You are on page 1of 4

ChemComm

View Article Online


COMMUNICATION View Journal | View Issue

A large room-temperature entropy change in a


new hybrid ferroelastic with an unconventional
Published on 17 July 2020. Downloaded by Universidad de la Coruna on 5/19/2022 9:40:56 AM.

Cite this: Chem. Commun., 2020,


56, 10054 bond-switching mechanism†
Received 11th June 2020,
Accepted 17th July 2020 Wei-Jian Xu, ab Ying Zeng,a Wei Yuan,a Wei-Xiong Zhang *a and
Xiao-Ming Chen a
DOI: 10.1039/d0cc04092d

rsc.li/chemcomm

We showed a reversible bond-switching ferroelastic phase transi- ferroelastics that display very large barocaloric effects generally
tion at near room temperature in a novel organic–inorganic hybrid, require large reversible entropy changes and large levels of
(Me3NNH2)2[Co(CN)6Na(H2O)], to afford a latent heat of 43.7 kJ kg1 dependence of the transition temperature on pressure, i.e.,
and a large entropy change of 146 J K1 kg1. barocaloric coefficient (dTc/dP).10,11 In this sense, it is funda-
mentally important to search for new sources of ferroelastics to
A large barocaloric effect (isothermal entropy change or adia- promote the exploration of hybrid ferroelastic materials with
batic temperature change upon application or withdrawal of large entropy changes.
external hydrostatic pressure) is expected in the region where Bond-switching structural phase transition involving rever-
ferroelasticity spontaneously emerges during a phase transi- sible generation/cleavage of coordination bonds, with a certain
tion, which makes ferroelastics promising candidates for appli- extent of order–disorder and displacement characteristics, is
cations as solid-state cooling devices to replace present not only of great academic interest but also of technological
refrigeration technologies.1–5 In the past decade, the explora- importance. Such bond-switching transition with changes in
tion of ferroelastics was extended to organic–inorganic hybrid the coordination environments of metal ions would bring
materials, profiting from their low weight, low acoustical drastic symmetry breaking that would facilitate a triggering of
impedance, environmentally benign synthesis and structural the development ferroic-related properties and interconversion
tunability. In particular, the past few years have witnessed of the spin state.12–17 Moreover, the entropy changes resulting
the emergence of numerous hybrid ferroelastics, such as from a first-order bond-switching transition are substantially
[Fe(Cp)2][FeCl4] with a narrow band gap of 1.61 eV,6 a chiral larger than those resulting from conventional phase transitions.
hybrid ferroelastic with seven physical channel switches,7 a For instance, we previously reported the first molecular double-
spin-crossover hybrid ferroelastic displaying domain wall perovskite multiaxial ferroelectric, (Me3NOH)2[KFe(CN)6] (named
motion,8 and (Me3NOH)2[ZnCl4] with a large spontaneous TMC-1), showing a bond-switching transition from a monoclinic
strain of 0.129.9 However, the hybrid ferroelastic displaying a ferroelectric phase to a cubic paraelectric phase.12 Understanding
large barocaloric effect is scarce. The main difficulty lies in the the mechanism of such an unconventional bond switching is thus
understanding that the underlying phase transition mechan- crucial for designing ferroic–caloric materials. Herein, we demon-
isms found in hybrid ferroelastics usually originate from con- strated how to make use of the weaker and switchable coordina-
ventional order–disorder and/or displacive types, which are not tion bonds to yield a reversible bond-switching ferroelastic
involved in breaking of the coordination bonds, and thus transition in a novel organic–inorganic hybrid, (Me3NNH2)2
generally contribute to small entropy changes. Nevertheless, [Co(CN)6Na(H2O)] (TMC-2), and afford a large entropy change
of 146 J K1 kg1 at room temperature. And a large volume
a
MOE Key Laboratory of Bioinorganic and Synthetic Chemistry, School of
change during bond-switching transition gave rise to a large
Chemistry, Sun Yat-Sen University, Guangzhou 510275, China. barocaloric coefficient of 22.2 K kbar1, making TMC-2 of
E-mail: zhangwx6@mail.sysu.edu.cn particular interest for solid-state cooling applications at or near
b
Department of Chemistry & CICECO-Aveiro Institute of Materials, University of room temperature. Such an unconventional bond-switching
Aveiro, 3810-193 Aveiro, Portugal
transition being capable of affording a large entropy change
† Electronic supplementary information (ESI) available: Single-crystal structures,
thermogravimetric analysis, and powder X-ray diffraction patterns. CCDC
in a hybrid ferroelastic opens up a new strategy for the design of
2003376 and 2003377. For ESI and crystallographic data in CIF or other electronic hybrid ferroic–caloric materials for use in next-generation eco-
format see DOI: 10.1039/d0cc04092d friendly solid-state refrigeration technologies.

