Colossal barocaloric effects in the complex hydride Li2B12H12
Traditional refrigeration technologies based on compression cycles of greenhouse gases
pose serious threats to the environment and cannot be downscaled to electronic device dimensions.
Solid-state cooling exploits the thermal response of caloric materials to external fields
and represents a promising alternative to current refrigeration methods. However, most of
the caloric materials known to date present relatively small adiabatic temperature changes
( K) and/or limiting irreversibility issues resulting from significant
phase-transition hysteresis. Here, we predict the existence of colossal barocaloric effects
(isothermal entropy changes of JK-1kg-1) in the energy material
Li2B12H12 by means of molecular dynamics simulations. Specifically, we estimate
JK-1kg-1 and K for an applied pressure of GPa at K. The disclosed colossal barocaloric effects are originated by an
order-disorder phase transformation that exhibits a fair degree of reversibility and involves
coexisting Li+ diffusion and (BH) reorientational motion at high temperatures.
Solid-state cooling is an environmentally friendly, energy efficient, and highly scalable technology that can solve most of the problems associated with conventional refrigeration methods based on compression cycles of greenhouse gases (i.e., environmental harm and lack of downsize scaling). Upon application of magnetic, electric or stress fields good caloric materials undergo noticeable temperature changes (– K) as a result of induced phase transformations that involve large entropy variations (– JK-1kg-1) manosa17 ; moya14 ; cazorla19 . Solid-state cooling capitalizes on such caloric effects to engineer refrigeration cycles. From a performance point of view, that is, largest and (although these are not the only parameters defining cooling efficiency lloveras20 ), barocaloric effects driven by small hydrostatic pressure shifts appear to be the most promising manosa17 ; moya14 ; cazorla19 .
Recently, colossal barocaloric effects (defined here as JK-1kg-1) have been measured in two different families of materials that display intriguing order-disorder phase transitions cazorla17a ; li19 ; lloveras19 . First, giant barocaloric effects have been theoretically predicted cazorla17b and experimentally observed in the archetypal superionic compound AgI cazorla17a . AgI exhibits a first-order normal (low-entropy) to superionic (high-entropy) phase transition that responds to both temperature and pressure sagotra17 and which involves the presence of highly mobile silver ions in the high– superionic state hull04 . The entropy changes estimated for other normal to superionic phase transitions in general are large as well cazorla16 ; cazorla18 ; min20 ; cazorla19a . And second, colossal barocaloric effects have been reported for the molecular solid neopentylglycol li19 ; lloveras19 , (CH3)2C(CH2OH)2, and other plastic crystals lloveras20 . In these solids molecules reorient almost freely around their centers of mass, which remain localized at well-defined lattice positions. Molecular rotations lead to orientational disorder, which renders high entropy. By using hydrostatic pressure, it is possible to block such molecular reorientational motion and thus induce a fully ordered state characterized by low entropy cazorla19a . The barocaloric effects resulting from this class of order-disorder phase transition are huge and comparable in magnitude to those achieved in conventional refrigerators with environmentally harmful fluids lloveras20 ; li19 ; lloveras19 .
Here, we report the prediction of colossal barocaloric effects ( JK-1kg-1) in the energy material Li2B12H12 (LBH), a complex hydride that is already known from the fields of hydrogen storage her08 ; lai19 ; shevlin12 and solid-state batteries paskevicius13 ; luo20 ; mohtadi16 . By using molecular dynamics simulations, we identify a pressure-induced isothermal entropy change of JK-1kg-1 and adiabatic temperature change of K at K. These colossal entropy and temperature changes are driven by moderate hydrostatic pressure shifts of GPa, thus yielding huge barocaloric strengths of JK-1kg-1GPa-1 and K GPa-1. The colossal barocaloric effects disclosed in bulk LBH are originated by simultaneous -driven frustration and activation of Li+ diffusion and (BH) icosahedra reorientational motion. Thus, alkali-metal complex borohydrides (B12H12, Li, Na, K, Cs udovic14 ; udovic20 ) emerge as a promising new family of barocaloric materials in which the salient phase-transition features of fast-ion conductors and plastic crystals coexist.
