237 Original scientific paper Journal of Microelectronics, Electronic Components and Materials Vol. 54, No. 4(2024), 237 – 245 https://doi.org/10.33180/InfMIDEM2024.401 How to cite: S. Merselmiz et al., “Electrocaloric and pyroelectric properties of 0.6Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 –0.4BaTi 0.89 Sn 0.11 O 3 ceramics" , Inf. Midem-J. Micro- electron. Electron. Compon. Mater., Vol. 54, No. 4(2024), pp. 237–245 Electrocaloric and pyroelectric properties of 0.6Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 –0.4BaTi 0.89 Sn 0.11 O 3 ceramics Soukaina Merselmiz 1,2 , Zouhair Hanani 2 , Hana Uršič 2,3 , Uroš Prah 2 , Daoud Mezzane 1,4 , El-houssaine Ablouh 5 , Matjaž Spreitzer 2 , Lahoucine Hajji 1 , Zahra Abkhar 1 , Brigita Rožič 2 , Mimoun El Marssi 4 and Zdravko Kutnjak 2 1 IMAD-Lab, Cadi Ayyad University, Marrakesh, Morocco 2 Jožef Stefan Institute, Ljubljana, Slovenia 3 Jožef Stefan International Postgraduate School, Ljubljana, Slovenia 4 LPMC, University of Picardy Jules Verne, Amiens, France 5 MSN, Mohammed VI Polytechnic University (UM6P), Benguerir, Morocco Abstract: Ferroelectric materials are gaining considerable attention for energy storage, electrocaloric and pyroelectric energy harvesting applications. In particular, Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 (BCZT) and BaTi 0.89 Sn 0.11 O 3 (BTSn) ceramics are among the best-studied lead- free BaTiO 3 -based ferroelectrics with high piezoelectric and electrocaloric properties. In this work, we prepared a 0.6BCZT–0.4BTSn solid solution. The structural, energy storage, electrocaloric, and pyroelectric properties are investigated. An energy density of 61.4 mJ cm -3 with a high energy efficiency of 82.4 % at 90 °C is achieved. The electrocaloric temperature change, which is determined indirectly via the Maxwell relation, is 0.5 K at 86 °C and 25 kV cm -1 . It is stable over a wide temperature range of around 65 °C and has a coefficient of performance of 15. Moreover, a pyroelectric energy density of 124.1 mJ cm -3 is achieved. The results of this study show that the 0.6BCZT–0.4BTSn ceramics is a multifunctional material with energy storage, electrocaloric and pyroelectric properties. Keywords: Lead-free; ceramic; BCZT; energy storage; electrocaloric; pyroelectric; energy harvesting Elektrokalorične in piroelektrične lastnosti 0.6Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 –0.4BaTi 0.89 Sn 0.11 O 3 keramike Izvleček: Feroelektrični materiali pridobivajo veliko pozornost v raziskavah, ki se osredotočajo na elektrokalorične in piroelektrične pojave ter na shranjevanje energije. Zlasti keramiki Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 (BCZT) in BaTi 0.89 Sn 0.11 O 3 (BTSn) sodita med najbolj raziskane keramične materiale brez svinca na osnovi BaTiO 3 . V tem delu smo pripravili trdno raztopino 0.6BCZT–0.4BTSn. Raziskali smo strukturne, elektrokalorične in piroelektrične lastnosti keramike 0.6BCZT-0.4BTSn ter njeno zmožnost shranjevanja energije. Keramika izkazuje gostoto shranjevanja energije v višini 61.4 mJ cm -3 z najvišjim energijskim izkoristkom 82.4 % pri temperaturi 90 °C. Elektrokalorična temperaturna sprememba določena preko Maxwellove enačbe, znaša 0.5 K pri temperature 86 °C in električnem polju 25 kV cm -1 ter je stabilna v širokem temperaturnem območju 65 °C s koeficientom učinkovitosti 15. Keramika izkazuje tudi piroelektrično gostoto energije 124.1 mJ cm -3 . Rezultati kažejo, da je keramika 0.6BCZT-0.4BTSn večfunkcijski material, ki izkazuje elektrokalorične in piroelektrične lastnosti ter zmožnost shranjevanja energije. Ključne besede: keramika brez svinca; BCZT; shranjevanje energije; elektrokalorik; piroelektrik; zbiranje energije * Corresponding Author’s e-mail: soukaina.merselmiz@ijs.si; hana.ursic@ijs.si 238 S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 1 Introduction To alleviate growing environmental concerns, the green energy