10054 | Chem. Commun., 2020, 56, 10054--10057 This journal is © The Royal Society of Chemistry 2020
View Article Online

Communication ChemComm
Published on 17 July 2020. Downloaded by Universidad de la Coruna on 5/19/2022 9:40:56 AM.

Fig. 1 (a and d) The asymmetric units of TMC-2 in the (a) FP and (d) PP; symmetry codes: (A) 1 + x, y, z; (B) 1  x, 1/2 + y, 1/2  z; (C) 1  x, 1/2 + y,
1/2  z; (D) 1 + x, y, z. (b and e) Simplified structural illustrations showing perspective views along the respective b-axes of the TMC-2 crystals in the (b) FP
and (e) PP; in each case the linkage of cyano groups is depicted by brown lines. (c and f) A comparison of selected structural units in the (c) FP and (f) PP.

TMC-2 was crystallized from an aqueous solution of Na3 To visualize the ferroelastic transition for TMC-2, variable-
[Co(CN)6] and (Me3NNH2)I in a ratio of 1 : 2 by carrying out a temperature polarization microscopy was performed on an
slow evaporation at room temperature. According to an in situ as-grown single crystal. After the first heating–cooling cycle,
variable-temperature X-ray diffraction analysis, TMC-2 crystal- stripe-like domain structures appeared in the (012) plane,
lized in the monoclinic space group P21/c at 200 K (ferroelastic confirming the occurrence of ferroelastic domains in the FP
phase, FP) (Table S1, ESI†). The asymmetric unit was found to (Fig. S6, ESI†). These stripes became clearer in the vicinity of
consist of two Me3NNH2+ cations, one [Co(CN)6]3 anion, one the phase transition point in a subsequent heating run. As
Na+, and one H2O (Fig. 1a). Here, each Na+ ion was located in a illustrated in Fig. 2b and d, well-defined diagonally aligned
distorted NaN5O octahedron and coordinated with five species: stripes were observed at 295 K, which well matched the predic-
two –NH2 groups from two Me3NNH2+ cations (Na–N10 = tions made by Aizu on the domain boundary orientation in the
3.023(1) Å, Na–N8 = 2.825(1) Å, Table S5, ESI†), three N atoms mmmF2/m species (see ESI† for the details of the deduction of
from the CN ligands (Na–N E 2.4 Å) and one H2O. Each domain orientation).23 Above the Tc, domain walls almost dis-
[Co(CN)6]3 anionic octahedron was observed to be connected appeared (Fig. 2c and e), while diagonal domain structures re-
to three neighboring Na+ ions via three bridging cyano groups, emerged upon further cooling to 280 K in the FP (Fig. 2f and g).
forming a two-dimensional (2D) bimetallic anionic layer in the Such reversible disappearance and appearance of domain walls
ab plane (Fig. 1b and Fig. S6a, ESI†). Both Me3NNH2+ cations provided evidence for the ferroelastic phase transition in TMC-2,
were found to interact with the anionic framework via weak
hydrogen bonds (N–H  N E 3.1 Å, Table S4, ESI†).
Upon heating above 300 K, the crystal structure transformed
into a paraelastic phase (PP) in the orthorhombic space group
Pmmn. Each Na+ became involved in a disordered tetrahedral
coordination geometry (NaN3O), thus resulting in a disordered
2D anionic layer as required by the imposed orthorhombic
symmetry in the PP (Fig. 1d). Interestingly, in the PP, all
Me3NNH2+ cations located in the space between the 2D layers
became coordination-free and showed two-fold disorder about
a mirror plane (Fig. 1e and f). Besides, the hydrogen bonds
between Me3NNH2+ cations and the 2D host layer became weaker
(N–H  N E 3.2 Å, Table S7, ESI†). In short, the observation of
switchable weak Na–N coordination bonds revealed for TMC-2 an
unconventional bond-switching phase transition quite different Fig. 2 (a–g) Evolution of the domain structure of TMC-2 in the heating–
from the common order–disorder and displacive transitions in cooling cycle. (h) The crystal morphology of TMC-2 (size: 0.6  0.4 
other hybrid ferroic-materials and plastic crystals.18–22 0.1 mm).