RESULTS
At ambient conditions, lithium dodecahydrododecaborate (Li2B12H12), LBH, presents an ordered cubic phase, referred to as hereafter, which is characterized by Li+ cations residing on near-trigonal-planar sites surrounded by three (BH) icosahedron anions. In turn, each (BH) anion resides in an octahedral cage surrounded by six Li+ cations (Fig.1a) her08 . A symmetry preserving order-disorder phase transition occurs at high temperatures ( K) that stabilises a disordered state, referred to as hereafter, in which the Li+ cations are mobile and the (BH) anions present reorientational motion (Fig.1b) paskevicius13 . The relative volume expansion that has been experimentally measured for such an order-disorder phase transition is % paskevicius13 . This huge volume variation along with the accompanying, and pressumably also large, phase-transition entropy change could be propitious for barocaloric purposes if the involved phase transformation was responsive to moderate external pressures of GPa. To the best of our knowledge, this possibility has not been hitherto explored. We performed classical molecular dynamics (MD) simulations based on a recently proposed LBH force field sau19 to fill up such a knowledge gap (Methods and Supplementary Methods), which has clear implications for potential solid-state cooling applications.
Figure 2a shows the – phase diagram that we estimated for bulk LBH using atomistic MD simulations. It was found that the temperature of the phase transition is certainly sensitive to external pressure. Specifically, the derivative of the corresponding phase boundary amounts to GPa K-1 at zero pressure and to GPa K-1 at GPa. Likewise, the relative volume change ascribed to the transformation is, according to our simulations, % at zero pressure and % at GPa (Fig.2b). By using these thermodynamic data and the Clausius-Clapeyron relation moya14 , we roughly estimated an entropy change of JK-1kg-1 for the order-disorder transition occurring in LBH at GPa. In view of these promising barocaloric descriptor values, we proceeded to accurately calculate the barocaloric isothermal entropy and adiabatic temperature changes, and , induced by pressures GPa. To this end, we followed the numerical protocols described in the Methods section, which essentially involve the determination of the volume and heat capacity of bulk LBH (Fig.2c) as a function of pressure and temperature.
The results of our precise barocaloric calculations for temperatures and pressures in the intervals K and GPa are shown in Figs. 2d,e. The and values estimated for the transformation in fact render colossal barocaloric effects. For example, at K and GPa ( GPa) we calculated an isothermal entropy change of JK-1kg-1 ( JK-1kg-1) and an adiabatic temperature change of K ( K). The resulting barocaloric effects are direct, that is, , because the low-entropy ordered state is stabilized under pressure (). A maximum value of JK-1kg-1 was found at K and GPa (Fig.2d). For temperatures above K, we estimated noticeably smaller and values (e.g., JK-1kg-1 and K for GPa at K), a trend that we link to some anomalous pressure-induced ionic diffusion (explained below). In the Discussion section, we will compare the barocaloric performance of LBH with those of other well-known barocaloric materials. In what follows, the atomistic mechanisms leading to the extraordinary and results just reported are unravelled.
There are two possible sources of large entropy variation in LBH, one stemming from the Li+ ionic diffusion and the other from the (BH) icosahedra reorientational motion. When hydrostatic pressure is applied on the disordered phase at temperatures below K, both the ionic diffusion and molecular orientational disorder are reduced and thus the crystal entropy diminishes significantly. This conclusion is straightforwardly deduced from the -induced variation of the Li+ diffusion coefficient, , and reorientational (BH) frequency, , shown in Figs.3a,b (Methods). For instance, at K and zero pressure and amount to cm2s-1 and s-1, respectively, whereas at GPa both quantities are practically zero (Fig.3). The two resulting contributions to the system entropy variation are of the same sign and make huge.