industry is developing rapidly [1]–[4]. In particular, high-efficiency electrocaloric (EC) cooling technologies have attracted much attention, especially in ferroelectric materials. This is due to their ability to be efficiently driven by electric fields that are readily available, making them promising for use in solid-state cooling systems to cool microelectronic devices [5]–[7]. This is due to their polarization and entropy change near the ferroelectric phase transition upon applica- tion/removal of an electric field, resulting in an adia- batic temperature change (ΔT), known as the EC effect [8]–[10]. In addition, dielectric capacitors such as fer- roelectric materials have been widely used in energy- scavenging technologies based solely on their intrinsic polarization [11], [12]. The waste heat produced by many electronic devices presents an opportunity for energy harvesting tech- nologies, which can convert it in various ways [13]. One approach to enhance device efficiency, involves harvesting and converting this wasted heat through pyroelectric energy harvesting [14], [15]. This method requires converting heat energy into clean electricity using materials exhibiting the pyroelectric effect [16]. This effect known as the converse of the EC effect, in- volves the transformation of waste or heat energy into electrical voltage when subjected to temperature vari- ations [17]. Its magnitude can be assessed using the Olsen cycle, similar to the Ericson cycle [18], [19]. Ac- cordingly, the density of pyroelectric energy harvest- ing (U pyro ) can be calculated from the recorded polariza- tion–electric field (P–E) hysteresis loop of ferroelectric materials. Since ferroelectrics are a subgroup of pyro- electrics, BaTiO 3 (BT)-based materials are considered promising candidates for pyroelectric energy harvest- ing. These materials exhibit significant spontaneous po- larization and can undergo polarization changes across a broad temperature range, fulfilling the requirements of the EC effect. For example, U pyro value of 229 mJ cm -3 was found in 0.5BaZr 0.2 Ti 0.8 O 3 –0.5Ba 0.7 Ca 0.3 TiO 3 ceramics [20]. In addition, a comparable U pyro value of 210 mJ cm -3 was obtained in BaTi 0.91 Sn 0.09 O 3 ceramics [21]. Ceramic dielectric capacitors play crucial roles as en- ergy conversion and storage devices by absorbing and releasing large voltages or current pulses within a short lifetime between microseconds and milliseconds [22]. This property makes them promising candidates for energy-storage devices within pulsed-power and pow- er-conditioning electronic applications [2], [23]. Pure BT ceramics capacitors exhibit a ferroelectric tetrago- nal phase with high dielectric permittivity close to the Curie temperature (T c ) and a relatively square-like P–E hysteresis loop, with both large remanent polarization (P r ) and coercive field (E c ). These properties lead to high energy loss (U loss ), low recovered energy density (U rec ) as well as low energy storage efficiency (η), limiting BT ceramics from practical application in energy storage devices [24]. Doping BT material with Ca 2+ at the A-site and with Zr 4+ /Sn 4+ at the B-site could be beneficial to adjust the P–E hysteresis loop by reducing the P r and increasing the difference ΔP between the maximal po- larization (P max ) and P r , thereby enhancing simultane- ously its U rec and η [22], [24]–[27]. In 2009, Liu et al. reported a high piezoelectric coeffi- cient of d 33 ~ 620 pC N -1 in Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 (abbre- viated as BCZT) ceramics related to the morphotropic phase boundary occurring at room temperature. Sub- sequently, BaTi 0.89 Sn 0.11 O 3 (abbreviated as BTSn) with a quasi-quadruple point (coexistence of cubic–tetrago- nal–orthorhombic–rhombohedral phases) was found to have high dielectric permittivity (~ 75 000) and im- proved piezoelectric coefficient of d 33 ~ 697 pC N -1 at ~ 42 °C [28]. As a result, the chemical modification of BT (e.g., Ca, Zr, Sn, etc.) enhance further the dielectric and piezoelectric poperties [29]–[35]. With the chemi- cal modification, the thermal stability of the properties can be tailored by approaching the rhombohedral–or- thorhombic (R–O, T R–O ) and orthorhombic–tetragonal (O–T, T O–T ) phase boundaries with the corresponding phase transition temperatures to the T c peak tempera- ture together with shifting T c to room temperature [31]. The sequence of phase boundaries enhance the ther- mal stability of the properties over a wide temperature range, which is essential to achieve practical applica- tions [31]. We have previously reported the EC properties of BCZT and BTSn ceramics studied by the indirect Maxwell ap- proach [36], [37]. BTSn ceramics showed high ΔT~0.71 K at 40 °C at 25 kV cm -1 , but in a relatively narrow temper- ature span (T span ) [37]. Meanwhile, BTSn ceramic showed a U rec of 84.4 mJ cm -3 with high η of 91.0 %. In contrast, BCZT ceramics showed a ΔT~0.57 K at 100 °C at the same electric field in a relatively broader T span of 70 K [36]. However, the U rec was ~75 mJ cm -3 with a very low η of 37 %, limiting BCZT from practical applications. In order to prepare multifunctional material with both en- hanced electrical properties and thermal stability, we prepared (1−x)Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 –xBaTi 0.89 Sn 0.11 O 3 solid solution system (x = 0.2, 0.4 and 0.6) as previously reported in our previous work [38]. In this work, we investigated structural, energy storage, EC effect and pyroelectric energy harvesting properties of 0.6BCZT- 0.4BTSn ceramics (abbreviated as 0.4BTSn). 239 2 Materials and methods The 0.6Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 –0.4BaTi 0.89 Sn 0.11 O 3 (abbrevi- ated as 0.4BTSn) ceramics was prepared by conventional solid-state method, by homogenizing BCZT and BTSn calcined powders. The preparation process of 0.4BTSn ceramics is described in detail in our previous work [38]. The crystalline structure of crushed 0.4BTSn ceramic pel- let at room-temperature (RT) was investigated by X-ray powder diffractometer (XRD, BRUKER AXS D4 ENDEAV- OR) equipped with Cu-Kα-radiation. Diffraction patterns were recorded in the 10–80° 2θ-range with a step size of 0.02° using Cu-Kα-radiation. Phase identification was per- formed with the COD-2020 database using the standard diffraction peaks of BaTiO 3 with orthorhombic (PDF#81– 2200) and tetragonal (PDF#05–0626) symmetries [39]. The microstructure of sintered ceramics was examined using a scanning electron microscope (SEM, Zeiss EVO 10 SEM, Carl Zeiss Microscopy, Germany) equipped with an energy dispersive X-ray spectrometer (EDXS, ZEISS SmartEDX Instrument, Carl Zeiss Microscopy, Germany). Prior to the microstructural analysis, the samples were ground and finely polished using a col- loidal silica suspension. The bulk density of the sintered ceramics was determined by the Archimedes’ method using deionized water as medium. In addition, the av- erage grain size was determined from the digitized im- ages of the polished surfaces processed with ImageJ software (version 1.52a, National Institutes of Health, USA) by measuring more than 300 grains using the av- erage grain intercept (AGI) method. For the electrical properties, the ceramic pellets were cut, thinned, and polished to a thickness of about 400 μm and then the Cr/Au electrodes were sputtered on sample’ surfaces. The dielectric properties were meas- ured using a precision LCR Meter instrument (Agilent, 4284A, USA) in the temperature range from −50 to 200 °C. The polarization versus electric field (P–E) hysteresis loops were recorded using Aixacct TF analyzer 2000 (Aixacct, Aachen, Germany) from 30 to 140 °C using a triangular excitation signal with a frequency of 10 Hz. 3 Results The XRD patterns of the 0.4BTSn ceramic at room tem- perature is shown in