This journal is © The Royal Society of Chemistry 2020 Chem. Commun., 2020, 56, 10054--10057 | 10055
View Article Online

ChemComm Communication

consistent with the symmetry reduction derived from the X-ray indirect Clausius–Clapeyron equation dTc/dP = DV/DS, where
diffraction structural analysis. The total spontaneous strain (ess) DS is the aforementioned entropy change, and DV is the volume
was estimated to be 0.036, comparable to those of cadmium change calculated from in situ single-crystal X-ray diffraction
thiocyanate hexagonal-perovskite ferroelastics (0.043–0.078),24 but (Table S8, ESI†). As shown in Fig. 3c, this entropy change of TMC-
lower than those of (Me3NOH)2[ZnCl4] (0.129)9 and [Fe(Cp)2] 2 was found to be greater than those, at ambient pressure, of
[FeCl4] (0.109).6 Such spontaneous strain can be mainly ascribed state-of-the-art barocaloric materials such as metal alloy Fe49Rh51
to a lattice distortion during the bond-switching phase transition. (12 J K1 kg1),31 ionic conductor AgI (64 J K1 kg1),32 spin
Temperature-dependent dielectric and differential scanning crossover compound [Fe(C11H9N5)](BF4)2 (86 J K1 kg1),33 and
calorimetry (DSC) measurements were taken to provide further ferroelectric compound (NH4)2SO4 (65 J K1 kg1),34 but smaller
than those of TMC-1 (178 J K1 kg1) and plastic crystal
Published on 17 July 2020. Downloaded by Universidad de la Coruna on 5/19/2022 9:40:56 AM.

