A New Playground for Organic–Inorganic Hybrids ... - ACS Publications

University of A Coruna, QuiMolMat Group, Department of Chemistry, Faculty of Science and Advanced Scientific Research Center (CICA), Zapateira, 15071 ...
1 downloads 8 Views 835KB Size
Viewpoint pubs.acs.org/JPCL

A New Playground for Organic−Inorganic Hybrids: Barocaloric Materials for Pressure-Induced Solid-State Cooling

I

ligands can display a linear N1−5-coordination when M = Au+ or Ag+,24 and a V-shaped N1−5-coordination when M = N.25−32 The resulting compounds also exhibit functional and multifunctional properties (such as ferromagnetism, dielectric and optical properties, colossal thermal expansion, etc.) in addition to multistimuli responsiveness.25−32 In this context, we have very recently reported that the hybrid perovskite [TPrA][Mn(dca)3] (TPrA = (CH3CH2CH2)4N+, dca = [N(CN)2]−) is the very first example of an organic−inorganic hybrid with a giant barocaloric effect (that is, a giant isothermal entropy change driven by applied pressure) near room temperature and under easily accessible pressure (P < 70 bar).32 This finding opens up a new bright horizon for the family of hybrid perovskites, since materials with large solid-state caloric effects induced by external stimuli (mechanic, magnetic and/or electric field) are urgently needed for the development of a new generation of eco-friendly solid-state refrigeration technologies.33,34 Such mechano-, magneto- and electrocaloric materials are called to replace the conventional cooling systems based on the vapor compression of greenhouse gases such as hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs) gases, which will be forbidden in Europe in 2020 (EU Regulation No 517/2014). To date, the scarce barocaloric effects described in the literature have been experimentally observed in expensive rareearth alloys and in some inorganic salts.33,34 And still very few published works have been devoted to computational estimations of caloric effects,35,36 including a recent estimation of the adiabatic temperature change of the (CH3NH3)PbI3 by molecular dynamics.37 Taking into account that the experimentally observed giant barocaloric effect of [TPrA][Mn(dca)3] (ref 32) and the computationally estimated adiabatic temperature change of (CH3NH3)PbI3 (ref 37) are probably not isolated examples of such extraordinary caloric behaviors, but only the tip of an iceberg, we propose to explore the relatively new family of organic−inorganic compounds in the search for other barocaloric materials. For this purpose, we start by identifying the main requirements that materials should fulfill to display large barocaloric effects, which can be of interest for cooling applications. Subsequently, we use this approach as an easy guiding tool to identify potential outstanding candidates within the particular case of hybrid organic−inorganic perovskites. This methodology is also applicable for searching in any other series. Out of the rather scarce barocaloric studies carried out on classic inorganic materials33,34 and our recent research on

n recent years, organic−inorganic hybrid perovskites have attracted increased interest among scientist working in diverse areas, such as solid-state chemistry, condensed matter physics, and materials science.1−4 This family of materials exhibits a wide chemical diversity based on the general formula ABX3, and the perovskite structure, where the A-site can be midsized protonated amines (CH 3 NH 3 +, (CH 3 ) 2 NH 2 + , (CH3)4N+, etc.), the B-site can be different divalent transition metal cations, and the X-site can be different halides or bidentate-bridge ligands (HCOO−, N3−, N(CN)2−, etc.) (see Figure 1).1−4

Figure 1. General ABX3 organic−inorganic hybrid perovskite structure, where the A-site (yellow sphere) is alkylammonium cations, B-site (pink spheres) is transition metal cations, and X-site (green rods) is halides or polyatomic bidentate-bridge ligands.

The presence in these compounds of organic/inorganic polyatomic building-blocks enhance their chemical diversity, structural richness and flexibility, giving rise to novel multifunctional properties. This is the case, for example, of (CH3NH3)PbI3 and related materials, which present remarkable photoconductivity and very large dielectric permittivity, among others.5−10 As for compounds with polyatomic linkers, probably the most studied family is that of the so-called formate-perovskites, [AmineH][M(HCOO)3].11−18 This family, which contains the V-shaped bridging formate ligand (HCOO−), displays very rich and uncommon structural features and phase transitions. These compounds also display very interesting properties such as anomalous thermal expansion, electric and/or magnetic order, which can give rise to type-I multiferroicity, and even magnetically induced polarization (type-II multiferroicity) in the case of [CH3NH3][Co(HCOO)3].18 Another widely studied family is that of azide-perovskites, which contain the linear azide-linker, N3−. These compounds also display interesting functional properties associated with ferroic structural transitions.19−23 As a representative example, the [(CH3)4N][Mn(N3)3] exhibits a multiferroic behavior with coexistence of magnetic, electric, and elastic order.21 Also, the M(CN)2− anions have recently attracted increased attention as linkers for hybrid perovskites. These M(CN)2− © 2017 American Chemical Society