Figure 1 (a). It shows peaks charac- teristic of the perovskite phase. The XRD fitting pattern extended by 2θ ≈ 45° using the Lorentz fitting method is shown in the inset (Figure 1 (b)). The enlarged peak pre- sents a splitting of two peaks that could be the coexist- ence of orthorhombic (022) O and tetragonal (200) T peaks Figure 2: SEM image and EDXS maps show the distribution of elements Ba, Ca, Zr, Ti, Sn and O on the surface of the 0.4BTSn ceramic (Scale bar: 50 µm). Figure 1: (a) Room-temperature XRD pattern of the 0.4BTSn ceramics, and (b) the enlarged view of the peak splitting at 2θ ≈ 45°. S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 240 forming a (022) O /(200) T doublet [40]. These results were confirmed by using Rietveld refinement, as reported in our previous work [38]. To gain insight into the microstructure and chemical composition of the 0.4BTSn ceramics, Figure 2 shows the SEM and the elemental mapping images on the polished surface of the sample. A compact and dense microstructure with an average grain size of (12.0 ± 4.8) µm were observed. The density of the ceramic was 5.5 g cm -3 , which corresponds to 93 % of the theoretical density. Furthermore, the EDXS mapping images show a homogeneous distribution of all contained elements (Ba, Ca, Zr, Ti, Sn and O). The temperature dependence of the dielectric per- mittivity (ε´) and dielectric loss (tanδ) of the 0.4BTSn sample are shown in Figure 3 (a). Sequential anomalies corresponding to R–O (T R–O ), O–T (T O–T ), and tetragonal- cubic (T c ) phase transitions at about –23, 37, and 75 °C, respectively, are observed. The maximum value of per- mittivity (ε´ max ) and the peak-permittivity temperature (T m ) were found to be ~ 10630 at ~ 77 °C and 1 kHz, cor- responding to a dielectric loss of tanδ ~ 0.04. P–E hysteresis loops at different temperatures are shown in Figure 3 (b). As the temperature increases, the P max decreases continuously due to the ferroelec- tric-paraelectric phase transition above temperatures around ~ 80 °C. To further investigate the energy stor- age properties, the recorded P–E hysteresis loops as a function of applied electric field and temperature were used. Inset in Figure 3 (b) shows schematically the ar- eas presenting the U rec and the U loss in blue and gray colors, respectively. The total energy density (U tot ) can be calculated by integrating and gathering U rec and U loss areas using equations (1) and (2). Therefore, the η can be estimated using equation (3) [41]. The temperature dependence of the energy storage properties is plot- ted in Figure 3 (c). At room temperature, the U rec value was found to be ~ 55 mJ cm -3 with η ~ 65 %, which is twice as high as that of pure BCZT (η ~ 37 %) at 25 kV cm -1 [36]. At 120 °C, high η value of 86 % was found in 0.4BTSn ceramic, exceeding that of pure BCZT (η ~ 72%) [36]. 0 U max P tot EdP   (1) U max r P rec P EdP   (2)  % 100 100 recr ec totr ec loss UU UU U     (3) For environmentally friendly solid-state cooling de- vices, the electrocaloric properties of 0.4BTSn ceram- ics were indirectly evaluated via the Maxwell relation using the measured electric polarization P (T, E). First, a fifth-order polynomial fit of the upper polarization branches was performed at each fixed applied electric field [5]. The thermal evolution of the polarization was Figure 3: Temperature dependence of (a) ε´ and tanδ, and (b) P–E hysteresis loops for 0.4BTSn ceramics. Inset: A schematic depiction of the relevant U rec and U loss de- termined via P–E hysteresis loops. (c) The corresponding energy storage properties as a function of temperature. S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 241 derived. The polarization (P) decreases continuously with increasing temperature as presented in Figure 4 (a). The isothermal entropy change (ΔS) and the ΔT were estimated using Maxwell relation with equations (4) and (5), where E, ρ and C p denote the applied elec- tric field, mass density and specific heat of the sample, respectively [5]. The value of C p (0.48 J g −1 K −1 ) was taken from Ref. [42]. 