disorder transition influenced by Jahn-Teller distortion and above room-temperature switching of the  nonlinear optical response between two SHG-active states, J. Phys. Chem. C 124(34), 18714–18723 (2020).

[61] M.  Mączka, K.  Szymborska-Małek, A.  Ciupa, and J. Hanuza, Comparative studies of vibrational properties and phase transitions in metal-organic frameworks of [NH4][M(HCOO)3] with M = Mg, Zn, Ni, Fe, Mn, Vib. Spectrosc. 77, 17–24 (2015).

[62] M. Maczka, A. Sieradzki, B. Bondzior, P. Dereń, J. Hanuza, and K. Hermanowicz, Effect of aliovalent doping on the properties of perovskite-like multiferroic formates, J. Mater. Chem. C 3(36), 9337–9345 (2015).

[63] P.  Peksa, A.  Nowok, F.  Formalik, J.K.  Zaręba, J. Trzmiel, A. Gągor, M. Mączka, and A. Sieradzki, More complex than originally thought: revisiting the origins of the relaxation processes in dimethylammonium zinc formate, J. Mater. Chem. C 10(17), 6866–6877 (2022).

[64] Z.  Wang, P.  Jain, K.-Y.  Choi, J.  van Tol, A.K. Cheetham, H.W. Kroto, H.-J. Koo, H. Zhou, J.  Hwang, E.S.  Choi, M.-H.  Whangbo, and N.S.  Dalal, Dimethylammonium copper formate [(CH3)2NH2]Cu(HCOO)3: A metal–organic framework with quasi-one-dimensional antiferromagnetism and magnetostriction, Phys. Rev. B 87, 224406 (2013).

[65] M.  Mązka, A.  Gągor, B.  Macalik, A.  Pikul, M. Ptak, and J. Hanuza, Order–disorder transition and weak ferromagnetism in the  perovskite metal formate frameworks of  [(CH3)2NH2] [M(HCOO)3] and  [(CH3)2ND2][M(HCOO)3] (M  =  Ni, Mn), Inorg. Chem. 53(1), 457–467 (2014).

[66] A. Sieradzki, M. Mączka, M. Simenas, J.K. Zaręba, A.  Gągor, S.  Balciunas, M.  Kinka, A.  Ciupa, M. Nyk, V. Samulionis, J. Banys, M. Paluch, and S. Pawlus, On the origin of ferroelectric structural phases in perovskite-like metal–organic formate, J. Mater. Chem. C 6(35), 9420–9429 (2018).

[67] M.  Šimėnas, S.  Balčiūnas, M.  Trzebiatowska, M.  Ptak, M.  Mączka, G.  Völkel, A.  Pöppl, and J.  Banys, Electron paramagnetic resonance and electric characterization of a  [CH3NH2NH2] [Zn(HCOO)3] perovskite metal formate framework, J. Mater. Chem. C 5(18), 4526–4536 (2017).

[68] M.  Mączka, A.  Ciupa, A.  Gągor, A.  Sieradzki, A. Pikul, B. Macalik, and M. Drozd, Perovskite metal formate framework of  [NH2-CH+-NH2] Mn(HCOO)3]: Phase transition, magnetic, dielectric, and phonon properties, Inorg. Chem. 53(10), 5260–5268 (2014).

[69] M.  Mączka, J.  Janczak, M.  Trzebiatowska, A. Sieradzki, S. Pawlus, and A. Pikul, Synthesis and temperature-dependent studies of a perovskite-like manganese formate framework templated with protonated acetamidine, Dalt. Trans. 46(26), 8476–8485 (2017).

[70] Y. Imai, B. Zhou, Y. Ito, H. Fijimori, A. Kobayashi, Z.M. Wang, and H. Kobayashi, Freezing of ring-puckering molecular motion and giant dielectric anomalies in metal–organic perovskites, Chem. Asian J. 7(12), 2786–2790 (2012).

[71] T.  Asaji, Y.  Ito, H.  Fujimori, and B.  Zhou, Ring-puckering motion of azetidinium cations in a metal–organic perovskite [(CH2)3NH2] [M(HCOO)3] (M = Zn, Mg) – A thermal and 1H NMR relaxation study, J. Phys. Chem. C 123(7), 4291–4298 (2019).

[72] R. Shang, S. Chen, B.W. Wang, Z.M. Wang, and S.  Gao, Temperature-induced irreversible phase transition from perovskite to diamond but pressure-driven back-transition in an ammonium copper formate, Angew. Chem. Int. Ed. 55(6), 2097–2100 (2016).

[73] R. Shang, G.C. Xu, Z.M. Wang, and S. Gao, Phase transitions, prominent dielectric anomalies, and negative thermal expansion in three high thermally stable ammonium magnesium-formate frameworks, Chem. Eur. J. 20(4), 1146–1158 (2014).

[74] S.  Chen, R.  Shang, K.L.  Hu, Z.M.  Wang, and S.  Gao,  [NH2NH3][M(HCOO)3] (M  =  Mn2+, Zn2+, Co2+ and Mg2+): structural phase transitions, prominent dielectric anomalies and negative thermal expansion, and magnetic ordering, Inorg. Chem. Front. 1(1), 83–98 (2014).

[75] J. Trzmiel, A. Sieradzki, S. Pawlus, and M. Mączka, Insight into understanding structural relaxation dynamics of [NH2NH3][Mn(HCOO)3] metal-organic formate, Mater. Sci. Eng. B 236–237, 24–31 (2018).

[76] M. Mączka, K. Pasińska, M. Ptak, W. Paraguassu, T.A. da Silva, A. Sieradzki, and A. Pikul, Effect of solvent, temperature and pressure on the stability of chiral and perovskite metal formate frameworks of [NH2NH3][M(HCOO)3] (M = Mn, Fe, Zn), Phys. Chem. Chem. Phys. 18(46), 31653–31663 (2016).

[77] M. Mączka, N.L. Marinho Costa, A. Gągor, W. Paraguassu, A. Sieradzki, and J. Hanuza, Structural, thermal, dielectric and phonon properties of perovskite-like imidazolium magnesium formate, Phys. Chem. Chem. Phys. 18(20), 13993–14000 (2016).

[78] S. Horiuchi, F. Kagawa, K. Hatahara, K. Kobayashi, R. Kumai, Y. Murakami, and Y. Tokura, Above-room-temperature ferroelectricity and antiferroelectricity in benzimidazoles, Nat. Commun. 3, 1–6 (2012).

METALO–FORMATO DARINIŲ FEROELEKTRIŠKUMO KONTROVERSIJA

P. Peksa, A. Sieradzki

Vroclavo mokslo ir technologijų universiteto Eksperimentinės fizikos departamentas, Vroclavas, Lenkija