insight into the bond-switching transition. The dielectric con-


stant measurements revealed a well-defined dielectric bistabil- (CH3)2C(CH2OH)2 (383 J K1 kg1).28–30 It is worth noting that
ity with a thermal hysteresis of ca. 18 K (Fig. 3a, top). This the entropy change of TMC-1 gradually decreased after several
observation was in good agreement with the DSC observation of cycles, and the plastic crystals having an intrinsic high deform-
a pair of endothermic/exothermic peaks at 300/282 K during ability generally did not rapidly and completely recover after
the heating/cooling cycle, suggesting a first-order phase transi- withdrawing the external pressure. Such irreversible characteris-
tion for TMC-2 (Fig. 3a, bottom). tics are unfavorable for many refrigeration cycles. In this sense, by
Integration of the calorimetric peaks yielded latent heat benefiting from good reversibility of a large entropy change, a
(|Q0|) values of 43.7 kJ kg1 and 41.2 kJ kg1 on the heating moderate phase transition temperature (i.e., near room tempera-
and cooling runs, respectively. These values were much higher ture), and a large barocaloric coefficient, TMC-2 may be consid-
than those of elastocaloric/barocaloric compounds (Fig. 3b), such ered as an excellent barocaloric material candidate for solid-state
as inorganic BaTiO3 (0.13 kJ kg1),25 Mn3GaN (6.29 kJ kg1),3 refrigeration applications in the room-temperature region.
hybrid perovskite (TPrA)Mn(dca)3 (14.0 kJ kg1),26 and X-ray diffraction structural analysis revealed a ratio of the
(DMA)[Mg(HCOO)3] (14.5 kJ kg1),27 but lower than that of numbers of geometrically distinguishable cation orientations
plastic crystal (CH3)2C(CH2OH)2 (110 kJ kg1).28–30 in both FP and PP to be N = 2. Accordingly, if only taking into
After several heating–cooling cycles of DSC measurements, account the contribution from order–disorder transition of
the average entropy change DS around the Tc was estimated to cations, a calculated DS of 14.3 J K1 kg1 could be expected
be 146 J K1 kg1, indicating good reversibility of the entropy based on the Boltzmann equation DS = R ln(N), where R is the
change for TMC-2 (Fig. S8, ESI†). Furthermore, a barocaloric gas constant. However, this calculated DS value is much lower
coefficient (dTc/dP) of 22.2 K kbar1 was calculated by using the than the experimental one estimated from the DSC measurement.

Fig. 3 (a) Temperature dependence of the dielectric constant (e 0 ) of TMC-2 as measured using a pressed-powder pellet at 1 MHz in a heating–cooling
cycle (top) and DSC curves of TMC-2 (bottom). (b) Latent heats of TMC-2 and selected mechanocaloric materials. (c) Solid–solid phase transition entropy
change, transition temperature, and barocaloric coefficient data for state-of-the-art barocaloric materials and TMC-2. (d) A schematic illustration of a
refrigeration cycle of TMC-2 based on a bond-switching ferroelastic transition.

10056 | Chem. Commun., 2020, 56, 10054--10057 This journal is © The Royal Society of Chemistry 2020
View Article Online

Communication ChemComm

Unsurprisingly, the cleavage and re-formation of weak Na–N 9 W. Yuan, Y. Zeng, Y.-Y. Tan, J.-H. Zhou, W.-J. Xu, W.-X. Zhang and
coordination bonds between organic cations and the anionic X.-M. Chen, Chem. Commun., 2019, 55, 8983–8986.
10 J. M. Bermúdez-Garcı́a, M. Sánchez-Andújar and M. A. Señarı́s-
inorganic layer mainly contributed to the entropy change, in Rodrı́guez, J. Phys. Chem. Lett., 2017, 8, 4419–4423.
great contrast to the order–disorder transitions of molecular 11 L. Mañosa, A. Planes and M. Acet, J. Mater. Chem. A, 2013, 1,
configurations reported for the known barocaloric materials. 4925–4936.
12 W.-J. Xu, P.-F. Li, Y.-Y. Tang, W.-X. Zhang, R.-G. Xiong and X.-M.
Therefore, we proposed a refrigeration mechanism based on Chen, J. Am. Chem. Soc., 2017, 139, 6369–6375.
this new bond-switching ferroelastic transition (Fig. 3d). First, 13 P. Szuromi, Science, 2017, 356, 817.
applying sufficient pressure on TMC-2 at above 300 K (at the 14 D. Aguilà, P. Dechambenoit, M. Rouzières, C. Mathonière and
R. Clérac, Chem. Commun., 2017, 53, 11588–11591.
PP) would cause a pressure-induced phase transition from the 15 V. I. Ovcharenko, S. V. Fokin, E. T. Kostina, G. V. Romanenko,
Published on 17 July 2020. Downloaded by Universidad de la Coruna on 5/19/2022 9:40:56 AM.