2 1 E E E P Sd E T        (4) 2 1 E E E p TP Td E CT          (5) Figure 4 (b) shows the temperature dependence of ΔT at different applied electric fields. The T O–T and T c are visible and more pronounced with increasing the applied E. The maximum ΔT was found to be around the T c . As the E increases, ΔT increases and its maxima shift slightly to higher temperatures. At 25 kV cm -1 , ΔT reaches a maximum of 0.5 K at 86 °C, then gradually de- creases. A crucial parameter for evaluating the EC effect of a material is the EC responsivity, written as ζ = ΔT/ ΔE. This calculated coefficient was found to be ζ = 0.20 K mm kV -1 at the peak temperature. Table 1 presents comparable results for some of the previously pub- lished EC outcomes for lead-free ferroelectric materials compared to 0.4BTSn ceramics. For practical cooling applications, maintaining a sig- nificant EC effect over a wide temperature range (T span ) is of great importance. T span is usually specified as the full width at half maximum (FWHM) of the EC peak (at the FE–PE phase transition), which can exceed 45–60 °C at a high EC effect benchmark [43]. The diffuse phase transition has been found to be directly related to the broadened EC peaks at low electric fields [44]. Improved T span value of 65 °C is obtained, which could be explained by the successive phase transitions and to the diffuse phase transition. Another important parameter for evaluating the suit- ability of EC materials for use in solid-state refrigeration systems, is the refrigerant capacity RC = ΔS. T span [5]. This parameter was found to be 33.1 J kg -1 . In addition, the co- efficient of performance (COP = input power/output cool- ing power = |T. ΔS|/|U rec |) is considered a crucial parameter for estimating the refrigeration cycle performance and evaluating the efficiency of the material [5]. The calcu- lated COP value is 15 at 90 °C, which is higher than some other lead-free [45]–[47]. In summary, the 0.4BTSn ceram- ics could be an advantageous material for some specific EC cooling systems in a wide temperature range. Ferroelectric materials with enhanced polarization change upon heating have a high potential for use in pyroelectric energy harvesting [18]. For this reason, the pyroelectric energy harvesting performances of 0.4BTSn ceramic were evaluated. The magnitude of the pyroelectric effect can be evaluated using the Olsen cy- cle. Figure 5 (a) depicts a diagram illustrating the func- tioning of the pyroelectric energy harvesting effect employing the Olsen cycle. It involves two isothermal (AYB, CYD) and two isoelectric (BYC, DYA) process- es per cycle [18]. The pyroelectric energy density (U pyro ), which is achieved in certain temperature and electric field ranges, corresponds to the area AYBYCYDYA, which can be described by equation (6) [48]. The U pyro divided by the total heat energy that is absorbed by this process gives the pyroelectric energy efficiency (η pyro ) (see equation 7). Analogous to the EC effect, we define the pyroelectric responsivity (ζ pyro ) by the equa- tion (8), when the energy harvesting density can be rationalized by the temperature change of the corre- sponding Olsen cycle, Figure 4: Temperature dependence of (a) P and (b) ΔT of the 0.4 BTSn ceramics measured at different applied electric fields from 5 to 25 kV cm -1 , showing the transi- tion temperatures T O–T and T c . S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 242 pyro UE dP  (6)  pyro pyro pH L U CT T     (7) . pyro pyro U ET    (8) Figure 5 (b) shows the high-temperature dependence of U pyro and η pyro at T L = 30 °C, E L = 5 kV cm -1 and E H = 25 kV cm -1 in 0.4BTSn ceramic. It is observed that U pyro increases with T H , and the maximum U pyro value is 124.1 mJ cm -3 at T H = 140 °C. Meanwhile, η pyro increases an- dreaches a maximum value of 0.08 % at T H = 120 °C. Accordingly, ζ pyro is calculated to be 0.56 × 10 −7 J cm -2 V -1 K -1 . The obtained pyroelectric energy harvesting parameters are improved compared to some reported lead-free BaTiO 3 -based ceramics [48], [53], [54]. Accord- ingly, 0.4BTSn ceramic has the potentials to be used as a working material in pyroelectric energy harvesting applications. 