Received: July 17, 2017 Accepted: August 31, 2017 Published: September 21, 2017 4419

DOI: 10.1021/acs.jpclett.7b01845 J. Phys. Chem. Lett. 2017, 8, 4419−4423

Viewpoint

The Journal of Physical Chemistry Letters Table 1. Caloric Parameters of Selected Hybrid Perovskitesa material [TPrA][Mn(dca)3] [TPrA][Co(dca)3] [(CH3)4N][Mn(N3)3] [(CH3)4N]2[CrNa(N3)6] [(CH3)4N]2[CrK(N3)6] [(CH3)4N]2[FeNa(N3)6] [(CH3)4N]2[FeK(N3)6] [(NH2NH3)][Mn(HCOO)3] [(NH2NH3)0.92(CH3NH3)0.08] [Mn(HCOO)3] [(NH2NH3)0.85(CH3NH3)0.15] [Mn(HCOO)3] [(NH2NH3)0.74(CH3NH3)0.26] [Mn(HCOO)3] [(CH3)2NH2][Mn(HCOO)3] [(CH3)2NH2][Mg(HCOO)3] [(CH3)2NH2][Co(HCOO)3] (CH3NH3)PbI3 (CH3NH3)PbBr3 (CH3NH3)PbCl3

Tt (K) exp.

|ΔV| × 10−5 (m3 kg−1) exp.

|ΔS| (J kg−1 K−1) exp.

|δTt/δP| (K kbar−1) calc.

|δTt/δP| (K kbar−1) exp.

ref.

330 340 305 310 313 302 314 355 339

0.9937 0.2482b 1.3381 1.7175 1.7175 1.8645 1.5832 0.1087 0.0925

42.5 15.56 80.77 100.03 93.78 108.59 95.60 48.87 46.64

23.4 16.0 16.6 17.2 18.3 17.2 16.6 2.2 2.0

23.1 18.4 -

28, 32 29 21 22 22 22 22 41 41

322

0.0864

41.72

2.1

-

41

311

0.0777

33.21

2.3

-

41

190 261 156 330 149 171

0.1648 0.0466 0.0902 0.1190 0.2086 0.3043

37.90 22.39 22.49 15.65 23.38 28.93

4.3 2.1 4.0 7.6 8.9c 10.5c

3.5 9.6 5.7

42, 43 44, 46, 46, 46,

45 45 47 48 49

Tt = transition temperature, ΔV = phase transition volume change, ΔS = phase transition entropy change, δTt/δP = barocaloric coefficient, exp. = experimental values, calc. = values calculated by Clausius−Clapeyron. bPhase transition volume change (ΔV) obtained from the here reported synchrotron XRD data, see Figure S1 of SI. cPhase transition volume change (ΔV) obtained from neutron powder diffraction of deuterated samples. a

[TPrA][Mn(dca)3],32 as well as by transposition of studies performed on magneto- and electrocaloric materials,33 we identify the following main requirements that emergent barocaloric materials must meet in order to display giant barocaloric effects of interest for cooling applications: (i) a large and reversible entropy change related to a solid−solid phase transition, an effect that is expected to be especially strong in first-order ferroic transitions, (ii) a large pressure dependence of the transition temperature, δTt/δP, the so-called barocaloric coefficient, and (iii) an adequate transition temperature, Tt, which matches well with that of the desired application. The entropy change of the solid−solid phase transition, which in these hybrids can be made up of configurational, rotational, and/or vibrational contributions,38−40 sets up the maximum barocaloric effect attainable in terms of isothermal entropy change (ΔSit). Meanwhile, the barocaloric coefficient (δTt/δP), which arises from the flexibility of the material and its pressure responsiveness, establishes the operational temperature window in which the barocaloric effect takes place, also known as the temperature span of the material.34 The barocaloric coefficient sign can be positive or negative, giving rise to the so-called conventional and inverse barocaloric materials respectively, which independently from this sign can be used for cooling applications.33 In order to obtain an efficient refrigeration, both the reversible entropy change and barocaloric coefficient should be maximized in absolute values (with independency of the sign). As for Tt, it will depend on the relative stability of the low-temperature versus high-temperature polymorphs involved in the phase transition. In addition to these main requirements, from the technological and commercial point of view, barocaloric materials should be not only eco-friendly and nonhazardous, but also mechanically and thermally stable. Moreover, the synthetic methods to obtain them should be easily scalable and economically accessible. Taking into account that several