Which of these two -induced order-restoring effects is most relevant for the barocaloric performance of bulk LBH? To answer this question, we performed constrained MD simulations in which we forced the Li+ ions to remain localized around their equilibrium positions independently of temperature. This type of artificial condition in principle cannot be imposed in the experiments but can be easily enforced in the atomistic simulations. The values estimated in these constrained MD simulations were roughly half the value of the isothermal entropy changes obtained in the standard MD simulations. Therefore, we may conclude that at temperatures below K the pressure-induced entropy changes stemming from the Li+ ionic diffusion and (BH) icosahedra reorientational motion variations play both an equally important role in the global barocaloric response of LBH.
Figure 3a shows that at K the Li+ diffusion coefficient increases under increasing pressure. For example, at K and zero pressure we estimate cm2s-1 whereas at GPa and the same temperature we obtain cm2s-1. This ionic diffusion behaviour is highly anomalous because hydrostatic compression typically hinders ionic transport sagotra17 ; hull04 ; cazorla16 . On the other hand, the reorientational motion of the (BH) icosahedra behaves quite normally, that is, decreases under pressure cazorla19 ; lloveras20 . For instance, at K and zero pressure we estimate s-1 whereas at GPa and the same temperature we obtain s-1 (Fig.3b). We hypothesize that the anomalous -induced Li+ diffusion behaviour observed in our MD simulations is due to the high anionic reorientational motion, which makes the (BH) centers of mass to fluctuate and partially block the ionic current channels skripov13 . Consistently, when the frequency of the (BH) rotations is reduced by effect of compression the ions can flow more easily throughout the crystal and Li+ transport is enhanced. In this particular – region, the two contributions to the crystal entropy variation stemming from Li+ ionic diffusion and (BH) icosahedra reorientational motion have opposite signs hence decreases significantly. The identified anomalous lithium diffusion behaviour, however, ceases at GPa since beyond that point decreases systematically upon increasing pressure (Supplementary Fig.1).
taulats15a | ||||||||
taulats15 | ||||||||
lloveras15 | ||||||||
bermudez17 | ||||||||
vallone19 | ||||||||
li19 ; lloveras19 | ||||||||
cazorla17a | ||||||||
DISCUSSION
To date, large BC effects have been experimentally measured for a number of shape-memory alloys taulats15a ; taulats15 , polar compounds lloveras15 , organic-inorganic hybrid perovskites bermudez17 ; bermudez17b , fluoride-based materials gorev10 , polymers rodriguez82 , the fast-ion conductor AgI cazorla17a and molecular crystals li19 ; lloveras19 ; vallone19 . In Table I, we compare the barocaloric performance predicted for bulk LBH with those of some representative barocaloric materials manosa17 ; moya14 ; cazorla19 . The isothermal entropy change induced in LBH by a moderate hydrostatic pressure of GPa, JK-1kg-1, is comparable in magnitude to the record that has been recently reported for the plastic crystal neopentylglycol by considering a similar pressure shift, JK-1kg-1 li19 ; lloveras19 . The rest of materials in Table I present isothermal entropy changes that are appreciably smaller, made the exception of the polar crystal (NH4)2SO4 which registers JK-1kg-1. As regards , the clear contestants of LBH are the fast-ion conductor AgI ( K) and again the plastic crystal (CH3)2C(CH2OH)2 ( K). The reason for the smaller value estimated for LBH as compared to that of AgI is the significantly larger heat capacity of the former material, which results from a smaller molecular weight cazorla18 . In terms of the barocaloric strengths defined as and , LBH remains competitive with the best performers. For instance, the organic-inorganic hybrid perovskite [TPrA][Mn(dca)3] displays the largest BSS and BST coefficients of all crystals, JK-1kg-1GPa-1 and K GPa-1, respectively, while for bulk LBH we estimate JK-1kg-1GPa-1 and K GPa-1. Meanwhile, the barocaloric strengths reported for the plastic crystal neopentylglycol are comparable in magnitude to those predicted for LBH, which hints at their common order-disorder phase-transition origin.