4 Conclusions In summary, the multifunctional lead-free 0.4BTSn ce- ramic was prepared by the solid-state reaction method. The energy storage, electrocaloric and pyroelectric en- ergy harvesting properties were systematically inves- tigated. Increased energy storage performances (U rec = 61.4 mJ cm -3 and η = 82.4 % at 90 °C), electrocaloric properties (ΔT = 0.50 K, ζ = 0.20 K mm kV -1 , RC = 33.1 J kg -1 and COP = 15 at T c = 86 °C with T span = 64.9 °C) as well as pyroelectric energy harvesting performances (U pyro = 124.1 mJ cm -3 , η pyro = 0.08 % and ζ pyro = 0.56 × 10 −7 J cm -2 V -1 K -1 at T L = 30 °C and T H = 140 °C) were obtained. These results indicate that the 0.4BTSn sample is a good, eco- friendly, and thermally-stable multifunctional ferro- electric material for energy storage, electrocaloric, and pyroelectric applications. Table 1: Comparison of the electrocaloric properties of 0.4BTSn ceramics with other lead-free BT-based ceramics reported in the literature. Ceramic T c (°C) ΔT (K) ΔE (kV cm -1 ) ζ (K mm V -1 ) Ref. Ba 0.94 Ca 0.06 Ti 0.90 Sn 0.10 O 3 47 0.55 20 0.280 [49] Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 -0.4BaTi 0.89 Sn 0.11 O 3 86 0.50 25 0.200 This work BaTi 0.89 Sn 0.11 O 3 52 0.71 25 0.284 [37] Ba 0.85 Ca 0.15 Zr 0.10 Ti 0.90 O 3 100 0.57 25 0.228 [36] 0.8Ba(Ti 0.89 Sn 0.11 )O 3 –0.2(Ba 0.7 Ca 0.3 )TiO 3 65 0.63 25 0.025 [43] 0.3BaHf 0.2 Ti 0.8 O 3 –0.7Ba 0.94 Sm 0.04 TiO 3 64 0.46 30 0.180 [47] Ba 0.97 Ce 0.03 Ti 0.99 Mn 0.01 O 3 55 0.41 30 0.140 [50] Ba 0.85 Ca 0.15 Zr 0.1 Ti 0.88 Sn 0.02 O 3 80 0.84 32 0.262 [44] Ba 0.7 Sr 0.3 TiO 3 40 0.67 40 0.160 [51] Ba 0.82 Ca 0.18 Sn 0.065 Ti 0.935 O 3 30 0.59 50 0.118 [52] Figure 5: (a) A diagram illustrating the principle of py- roelectric energy harvesting using P–E hysteresis loops measured at two different temperatures based on the Olsen cycle. The green region represents the U pyro . (b) The high-temperature dependence of U pyro and η pyro of 0.4BTSn ceramics between 5 and 25 kV cm -1 . S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 243 5 Acknowledgments The authors would like to thank the CNRST Priority Program PPR 15/2015, the Slovenian research agency grants (No. P2-0105, No. N2-0212 and No. P1-0125, No. J1-9147, No. P2-0091) and the European Union Horizon 2020 Research and Innovation actions MSCA- RISE-ENGIMA (No. 778072) and MSCA-RISE-MELON (No. 872631). 6 Conflict of interest The authors declare no conflict of interest. 7 References 1. B. Malič, M. Otoničar, K. Radan, and J. Koruza, “Lead-Free Piezoelectric Ceramics,” in Encyclope- dia of Materials: Technical Ceramics and Glasses, (Ed.: M. Pomeroy), Ed. Amsterdam: Elsevier BV, 2021, pp. V3-358-V3-368, https://doi.org/10.1016/B978-0-12-803581-8.12131-9. 2. S. Zhang, B. Malič, J. F. Li, and J. Rödel, “Lead-free ferroelectric materials: Prospective applications,” J. Mater. Res., vol. 36, no. 5, pp. 985–995, Mar. 2021, https://doi.org/10.1557/s43578-021-00180-y. 3. J. Rödel, W. Jo, K. T. P . Seifert, E. M. Anton, T. Gran- zow, and D. Damjanovic, “Perspective on the de- velopment of lead-free piezoceramics, ” J. Am. Cer- am. Soc., vol. 92, no. 6, pp. 1153–1177, Jun. 2009, https://doi.org/10.1111/j.1551-2916.2009.03061.x. 4. J. Rödel and J. F. Li, “Lead-free piezoceramics: Sta- tus and perspectives,” MRS Bull., vol. 43, no. 8, pp. 576–580, Aug. 2018, https://doi.org/10.1557/mrs.2018.181. 