members of the fast growing family of hybrid perovskites could already fulfill the requirements to show giant barocaloric effects; and that this family has already exhibited tunable mechanical and thermal stability for different technological applications in addition to low-cost and scalable processing methods,1−4 we devote this work to identify from literature data potential outstanding candidates for barocaloric materials. For this purpose, we have scrutinized the literature in the search for hybrid perovskites that experience large and reversible entropy changes related to a solid−solid phase transition of first-order (which should preferentially take place near room temperature for wider practical applications) and for which thermal and structural data should be available. Subsequently, we have estimated the barocaloric coefficient of the preselected candidates by using the Clausius−Clapeyron eq 1, a widely used indirect method to predict caloric effects in classic caloric materials (normally expensive metals and alloys) with first-order transitions:33,34

∂Tt ΔV = ∂P ΔS

(1)

where ΔS is the phase transition entropy change and ΔV is the relative volume change in the vicinity of the transition temperature, Tt. For ΔS, we have used experimental data reported in the literature that were obtained by differential scanning calorimetry, to calculate the entropy change of their phase transitions in the units commonly used to refer to caloric effects (J kg−1 K−1). As for ΔV, we have calculated the relative volume change from published X-ray diffraction (XRD) data as a function of temperature in the vicinity of the corresponding structural transitions. In the case of the [TPrA][Co(dca)3] compound, the value of ΔV was obtained from here reported synchrotron XRD data (see experimental details and Figure S1 of the Supporting Information (SI)). The results obtained for the 4420

DOI: 10.1021/acs.jpclett.7b01845 J. Phys. Chem. Lett. 2017, 8, 4419−4423

Viewpoint

The Journal of Physical Chemistry Letters

already widely overpass the above-mentioned values, offering great potential as barocaloric materials. In addition, the here shown formate-perovskite families exhibit an entropy change comparable to that of the giant barocaloric [TPrA][Mn(dca)3] material. Nevertheless, their response toward pressure (barocaloric coefficient) is relatively smaller and therefore they would need higher pressures to exhibit useful barocaloric effects. On the other hand, the (CH3NH3)PbCl3 and (CH3NH3)PbBr3 compounds exhibit a similar entropy change and a larger barocaloric coefficient, although their transition temperature (171 and 149 K, respectively) is very low for conventional cooling applications. Meanwhile (CH3NH3)PbI3 displays a relatively smaller entropy change and pressure response even if near room temperature (330 K). Very interestingly, the azide-perovskite family [(CH3)4N][M(N3)3] exhibits an outstanding phase entropy change that is more than twice the value found in the [TPrA][Mn(dca)3] compound. In that regard, it should be noted that, in contrast to formate- and halide-perovskites, the azide-perovskites can display order−disorder processes and off-center shifts of both the A-site cations and the azide ligands, which can also bend and change their coordination mode.21 All these structural features, in addition to the existence of aliovalent M cations (Fe(III)Na(I), Fe(III)K(I), Cr(III)Na(I) and Cr(III)K(I)) in the B-site of the perovskite structure that increase their configurational entropy,38−40 would contribute to such large value of entropy change at the phase transition. Moreover, the estimated pressure response for these azide-peroskites is very large, with a barocaloric coefficient very similar to the [TPrA][M(dca)3] family, which to date displays the largest barocaloric coefficient experimentally observed in hybrid perovskites. Very remarkably both calorimetric parameters, ΔS and δTt/ δP, of this azide-perovskite family are even larger than those observed in the best barocaloric material known so far, the ammonium sulfate with ΔS ∼ 65 J kg−1 K−1 and δTt/δP ∼ 5.7 K kbar−1.55 Finally, it should be noted that, although the barocaloric effect of all these hybrid perovskites could be partially reduced by nonreversible processes or elastic heating contributions, such contributions can be avoided by controlling the rate and magnitude of the applied pressure as in the case of the [TPrA][Mn(dca)3] compound, where the elastic heating contribution is almost negligible for P < 70 bar.32 In summary, we identify the main requirements that materials should fulfill to display large barocaloric effects useful for cooling applications and use them as an easy search tool to detect potential candidates within the family of hybrid perovskite compounds. On this basis, we identify different families of hybrid perovskites as potential emergent barocaloric materials, in particular, the azide-perovskites, which exhibit remarkably large entropy changes related to first-order structural transitions. Moreover, we estimate their barocaloric coefficient by using the simple but robust Clausius−Clapeyron method, which is a widely used indirect method in the case of classic caloric materials. This method reveals that the proposed hybrid perovskite materials exhibit a very large pressure response and, in turn, display a wide operational temperature window for cooling applications, most of them near room temperature. Even more, these examples also show that there is plenty of room to play with the chemistry of these versatile and stimuli responsive compounds to obtain enhanced barocaloric