PP to the FP, i.e., a transition of the organic cations from a A. S. Bogomyakov and E. V. Tretyakov, Inorg. Chem., 2012, 51,
coordination-free state to a coordinatively bonded state, with a 12188–12194.
16 M. C. Bernini, F. Gándara, M. Iglesias, N. Snejko, E. Gutiérrez-
release of a large amount of latent heat. Second, TMC-2 would Puebla, E. V. Brusau, G. E. Narda and M. Á. Monge, Chem. – Eur. J.,
be made to return to its initial temperature by dissipating the 2009, 15, 4896–4905.
heat to the environment, and a heat sink is usually used to 17 P. Guionneau, F. Le Gac, A. Kaiba, J. S. Costa, D. Chasseau and J.-F.
Létard, Chem. Commun., 2007, 3723–3725.
facilitate such a process. Third, pressure on TMC-2 would be 18 Y. Tang, Z. Sun, C. Ji, L. Li, S. Zhang, T. Chen and J. Luo, Cryst.
released in order recover the coordination-free state of the organic Growth Des., 2015, 15, 457–464.
cations and to absorb a large amount of heat. Fourth, TMC-2 19 H.-Y. Ye, J.-Z. Ge, Y.-Y. Tang, P.-F. Li, Y. Zhang, Y.-M. You and R.-G.
Xiong, J. Am. Chem. Soc., 2016, 138, 13175–13178.
would be made to absorb heat from a heat source and return to its 20 Y.-J. Cao, L. Zhou, P.-P. Shi, Q. Ye and D.-W. Fu, Chem. Commun.,
initial temperature, thereby cooling the target heat source. 2019, 55, 8418–8421.
In summary, a new room-temperature hybrid ferroelastic, 21 S.-S. Wang, X.-X. Chen, B. Huang, R.-K. Huang, W.-X. Zhang and X.-M.
Chen, CCS Chem., 2019, 1, 448–454.
i.e., TMC-2, was shown to exhibit a reversible first-order bond- 22 H.-Y. Zhang, Y.-Y. Tang, P.-P. Shi and R.-G. Xiong, Acc. Chem. Res.,
switching phase transition accompanied by a large entropy 2019, 52, 1928–1938.
change of 146 J K1 kg1. Furthermore, electrocaloric and 23 K. Aizu, J. Phys. Soc. Jpn., 1969, 27, 387–396.
24 L. He, L. Zhou, P.-P. Shi, Q. Ye and D.-W. Fu, Chem. Mater., 2019, 31,
barocaloric effects were expected to be accomplished simulta- 10236–10242.
neously in the hybrid ferroelectric with large entropy changes 25 E. Stern-Taulats, P. Lloveras, M. Barrio, E. Defay, M. Egilmez,
for the development of multicaloric materials.35–37 This work is A. Planes, J. L. Tamarit, L. Mañosa, N. D. Mathur and X. Moya,
APL Mater., 2016, 4, 091102.
expected to not only be helpful for deepening our understand- 26 J. M. Bermúdez-Garcı́a, M. Sánchez-Andújar, S. Castro-Garcı́a,
ing of unconventional bond-switching phase transitions in J. López-Beceiro, R. Artiaga and M. A. Señarı́s-Rodrı́guez, Nat.
hybrid materials, but also for inspiring the pursuit of intriguing Commun., 2017, 8, 15715.
27 M. Szafrański, W.-J. Wei, Z.-M. Wang, W. Li and A. Katrusiak, APL
further research into novel ferroic–caloric functional materials Mater., 2018, 6, 100701.