5. Z. Kutnjak, B. Rožič, and R. Pirc, “Electrocaloric Ef- fect: Theory, Measurements, and Applications,” in Wiley Encyclopedia of Electrical and Electronics En- gineering, Hoboken, NJ, USA: John Wiley & Sons, Inc., 2015, pp. 1–19, https://doi.org/10.1002/047134608x.w8244. 6. X. Moya and N. D. Mathur, “Caloric materials for cooling and heating,” Science (80-. )., vol. 370, no. 6518, pp. 797–803, Nov. 2020, https://doi.org/10.1126/science.abb0973. 7. A. Torelló and E. Defay, “Electrocaloric Coolers: A Review,” Adv. Electron. Mater., vol. 8, no. 6, Jun. 2022, https://doi.org/10.1002/aelm.202101031. 8. Z. Kutnjak and B. Rožič, “Indirect and Direct Meas- urements of the Electrocaloric Effect, ” in Engineer- ing Materials, (Ed.:Tatiana Correia and Qi Zhang), Ed. Berlin, Heidelberg: Springer, 2014, pp. 147– 182, https://doi.org/10.1007/978-3-642-40264-7_7. 9. A. L. Kholkin, O. V. Pakhomov, A. A. Semenov, and A. Tselev, “The Electrocaloric Effect: Materials and Applications, ” Electrocaloric Eff. Mater. Appl., pp. 1–433, 2023, https://doi.org/10.1016/C2019-0-02843-9. 10. X. Chen, V. V. Shvartsman, D. C. Lupascu, and Q. M. Zhang, “Electrocaloric cooling - From materials to devices, ” J. Appl. Phys., vol. 132, no. 24, Dec. 2022, https://doi.org/10.1063/5.0132533. 11. Z. Fan, X. Liu, and X. Tan, “Large electrocaloric re- sponses in [Bi1/2(Na,K)1/2]TiO3-based ceramics with giant electro-strains,” J. Am. Ceram. Soc., vol. 100, no. 5, pp. 2088–2097, May 2017, https://doi.org/10.1111/jace.14777. 12. M. S. Habib et al., “Experimental determination of electrophoretic deposition parameters and elec- trical characterization for K 0.5 Na 0.5 NbO3 per- ovskite thick films for energy harvesting applica- tions, ” Mater. Chem. Phys., vol. 316, p. 129074, Apr. 2024, https://doi.org/10.1016/j.matchemphys.2024.129074. 13. S. Pandya et al., “New approach to waste-heat en- ergy harvesting: pyroelectric energy conversion,” NPG Asia Mater., vol. 11, no. 1, p. 26, Dec. 2019, https://doi.org/10.1038/s41427-019-0125-y. 14. F. Y. Lee, A. Navid, and L. Pilon, “Pyroelectric waste heat energy harvesting using heat conduction,” Appl. Therm. Eng., vol. 37, pp. 30–37, May 2012, https://doi.org/10.1016/j.applthermaleng.2011.12.034. 15. P. Lheritier et al., “Large harvested energy with non-linear pyroelectric modules, ” Nature, vol. 609, no. 7928, pp. 718–721, Sep. 2022, https://doi.org/10.1038/s41586-022-05069-2. 16. D. Zhang, H. Wu, C. R. Bowen, and Y. Yang, “Recent Advances in Pyroelectric Materials and Applica- tions, ” Small, vol. 17, no. 51, Dec. 2021, https://doi.org/10.1002/smll.202103960. 17. S. P. Alpay, J. Mantese, S. Trolier-McKinstry, Q. Zhang, and R. W. Whatmore, “Next-generation electrocaloric and pyroelectric materials for solid- state electrothermal energy interconversion, ” MRS Bull., vol. 39, no. 12, pp. 1099–1111, Dec. 2014, https://doi.org/10.1557/mrs.2014.256. 18. C. R. Bowen, J. Taylor, E. Leboulbar, D. Zabek, A. Chauhan, and R. Vaish, “Pyroelectric materials and devices for energy harvesting applications,” Ener- gy Environ. Sci., vol. 7, no. 12, pp. 3836–3856, 2014, https://doi.org/10.1039/c4ee01759e. 19. S. Patel et al., “Thermomechanical Energy Con- version Potential of Lead-Free 0.50Ba(Zr0.2Ti0.8) O3–0.50(Ba0.7Ca0.3)TiO3 Bulk Ceramics,” Energy Technol., vol. 6, no. 5, pp. 872–882, May 2018, https://doi.org/10.1002/ente.201700416. S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 244 20. D. Ando and K. ichi Kakimoto, “Pyroelectric en- ergy harvesting using low–TC (1–x)(Ba0.7Ca0.3) TiO3–xBa(Zr0.2Ti0.8)O3 bulk ceramics,” J. Am. Ceram. Soc., vol. 101, no. 11, pp. 5061–5070, Nov. 2018, https://doi.org/10.1111/jace.15746. 21. H. Kacem et al., “Relaxor characteristics and py- roelectric energy harvesting performance of BaTi 0.91 Sn 0.09 O 3 ceramic,” J. Alloys Compd., vol. 872, p. 159699, 2021, https://doi.org/10.1016/j.jallcom.2021.159699. 22. F. Yan, J. Qian, S. Wang, and J. Zhai, “Progress and outlook on lead-free ceramics for energy storage applications, ” Nano Energy, vol. 123, p. 109394, May 2024, https://doi.org/10.1016/j.nanoen.2024.109394. 