As it was mentioned in the Introduction, the magnitude of the and shifts are not the only parameters that define the barocaloric performance of a material. The degree of reversibility of the involved -induced phase transition, for instance, is another important barocaloric descriptor that provides information on the materials efficiency during successive pressure application/removal cycles. Specifically, the hysteresis of the transition makes the materials behaviour to depend on its cycling history and to increase the value of the external field that is required to bring the phase transition to completion lloveras20 . As a consequence, the barocaloric performance of a hysteretic material can be significantly worse than that of its ideal non-hysteretic counterpart. In order to quantify the degree of reversibility associated with the phase transition in LBH, we performed a series of long MD simulations ( ns) in which the pressure (temperature) was kept fixed while the temperature (pressure) was varied steadily first from up to K (from up to GPa) and subsequently from back to K (from back to GPa). The results of such field-changing simulations indicate that the degree of reversibility of the order-disorder phase transition is quite acceptable (Supplementary Fig.2). For instance, by monitoring the variation of the system volume, we found that at zero pressure the difference between the transition temperatures observed during the heating and cooling stages was K (Supplementary Fig.2a). The size of , however, increases noticeably at higher pressures ( K at GPa). Meanwhile, at fixed temperature we found that the hysteresis of the phase transition as driven by pressure was practically null at K ( GPa) and equal to GPa at K (Supplementary Fig.2b).
Arguably the only weakness of bulk LBH in terms of barocaloric potential is that the critical temperature of the order-disorder phase transition is significantly higher than room temperature. However, this practical problem can be efficiently solved by means of doping and alloying strategies. In fact, recently it has been experimentally shown that carbon-doped LBH, LiCB11H12, presents a much lower transition temperature of K tang15 , and that the disordered phase is already stabilized at room temperature in Li(CB9H10)–Li(CB11H12) solid solutions kim19 . Moreover, the type of isosymmetric order-disorder phase transition underlying the exceptional barocaloric behaviour of LBH occurs also in analogous alkali-metal complex hydrides (A2B12H12, A = Na, K, Cs) verdal11 and other earth-abundant and non-toxic materials like KHPO4, NaAlSi3O8 and KNO3 christy95 . Bulk KNO3, for example, displays a staggering volume collapse of % for a room-temperature phase transformation induced by a modest pressure of GPa adams88 , which suggests great barocaloric potential as well.
In conclusion, we have predicted the existence of colossal barocaloric effects rendering isothermal entropy changes of JK-1kg-1 and adiabatic temperature shifts of K in the complex hydride Li2B12H12, which are driven by moderate hydrostatic pressures of GPa. The phase transition underlying such colossal barocaloric effects is remarkable as it combines key ingredients of fast-ion conductors (i.e., ionic diffusion) and molecular crystals (i.e., reorientational motion), materials that individually have been proven to be excellent barocaloric materials. This same type of isosymmetric order-disorder phase transition is likely to occur also in other economically affordable and innocuous compounds (e.g., Cs2B12H12 and KNO3), thus broadening significantly the spectrum of caloric materials with commercial potential for solid-state cooling applications. We believe that our simulation study will stimulate experimental research on this new family of barocaloric materials, namely, alkali-metal complex hydrides, which are already known from other technological disciplines (e.g., hydrogen storage and electrochemical devices) and are routinely synthesized in the laboratory.