different families under study are summarized in Table 1, Figure 2, and Table S1 of SI.

Figure 2. Comparison of the barocaloric coefficient (δTt/δP) versus (a) solid−solid phase transition entropy change and (b) transition temperature for different families of organic−inorganic hybrid perovskites. Note: (δTt/δP)calc. are used when (δTt/δP)exp. are not available. The red dash line indicates the magnetic-induced caloric effect of gadolinium in terms of isothermal entropy change (∼11 J kg−1 K−1).33 Inset in panel b: zoom of [(CH3)4N][M(N3)3] for a better view.

As it can be seen in Table 1, we have found more than a dozen hybrid perovskites that satisfy our search requirements. Also and very importantly, the barocaloric coefficient calculated by Clausius−Clapeyron is fully in agreement with the experimental results in those scarce examples where pressure experiments have already been performed, namely, [TPrA][M(dca)3] and (CH3NH3)PbX3 families. The slight differences between the δTt/δP calculated and experimental values for the halide-perovskites are attributable to the reproducibility issues between different batches of those samples, whose crystalline characteristics are well-known to be highly dependent on the synthetic processes and room humidity conditions (among other factors).50,51 Figure 2 shows the barocaloric coefficient (δTt/δP) versus entropy change of the solid−solid phase transition (ΔS), and also versus the transition temperature (Tt) for the different families of hybrid perovskites. As it can be seen there, all of them exhibit an entropy change that is larger than the one observed in gadolinium (Gd), ∼11 J kg1− K−1, which is one of the best magnetocaloric materials for solid-state magnetic refrigeration known today and can be used as reference.33 As for the barocaloric coefficient, most of the reported barocaloric materials to date exhibit a δTt/δP ranging between 1.8 (ref 52) and 6.5 (ref 53) K kbar−1, Gd5Si2Ge2 being one of the most well-known reference barocaloric materials, with δTt/ δP = 3.8 K kbar−1 and ΔS = 21.0 J kg1− K−1.54 As it can be observed in Table 1, most of the described hybrid perovskites 4421

DOI: 10.1021/acs.jpclett.7b01845 J. Phys. Chem. Lett. 2017, 8, 4419−4423

Viewpoint

The Journal of Physical Chemistry Letters materials: finally it is just a question of the right choice and combination of organic/inorganic building blocks, among the innumerable possibilities, to obtain optimized configurational, vibrational, and rotational entropies as well as adequate transition temperatures. In conclusion, we present several hybrid perovskite families with promising barocaloric effects as a starting point to explore already known or novel organic−inorganic hybrids for solidstate cooling applications nonbased on greenhouse gases.