based on the bond-switching mechanism. 28 B. Li, Y. Kawakita, S. Ohira-Kawamura, T. Sugahara, H. Wang,
This work was supported by the NSFC (21722107, 21671202, J. Wang, Y. Chen, S. I. Kawaguchi, S. Kawaguchi, K. Ohara, K. Li,
D. Yu, R. Mole, T. Hattori, T. Kikuchi, S.-I. Yano, Z. Zhang, Z. Zhang,
21805312 and 21821003), and Local Innovative and Research W. Ren, S. Lin, O. Sakata, K. Nakajima and Z. Zhang, Nature, 2019,
Teams Project of Guangdong Pearl River Talents Program 567, 506–510.
(2017BT01C161). 29 P. Lloveras, A. Aznar, M. Barrio, P. Negrier, C. Popescu, A. Planes,
L. Mañosa, E. Stern-Taulats, A. Avramenko, N. D. Mathur, X. Moya
and J. L. Tamarit, Nat. Commun., 2019, 10, 1803.
30 A. Aznar, P. Lloveras, M. Barrio, P. Negrier, A. Planes, L. Mañosa,
Conflicts of interest N. D. Mathur, X. Moya and J.-L. Tamarit, J. Mater. Chem. A, 2020, 8,
639–647.
There are no conflicts to declare. 31 E. Stern-Taulats, A. Planes, P. Lloveras, M. Barrio, J.-L. Tamarit,
S. Pramanick, S. Majumdar, C. Frontera and L. Mañosa, Phys. Rev. B:
Notes and references Condens. Matter Mater. Phys., 2014, 89, 214105.
32 A. Aznar, P. Lloveras, M. Romanini, M. Barrio, J.-L. Tamarit,
1 X. Moya, S. Kar-Narayan and N. D. Mathur, Nat. Mater., 2014, 13, 439–450. C. Cazorla, D. Errandonea, N. D. Mathur, A. Planes, X. Moya and
2 T. Ichiro and K. Sandeman, Phys. Today, 2015, 68, 48–53. L. Mañosa, Nat. Commun., 2017, 8, 1851.
3 D. Matsunami, A. Fujita, K. Takenaka and M. Kano, Nat. Mater., 33 S. P. Vallone, A. N. Tantillo, A. M. dos Santos, J. J. Molaison,
2015, 14, 73–78. R. Kulmaczewski, A. Chapoy, P. Ahmadi, M. A. Halcrow and
4 L. Mañosa and A. Planes, Adv. Mater., 2017, 29, 1603607. K. G. Sandeman, Adv. Mater., 2019, 31, 1807334.
5 B. Lu and J. Liu, Sci. Bull., 2015, 60, 1638. 34 P. Lloveras, E. Stern-Taulats, M. Barrio, J. L. Tamarit, S. Crossley,
6 H.-Y. Zhang, C.-L. Hu, Z.-B. Hu, J.-G. Mao, Y. Song and R.-G. Xiong, W. Li, V. Pomjakushin, A. Planes, L. Mañosa, N. D. Mathur and
J. Am. Chem. Soc., 2020, 142, 3240–3245. X. Moya, Nat. Commun., 2015, 6, 8801.
7 S.-Q. Lu, Z.-X. Zhang, H. Cheng, P.-F. Li, W.-Q. Liao and R.-G. Xiong, 35 H. Meng, B. Li, W. Ren and Z. Zhang, Phys. Lett. A, 2013, 377,
Angew. Chem., Int. Ed., 2020, 59, 9574–9578. 567–571.
8 G. G. Morgan, V. B. Jakobsen, E. Trzop, L. C. Gavin, E. Dobbelaar, 36 Y. Liu, G. Zhang, Q. Li, L. Bellaiche, J. F. Scott, B. Dkhil and
S. Chikara, X. Ding, K. Esien, H. Müller-Bunz, S. Felton, V. S. Zapf, Q. Wang, Phys. Rev. B, 2016, 94, 214113.
E. Collet and M. A. Carpenter, Angew. Chem., Int. Ed., 2020, DOI: 37 E. Stern-Taulats, T. Castán, L. Mañosa, A. Planes, N. D. Mathur and
10.1002/anie.202003041. X. Moya, MRS Bull., 2018, 43, 295–299.

This journal is © The Royal Society of Chemistry 2020 Chem. Commun., 2020, 56, 10054--10057 | 10057

You might also like