METHODS
Classical molecular dynamics simulations. Molecular dynamics (MD) simulations were
performed with the LAMMPS code lammps . The pressure and temperature in the system were kept
fluctuating around a set-point value by using thermostatting and barostatting techniques in which some
dynamic variables are coupled to the particle velocities and simulation box dimensions. The interactions
between atoms were modeled with the harmonic Coulomb-Buckingham interatomic potential reported in work
sau19 , the details of which are provided in the Supplementary Methods. The employed interatomic
potential reproduces satisfactorily the vibrational spectra, structure and lithium diffusion coefficients
of bulk LBH sau19 (Supplementary Discussion). We employed simulation boxes containing atoms
and applied periodic boundary conditions along the three Cartesian directions. Newton’s equations of motion
were integrated using the customary Verlet’s algorithm with a time-step length of fs. The typical
duration of a MD run was of ns. A particle-particle particle-mesh -space solver was
used to compute long-range van der Waals and Coulomb interactions beyond a cut-off distance of Å
at each time step.
Density functional theory calculations. First-principles calculations based on density functional theory (DFT) cazorla17 were performed to analyse the energy, structural, vibrational, and ionic transport properties of Li2B12H12. We performed these calculations with the VASP software vasp by following the generalized gradient approximation to the exchange-correlation energy due to Perdew et al. pbe96 . The projector augmented-wave method was used to represent the ionic cores bloch94 , and the electronic states - Li, -- B and H were considered as valence. Wave functions were represented in a plane-wave basis set truncated at eV. By using these parameters and dense -point grids for Brillouin zone integration, the resulting energies were converged to within meV per formula unit. In the geometry relaxations, a tolerance of eVÅ-1 was imposed on the atomic forces.
Ab initio molecular dynamics (AIMD) simulations based on DFT were carried out to assess the reliability of
the interatomic potential model employed in the classical molecular dynamics simulations (Supplementary Fig.3 and
Supplementary Discussion). The AIMD simulations were performed in the canonical ensemble considering
constant number of particles, volume and temperature. The constrained volumes were equal to the equilibrium volumes
determined at zero temperature, thus we neglected possible thermal expansion effects. Nevertheless, in view of
previous first-principles work cazorla19b , it is reasonable to expect that thermal expansion effects do not
affect significantly the estimation of lithium diffusion coefficients at the considered temperatures. The temperature
in the AIMD simulations was kept fluctuating around a set-point value by using Nose-Hoover thermostats. A large
simulation box containing atoms was employed in all the simulations, and periodic boundary conditions were
applied along the three Cartesian directions. Newton’s equations of motion were integrated by using the customary
Verlet’s algorithm and a time-step length of ps. -point sampling for integration within
the first Brillouin zone was employed in all the AIMD simulations. The AIMD simulations comprised long simulation
times of ps.
Estimation of key quantities. The mean square displacement of lithium ions was estimated with the formula cazorla19b :
where is the position of the migrating ion at time (), represents a lag time, , is the total number of mobile ions, and the total number of time steps. The maximum was chosen equal to , hence we could accumulate enough statistics to reduce significantly the fluctuations in at large ’s. The diffusion coefficient of lithium ions then was obtained with the Einstein relation:
(2) |
by performing linear fits to the averaged values calculated at long .
The angular autocorrelation function of the closoborane (BH) icosahedra was estimated according to the expression sau19 :
(3) |
where is a unitary vector connecting the center of mass of each closoborane unit with one of its edges and denotes thermal average considering all the closoborane icosahedra. This autocorrelation function typically decays as , where the parameter represents a characteristic reorientational frequency. When the (BH) reorientational motion is significant, that is, is large, the function decreases rapidly to zero with time.
Isothermal entropy changes associated with the barocaloric effect were estimated with the formula moya14 ; cazorla19 :
(4) |
where represents the maximum applied hydrostatic pressure and the volume of the system. Likewise, the accompanying adiabatic temperature shift was calculated as:
(5) |
where is the heat capacity of the crystal obtained at constant pressure and temperature conditions.