(6) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643−647. (7) Juarez-Perez, E. J.; Sanchez, R. S.; Badia, L.; Garcia-Belmonte, G.; Kang, Y. S.; Mora-Sero, I.; Bisquert, J. Photoinduced Giant Dielectric Constant in Lead Halide Perovskite Solar Cells. J. Phys. Chem. Lett. 2014, 5, 2390−2394. (8) Bisquert, J. Consolidation and Expansion of Perovskite Solar Cell Research. J. Phys. Chem. Lett. 2016, 7, 775−775. (9) Stoumpos, C. C.; Mao, L.; Malliakas, C. D.; Kanatzidis, M. G. Structure−Band Gap Relationships in Hexagonal Polytypes and LowDimensional Structures of Hybrid Tin Iodide Perovskites. Inorg. Chem. 2017, 56, 56−73. (10) Shi, Z.; Guo, J.; Chen, Y.; Li, Q.; Pan, Y.; Zhang, H.; Xia, Y.; Huang, W. Lead-Free Organic−Inorganic Hybrid Perovskites for Photovoltaic Applications: Recent Advances and Perspectives. Adv. Mater. 2017, 29, 1605005. (11) Jain, P.; Ramachandran, V.; Clark, R. J.; Zhou, H. D.; Toby, B. H.; Dalal, N. S.; Kroto, H. W.; Cheetham, A. K. Multiferroic Behavior Associated with an Order−Disorder Hydrogen Bonding Transition in Metal-Organic Frameworks (MOFs) with the Perovskite ABX3 Architecture. J. Am. Chem. Soc. 2009, 131, 13625−13627. (12) Sánchez-Andújar, M.; Presedo, S.; Yáñez-Vilar, S.; CastroGarcía, S.; Shamir, J.; Señarís-Rodríguez, M. A. Characterization of the Order-Disorder Dielectric Transition in the Hybrid Organic-Inorganic Perovskite-Like Formate Mn(HCOO)3[(CH3)2NH2]. Inorg. Chem. 2010, 49, 1510−1516. (13) Zhang, Z.; Li, W.; Carpenter, M. A.; Howard, C. J.; Cheetham, A. K. Elastic properties and acoustic dissipation associated with a disorder−order ferroelectric transition in a metal-organic framework. CrystEngComm 2015, 17, 370−374. (14) Chen, S.; Shang, R.; Hu, K.-L.; Wang, Z.-M.; Gao, S. [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. 2014, 1, 83−98. (15) Zhou, B.; Imai, Y.; Kobayashi, A.; Wang, Z.-M.; Kobayashi, H. Giant Dielectric Anomaly of a Metal-Organic Perovskite with FourMembered Ring Ammonium Cations. Angew. Chem., Int. Ed. 2011, 50, 11441−11445. (16) Maczka, M.; Gagor, A.; Ptak, M.; Paraguassu, W.; da Silva, T. A.; Sieradzki, A.; Pikul, A. Phase Transitions and Coexistence of Magnetic and Electric Orders in the Methylhydrazinium Metal Formate Frameworks. Chem. Mater. 2017, 29, 2264−2275. (17) Stroppa, A.; Barone, P.; Jain, P.; Perez-Mato, J. M.; Picozzi, S. Hybrid Improper Ferroelectricity in a Multiferroic and Magnetoelectric Metal-Organic Framework. Adv. Mater. 2013, 25, 2284−2290. (18) Gómez-Aguirre, L. C.; Pato-Doldán, B.; Mira, J.; Castro-García, S.; Señarís-Rodríguez, M. A.; Sánchez-Andújar, M.; Singleton, J.; Zapf, V. S. Magnetic Ordering-Induced Multiferroic Behavior in [CH3NH3][Co(HCOO)3] Metal-Organic Framework. J. Am. Chem. Soc. 2016, 138, 1122−1125. (19) Mautner, F. A.; Cortés, R.; Lezama, L.; Rojo, T. [N(CH3)4][Mn(N3)3]: A Compound with a Distorted Perovskite Structure through Azido Ligands. Angew. Chem., Int. Ed. Engl. 1996, 35, 78−80. (20) Zhao, X.-H.; Huang, X.-C.; Zhang, S.-L.; Shao, D.; Wei, H.-Y.; Wang, X.-Y. Cation-Dependent Magnetic Ordering and RoomTemperature Bistability in Azido-Bridged Perovskite-Type Compounds. J. Am. Chem. Soc. 2013, 135, 16006−16009. (21) Gómez-Aguirre, L. C.; Pato-Doldán, B.; Stroppa, A.; Yang, L.M.; Frauenheim, T.; Mira, J.; Yáñez-Vilar, S.; Artiaga, R.; CastroGarcía, S.; Sánchez-Andújar, M.; et al. Coexistence of Three Ferroic Orders in the Multiferroic Compound [(CH3)4N][Mn(N3)3] with Perovskite-Like Structure. Chem. - Eur. J. 2016, 22, 7863−7870. (22) Du, Z.-Y.; Zhao, Y.-P.; He, C.-T.; Wang, B.-Y.; Xue, W.; Zhou, H.-L.; Bai, J.; Huang, B.; Zhang, W.-X.; Chen, X.-M. Structural Transition in the Perovskite-like Bimetallic Azido Coordination Polymers: (NMe4)2[B′·B″(N3)6] (B′ = Cr3+, Fe3+; B″ = Na+, K+). Cryst. Growth Des. 2014, 14, 3903−3909.

Juan M. Bermúdez-García* Manuel Sánchez-Andújar María A. Senã rís-Rodríguez*



University of A Coruna, QuiMolMat Group, Department of Chemistry, Faculty of Science and Advanced Scientific Research Center (CICA), Zapateira, 15071 A Coruna, Spain

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpclett.7b01845. Experimental details, variable-temperature synchrotron XRD data, and raw data used for calculating δTt/δP (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Juan M. Bermúdez-García: 0000-0001-7381-4409 Manuel Sánchez-Andújar: 0000-0002-3441-0994 María A. Señarís-Rodríguez: 0000-0002-0117-6855 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported by the Ministerio de Economiá y Competitividad and EU-FEDER (ENE2014-56237-C4-4-R) and Xunta de Galicia (GRC2014/042). Synchrotron experiments (exp. MA2324) were performed on the ID22 beamline at the European Synchrotron Radiation Facility (ESRF), France. J.M.B.-G. thanks Fundación Barrié for a predoctoral fellowship.



REFERENCES

(1) Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R. H.; Cheetham, A. K. Chemically diverse and multifunctional hybrid organic−inorganic perovskites. Nat. Rev. Mater. 2017, 2, 16099. (2) Xu, W.-J.; Du, Z.-Y.; Zhang, W.-X.; Chen, X.-M. Structural phase transitions in perovskite compounds based on diatomic or multiatomic bridges. CrystEngComm 2016, 18, 7915−7928. (3) Saparov, B.; Mitzi, D. B. Organic−Inorganic Perovskites: Structural Versatility for Functional Materials Design. Chem. Rev. 2016, 116, 4558−4596. (4) Boström, H. L. B.; Hill, J. A.; Goodwin, A. L. Columnar shifts as symmetry-breaking degrees of freedom in molecular perovskites. Phys. Chem. Chem. Phys. 2016, 18, 31881−31894. (5) Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Grätzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316−319. 4422