In order to accurately compute the and shifts induced by pressure, we calculated the corresponding volumes and heat capacities over dense grids of points spaced by GPa and K. Spline interpolations were subsequently applied to the calculated sets of points, which allowed for accurate determination of and heat capacities. The and values appearing in Fig.2d–e were obtained by numerically integrating those spline functions with respect to pressure.
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author (C.C.) upon reasonable request.
ACKNOWLEDGEMENTS
C. C. acknowledges support from the Spanish Ministry of Science, Innovation and Universities under the “Ramón y Cajal” fellowship RYC2018-024947-I. D. E. acknowledges support from the Spanish Ministry of Science, Innovation and Universities under the Grant PID2019-106383GB-C41 and the Generalitat Valenciana under the Grant Prometeo/2018/123 (EFIMAT). Computational resources and technical assistance were provided by the Informatics Service of the University of Valencia through the Tirant III cluster and the Center for Computational Materials Science of the Institute for Materials Research, Tohoku University (MAterial science Supercomputing system for Advanced MUltiscale simulations towards NExt-generation-Institute of Material Research) (Project No-19S0010).
References
- (1) Maosa, Ll. Planes, A. Materials with Giant Mechanocaloric Effects: Cooling by Strength. Adv. Mater. 29, 1603607 (2017).
- (2) Moya, X., Kar-Narayan, S. Mathur, N. D. Caloric materials near ferroic phase transitions. Nat. Mater. 13, 439 (2014).
- (3) Cazorla, C. Novel mechanocaloric materials for solid-state cooling applications. Appl. Phys. Rev. 6, 041316 (2019).
- (4) Aznar, A., Lloveras, P., Barrio, M., Negrier, Ph., Planes, A., Ll. Mañosa, Mathur, N. D., Moya, X. Tamarit, J.-Ll. Reversible and irreversible colossal barocaloric effects in plastic crystals. J. Mater. Chem. A 8, 639 (2020).
- (5) Aznar, A., Lloveras, P., Romanini, M., Barrio, M., Tamarit, J. Ll., Cazorla, C., Errandonea, D., Mathur, N. D., Planes, A., Moya, X. Maosa, Ll. Giant barocaloric effects over a wide temperature range in superionic conductor AgI. Nat. Commun. 8, 1851 (2017).
- (6) Li, B. et al. Colossal barocaloric effects in plastic crystals. Nature 567, 506 (2019).
- (7) Lloveras, P., Aznar, A., Barrio, M., Negrier, Ph., Popescu, C., Planes, A., Ll. Mañosa, Stern-Taulats, E., Avramenko, A., Mathur, N. D., Moya, X. Tamarit, J.-Ll. Colossal barocaloric effects near room temperature in plastic crystals of neopentylglycol. Nat. Commun. 10, 1803 (2019).
- (8) Sagotra, A. K., Errandonea, D. Cazorla, C. Mechanocaloric effects in superionic thin films from atomistic simulations. Nat. Commun. 8, 963 (2017).
- (9) Sagotra, A. K. Cazorla, C. Stress-mediated enhancement of ionic conductivity in fast-ion conductors. ACS Appl. Mater. Interfaces 9, 38773 (2017).
- (10) Hull, S. Superionics: crystal structures and conduction processes. Rep. Prog. Phys. 67, 1233 (2004).
- (11) C. Cazorla Errandonea, D. Giant mechanocaloric effects in fluorite-structured superionic materials. Nano Lett. 16, 3124 (2016).
- (12) Sagotra, A. K., Chu, D. Cazorla C. Room-temperature mechanocaloric effects in lithium-based superionic materials. Nat. Commun. 9, 3337 (2018).
- (13) Min, J., Sagotra, A. K. Cazorla, C. Large barocaloric effects in thermoelectric superionic materials. Phys. Rev. Mater. 4, 015403 (2020).
- (14) Cazorla, C. Refrigeration based on plastic crystals. Nature 567, 470 (2019).