DOI: 10.1021/acs.jpclett.7b01845 J. Phys. Chem. Lett. 2017, 8, 4419−4423

Viewpoint

The Journal of Physical Chemistry Letters

inorganic perovskite materials. Phys. Rev. B: Condens. Matter Mater. Phys. 2016, 94, 180103. (41) Chen, S.; Shang, R.; Wang, B.-W.; Wang, Z.-M.; Gao, S. An ASite Mixed-Ammonium Solid Solution Perovskite Series of [(NH 2 NH 3 )x(CH3 NH3 ) 1− x][Mn(HCOO) 3 ] (x = 1.00−0.67). Angew. Chem. 2015, 127, 11245−11096. (42) Sánchez-Andújar, M.; Gómez-Aguirre, L. C.; Pato Doldán, B.; Yáñez-Vilar, S.; Artiaga, R.; Llamas-Saiz, A. L.; Manna, R. S.; Schnelle, F.; Lang, M.; Ritter, F.; et al. First-order structural transition in the multiferroic perovskite-like formate [(CH3)2NH2][Mn(HCOO)3]. CrystEngComm 2014, 16, 3558−3566. (43) Pato-Doldán, B.; Sánchez-Andújar, M.; Gómez-Aguirre, L. C.; Yáñez-Vilar, S.; López-Beceiro, J.; Gracia-Fernández, C.; Haghighirad, A. A.; Ritter, F.; Castro-García, S.; Señarís-Rodríguez, M. A. Near room temperature dielectric transition in the perovskite formate framework [(CH3)2NH2][Mg(HCOO)3]. Phys. Chem. Chem. Phys. 2012, 14, 8498−8501. (44) López-Beceiro, J.; Gracia-Fernández, C.; Gómez-Barreiro, S.; Castro-García, S.; Sánchez-Andújar, M.; Artiaga, R. Kinetic Study of the Low Temperature Transformation of Co(HCOO)3[(CH3)2NH2]. J. Phys. Chem. C 2012, 116, 1219−1224. (45) Collings, I. E.; Hill, J. A.; Cairns, A. B.; Cooper, R. I.; Thompson, A. L.; Parker, J. E.; Tang, C. C.; Goodwin, A. L. Compositional dependence of anomalous thermal expansion in perovskite-like ABX3 formates. Dalton Trans. 2016, 45, 4169−4178. (46) Onoda-Yamamuro, N.; Yamamuro, O.; Matsuo, T.; Suga, H. p-T phase relations of CH3NH3PbX3 (X = Cl, Br, I) crystals. J. Phys. Chem. Solids 1992, 53, 277−281. (47) Baikie, T.; Fang, Y.; Kadro, J. M.; Schreyer, M.; Wei, F.; Mhaisalkar, S. G.; Graetzel, M.; White, T. J. Synthesis and crystal chemistry of the hybrid perovskite (CH3NH3)PbI3 for solid-state sensitised solar cell applications. J. Mater. Chem. A 2013, 1, 5628− 5641. (48) Swainson, I. P.; Hammond, R. P.; Soullière, C.; Knop, O.; Massa, W. Phase transitions in the perovskite methylammonium lead bromide, CH3ND3PbBr3. J. Solid State Chem. 2003, 176, 97−104. (49) Chi, L.; Swainson, I.; Cranswick, L.; Her, J.-H.; Stephens, P.; Knop, O. The ordered phase of methylammonium lead chloride CH3ND3PbCl3. J. Solid State Chem. 2005, 178, 1376−1385. (50) Grancini, G.; Marras, S.; Prato, M.; Giannini, C.; Quarti, C.; De Angelis, F.; De Bastiani, M.; Eperon, G. E.; Snaith, H. J.; Manna, L.; et al. The Impact of the Crystallization Processes on the Structural and Optical Properties of Hybrid Perovskite Films for Photovoltaics. J. Phys. Chem. Lett. 2014, 5, 3836−3842. (51) Leguy, A. M. A.; Hu, Y.; Campoy-Quiles, M.; Alonso, M. I.; Weber, O. J.; Azarhoosh, P.; van Schilfgaarde, M.; Weller, M. T.; Bein, T.; Nelson, J.; et al. Reversible Hydration of CH3NH3PbI3 in Films, Single Crystals, and Solar Cells. Chem. Mater. 2015, 27, 3397−3407. (52) Mañosa, Ll.; González-Alonso, D.; Planes, A.; Bonnot, E.; Barrio, M.; Tamarit, J.-Ll; Aksoy, S.; Acet, M. Giant solid-state barocaloric effect in the Ni-Mn-In magnetic shape-memory alloy. Nat. Mater. 2010, 9, 478−481. (53) Matsunami, D.; Fujita, A.; Takenaka, K.; Kano, M. Giant barocaloric effect enhanced by the frustration of the antiferromagnetic phase in Mn3GaN. Nat. Mater. 2015, 14, 73−78. (54) Yuce, S.; Barrio, M.; Emre, B.; Stern-Taulats, E.; Planes, A.; Tamarit, J.-Ll; Mudryk, Y.; Gschneidner, K. A., Jr.; Pecharsky, V. K.; Mañosa, Ll. Barocaloric effect in the magnetocaloric prototype Gd5Si2Ge2. Appl. Phys. Lett. 2012, 101, 071906. (55) Lloveras, P.; Stern-Taulats, E.; Barrio, M.; Tamarit, J.-Ll; Crossley, S.; Li, W.; Pomjakushin, V.; Planes, A.; Mañosa, Ll.; Mathur, N. D.; et al. Giant barocaloric effects at low pressure in ferrielectric ammonium sulphate. Nat. Commun. 2015, 6, 8801.

(23) Du, Z.-Y.; Zhao, Y.-P.; Zhang, W.-X.; Zhou, H.-L.; He, C.-T.; Xue, W.; Wang, B.-Y.; Chen, X.-M. Above-room-temperature ferroelastic phase transition in a perovskite-like compound [N(CH3)4][Cd(N3)3]. Chem. Commun. 2014, 50, 1989−1991. (24) Hill, J. A.; Thompson, A. L.; Goodwin, A. L. Dicyanometallates as Model Extended Frameworks. J. Am. Chem. Soc. 2016, 138, 5886− 5896. (25) Tong, M.-L.; Ru, J.; Wu, Y.-M.; Chen, X.-M.; Chang, H.-C.; Mochizuki, K.; Kitagawa, S. Cation-templated construction of threedimensional α-Po cubic-type [M(dca)3]− networks. Syntheses, structures and magnetic properties of A[M(dca)3](dca = dicyanamide; for A = benzyltributylammonium, M = Mn2+, Co2+; for A = benzyltriethylammonium, M = Mn2+, Fe2+). New J. Chem. 2003, 27, 779−782. (26) Schlueter, J. A.; Manson, J. L.; Hyzer, K. A.; Geiser, U. Spin Canting in the 3D Anionic Dicyanamide Structure (SPh3)Mn(dca)3 (Ph = Phenyl, dca = Dicyanamide). Inorg. Chem. 2004, 43, 4100− 4102. (27) Schlueter, J. A.; Manson, J. L.; Geiser, U. Structural and Magnetic Diversity in Tetraalkylammonium Salts of Anionic M[N(CN)2]3− (M = Mn and Ni) Three-Dimensional Coordination Polymers. Inorg. Chem. 2005, 44, 3194−3202. (28) Bermúdez-García, J. M.; Sánchez-Andújar, M.; Yáñez-Vilar, S.; Castro-García, S.; Artiaga, R.; López-Beceiro, J.; Botana, L.; Alegría, A.; Señarís-Rodríguez, M. A. Role of Temperature and Pressure on the Multisensitive Multiferroic Dicyanamide Framework [TPrA][Mn(dca)3] with Perovskite-like Structure. Inorg. Chem. 2015, 54, 11680−11687. (29) Bermúdez-García, J. M.; Sánchez-Andújar, M.; Yáñez-Vilar, S.; Castro-García, S.; Artiaga, R.; López-Beceiro, J.; Botana, L.; Alegría, A.; Señarís-Rodríguez, M. A. Multiple phase and dielectric transitions on a novel multi-sensitive [TPrA][M(dca)3] (M: Fe2+, Co2+ and Ni2+) hybrid inorganic-organic perovskite family. J. Mater. Chem. C 2016, 4, 4889−4898. (30) Geng, F.-J.; Zhou, L.; Shi, P.-P.; Wang, X.-L.; Zheng, X.; Zhang, Y.; Fu, D.-W.; Ye, Q. Perovskite-type organic-inorganic hybrid NLO switches tuned by guest cations. J. Mater. Chem. C 2017, 5, 1529− 1536. (31) Zhou, L.; Zheng, X.; Shi, P.-P.; Zafar, Z.; Ye, H.-Y.; Fu, D.-W.; Ye, Q. Switchable Nonlinear Optical and Tunable Luminescent Properties Triggered by Multiple Phase Transitions in a PerovskiteLike Compound. Inorg. Chem. 2017, 56, 3238−3244. (32) Bermúdez-García, J. M.; Sánchez-Andújar, M.; Castro-García, S.; López-Beceiro, J.; Artiaga, R.; Señ arís-Rodríguez, M. A. Giant barocaloric effect in the ferroic organic-inorganic hybrid [TPrA][Mn(dca)3] perovskite under easily accessible pressures. Nat. Commun. 2017, 8, 15715. (33) Moya, X.; Kar-Narayan, S.; Mathur, N. D. Caloric materials near ferroic phase transitions. Nat. Mater. 2014, 13, 439−450. (34) Mañosa, Ll.; Planes, A. Materials with Giant Mechanocaloric Effects: Cooling by Strength. Adv. Mater. 2017, 29, 1603607. (35) Cazorla, C.; Errandonea, D. Giant Mechanocaloric Effects in Fluorite-Structured Superionic Materials. Nano Lett. 2016, 16, 3124− 3129. (36) He, W.; Huang, H.; Ma, X. First-principles calculations on elastic and entropy properties in FeRh alloys. Mater. Lett. 2017, 195, 156−158. (37) Liu, S.; Cohen, R. E. Response of Methylammonium Lead Iodide to External Stimuli and Caloric Effects from Molecular Dynamics Simulations. J. Phys. Chem. C 2016, 120, 17274−17281. (38) Butler, K. T.; Walsh, A.; Cheetham, A. K.; Kieslich, G. Organised chaos: entropy in hybrid inorganic-organic systems and other materials. Chem. Sci. 2016, 7, 6316−6324. (39) Kieslich, G.; Kumagai, S.; Butler, K. T.; Okamura, T.; Hendon, C. H.; Sun, S.; Yamashita, M.; Walsh, A.; Cheetham, A. K. Role of entropic effects in controlling the polymorphism in formate ABX3 metal-organic frameworks. Chem. Commun. 2015, 51, 15538−15541. (40) Butler, K. T.; Svane, K.; Kieslich, G.; Cheetham, A. K.; Walsh, A. Microscopic origin of entropy-driven polymorphism in hybrid organic4423

DOI: 10.1021/acs.jpclett.7b01845 J. Phys. Chem. Lett. 2017, 8, 4419−4423