- (15) Her, J.-H., Yousufuddin, M., Zhou, W., Jalisatgi, S. S., Kulleck, J. G., Zan, J. A., Hwang, S.-J., Bowman, R. C. Udovic, T. J. Crystal structure of Li2B12H12: A possible intermediate species in the decomposition of LiBH4. Inorg. Chem. 47, 9757 (2008).
- (16) Lai, Q., Sun, Y., Modi, P., Cazorla, C., Demirci, U. B., Ares, J. R., Leardini, F. Aguey-Zinsou, K. F. How to design hydrogen storage materials? Fundamentals, synthesis, and storage tanks. Adv. Sustainable Syst. 3, 1900043 (2019).
- (17) Shevlin, S. A., Cazorla, C. Guo, Z. X. Structure and defect chemistry of low and high temperature phases of LiBH4. J. Phys. Chem. C 116, 13488 (2012).
- (18) Paskevicius, M., Pitt, M. P., Brown, D. H., Sheppard, D. A., Chumphongphan, S. Buckley, C. E. First-order phase transition in the Li2B12H12 system. Phys. Chem. Chem. Phys. 15, 15825 (2013).
- (19) Luo, X., Rawal, A., Cazorla, C. Aguey-Zinsou, K. F. Facile self-forming superionic conductors based on complex borohydrides surface oxidation. Adv. Sustainable Syst. 4, 1900113 (2020).
- (20) Mohtadi, R. Orimo, S. I. The renaissance of hydrides as energy materials. Nat. Rev. Mater. 2, 16091 (2016).
- (21) Udovic, T. J., Matsuo, M., Unemoto, A., Verdal, N., Stavila, V., Skripov, A. V., Rush, J. J., Takamura, H. Orimo, S.-I. Sodium superionic conduction in Na2B12H12. Chem. Commun. 50, 3750 (2014).
- (22) Jorgensen, M., Shea, P. T., Tomich, A. W., Varley, J. B., Bercx, M., Lovera, S., Cerny, R., Zhou, W., Udovic, T. J., Lavallo, V., Jensen, T. R., Wood, B. C. Stavila, V. Understanding Superionic Conductivity in Lithium and Sodium Salts of Weakly Coordinating Closo-Hexahalocarbaborate Anions. Chem. Mater. 32, 1475 (2020).
- (23) Sau, K., Ikeshoji, T., Kim, S., Takagi, S., Akgi, K. Orimo, S.-i. Reorientational motion and Li+ ion trasnport in Li2B12H12 system: Molecular dynamics study. Phys. Rev. Mater. 3, 075402 (2019).
- (24) Skripov, A. V., Soloninin, A. V., Ley, M. B., Jensen, T. R. Filinchuk, Y. Nuclear magnetic resonance studies of BH4 reorientations and Li diffusion in LiLa(BH4)3Cl. J. Phys. Chem. C 117, 14965 (2013).
- (25) Stern-Taulats, E., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-Ll., Pramanick, S., Majumdar, S., Yüce, S., Emre, B., Frontera, C. Maosa, Ll. Tailoring barocaloric and magnetocaloric properties in low-hysteresis magnetic shape memory alloys. Acta Mater. 96, 324 (2015).
- (26) Stern-Taulats, E., Gracia-Condal, A., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-Ll., Pramanick, S., Majumdar, S. Maosa, Ll. Reversible adiabatic temperature changes at the magnetocaloric and barocaloric effects in Fe49Rh51. App. Phys. Lett. 107, 152409 (2015).
- (27) Lloveras, P., Stern-Taulats, E., Barrio, M., Tamarit, J.-Ll., Crossley, S., Li, W., Pomjakushin, V., Planes, A., Mañosa, Ll., Mathur, N. D. Moya, X. Giant barocaloric effects at low temperature in ferrielectric ammonium sulphate. Nat. Commun. 6, 8801 (2015).
- (28) Bermúdez-García, J. M., Sánchez-Andújar, M., Castro-García, S., López-Beceiro, J., Artiaga, R. Señarís-Rodríguez, M. A. Giant barocaloric effect in the ferroic organic-inorganic hybrid [TPrA][Mn(dca)3] perovskite under easily accessible pressures. Nat. Commun. 8, 15715 (2017).
- (29) Bermúdez-García, J. M., Sánchez-Andújar, M. Señarís-Rodríguez, M. A. A new playground for organic-inorganic hybrids: Barocaloric materials for pressure-induced solid-state cooling. J. Phys. Chem. Lett. 8, 4419 (2017).
- (30) Gorev, M., Bogdanov, E., Flerov, I. N. Laptash, N. M. Thermal expansion, phase diagrams and barocaloric effects in (NH4)2NbOF5. J. Phys.: Condens. Matter 22, 185901 (2010).
- (31) Rodriguez, E, L. Filisko, F. E. Thermoelastic temperature changes in poly(methyl methacrylate) at high hydrostatic pressure: Experimental. J. Appl. Phys. 53, 6536 (1982).
- (32) Vallone, S. P., Tantillo, A. N., dos Santos, A. M., Molaison, J. J., Kulmaczewski, R., Chapoy, A., Ahmadi, P. Halcrow, M. A., Sandeman, K. G. Giant barocaloric effect at the spin crossover transition of a molecular crystal. Adv. Mater. 31, 1807334 (2019).
- (33) Tang, W. S., Unemoto, A., Zhou, W., Stavila, V., Matsuo, M., Wu, H., Orimo, S.-I. Udovic, T. J. Unparalleled lithium and sodium superionic conduction in solid electrolytes with large monovalent cage-like anions. Energy Environ. Sci. 8, 3637 (2015).
- (34) Kim, S., Oguchi, H., Toyama, N., Sato, T., Takagi, S., Otomo, T., Arunkumar, D., Kuwata, N., Kawamura, J. Orimo, S.-I. A complex hydride lithium superionic conductor for high-energy-density all-solid-state lithium metal batteries. Nat. Commun. 10, 1081 (2019).
- (35) Verdal, N., Wu, H., Udovic, T. J., Stavila, V., Zhou, W. Rush, J. J. Evidence of a transition to reorientational disorder in the cubic alkali-metal dodecahydro-closo-dodecaborates. J. Solid State Chem. 184, 3110 (2011).
- (36) Christy, A. G. Isosymmetric structural phase transitions: Phenomenology and examples. Acta Cryst. B51, 753 (1995).
- (37) Adams, D. M., Hatton, P. D., Heath, A. E. Russell, D. R. X-ray diffraction measurements on potassium nitrate under high pressure using synchrotron radiation. J. Phys. C: Solid State Phys. 21, 505 (1988).
- (38) Plimpton, S. J. Fast parallel algorithms for short-range molecular dynamics. J. Comp. Phys. 117, 1 (1995) http://lammps.sandia.gov.
- (39) Cazorla, C. Boronat, J. Simulation and understanding of atomic and molecular quantum crystals. Rev. Mod. Phys. 89, 035003 (2017).
- (40) Kresse, G. Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169 (1996).
- (41) Perdew, J. P., Burke, K. Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).
- (42) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994).
- (43) Sagotra, A. K., Chu, D. Cazorla, C. Influence of lattice dynamics on lithium-ion conductivity: A first-principles study. Phys. Rev. Mater. 3, 035405 (2019).
AUTHOR CONTRIBUTIONS
K.S. and C.C. conceived the study and planned the research. K.S. performed the molecular dynamics
simulations and C.C. the first-principles calculations and barocaloric analysis. Results were
discussed by all the authors. All the authors participated in the writing of the manuscript.
ADDITIONAL INFORMATION
Supplementary information is available in the online version of the paper.
COMPETING INTERESTS
The authors declare no competing interests.