23. L. Yang et al., “Perovskite lead-free dielectrics for energy storage applications,” Prog. Mater. Sci., vol. 102, pp. 72–108, May 2019, https://doi.org/10.1016/j.pmatsci.2018.12.005. 24. H. Zhang et al., “A review on the development of lead-free ferroelectric energy-storage ceramics and multilayer capacitors, ” J. Mater. Chem. C, vol. 8, no. 47, pp. 16648–16667, 2020, https://doi.org/10.1039/d0tc04381h. 25. A. Jain, Y. G. Wang, and L. N. Shi, “Recent devel- opments in BaTiO3 based lead-free materials for energy storage applications, ” J. Alloys Compd., vol. 928, 2022, https://doi.org/10.1016/j.jallcom.2022.167066. 26. V. Veerapandiyan, F. Benes, T. Gindel, and M. De- luca, “Strategies to improve the energy storage properties of perovskite lead-free relaxor ferro- electrics: A review,” Materials (Basel)., vol. 13, no. 24, pp. 1–47, Dec. 2020, https://doi.org/10.3390/ma13245742. 27. V. Veerapandiyan et al., “Origin of Relaxor Behav- ior in Barium-Titanate-Based Lead-Free Perovs- kites, ” Adv. Electron. Mater., vol. 8, no. 2, Feb. 2022, https://doi.org/10.1002/aelm.202100812. 28. Y. Yao et al., “Large piezoelectricity and dielectric permittivity in BaTiO 3-xBaSnO 3 system: The role of phase coexisting, ” Epl, vol. 98, no. 2, 2012, https://doi.org/10.1209/0295-5075/98/27008. 29. T. R. Shrout and S. J. Zhang, “Lead-free piezoelec- tric ceramics: Alternatives for PZT?,” J. Electrocer- amics, vol. 19, no. 1, pp. 111–124, Sep. 2007, https://doi.org/10.1007/s10832-007-9047-0. 30. Y. Zhang, L. Chen, H. Liu, S. Deng, H. Qi, and J. Chen, “High-performance ferroelectric based ma- terials via high-entropy strategy: Design, proper- ties, and mechanism,” InfoMat, vol. 5, no. 12, Dec. 2023, https://doi.org/10.1002/inf2.12488. 31. C. Zhao, B. Wu, and J. Wu, “Composition-driven broad phase boundary for optimizing properties and stability in lead-free barium titanate ceram- ics, ” J. Am. Ceram. Soc., vol. 102, no. 6, pp. 3477– 3487, 2019, https://doi.org/10.1111/jace.16194. 32. M. Acosta et al., “BaTiO3-based piezoelectrics: Fundamentals, current status, and perspectives,” Appl. Phys. Rev., vol. 4, no. 4, Dec. 2017, https://doi.org/10.1063/1.4990046. 33. G. Canu et al., “Structure-property correlations and origin of relaxor behaviour in BaCexTi1-xO3,” Acta Mater., vol. 152, pp. 258–268, Jun. 2018, https://doi.org/10.1016/j.actamat.2018.04.038. 34. S. W. Konsago, A. Debevec, J. Cilenšek, B. Kmet, and B. Malič, “Linear Thermal Expansion of 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3 Bulk Ce- ramic, ” Inf. MIDEM, vol. 53, no. 4, pp. 233–238, Feb. 2023, https://doi.org/10.33180/InfMIDEM2023.403. 35. M. S. Habib et al., “Improved sintering and im- pedance studies of CuO-doped multiferroic (0.98Ba0.85Ca0.15) (Zr0.1Ti0.9)·O30.02BiFeO3 ce- ramics, ” Appl. Phys. A Mater. Sci. Process., vol. 128, no. 3, p. 238, Mar. 2022, https://doi.org/10.1007/s00339-022-05370-x. 36. S. Merselmiz et al., “Thermal-stability of the en- hanced piezoelectric, energy storage and elec- trocaloric properties of a lead-free BCZT ceramic,” RSC Adv., vol. 11, no. 16, pp. 9459–9468, 2021, https://doi.org/10.1039/D0RA09707A. 37. S. Merselmiz et al., “High energy storage effi- ciency and large electrocaloric effect in lead-free BaTi0.89Sn0.11O3 ceramic, ” Ceram. Int., vol. 46, no. 15, pp. 23867–23876, Jun. 2020, https://doi.org/10.1016/j.ceramint.2020.06.163. 38. S. Merselmiz et al., “Design of lead-free BCZT- based ceramics with enhanced piezoelectric en- ergy harvesting performances, ” Phys. Chem. Chem. Phys., 2022, https://doi.org/10.1039/D1CP04723J. 39. L. F. Zhu, B. P. Zhang, L. Zhao, and J. F. Li, “High piezoelectricity of BaTiO3-CaTiO3-BaSnO 3 lead- free ceramics,” J. Mater. Chem. C, vol. 2, no. 24, pp. 4764–4771, 2014, https://doi.org/10.1039/c4tc00155a. 40. W. Cai et al., “Effects of oxygen partial pressure on the electrical properties and phase transitions in (Ba,Ca)(Ti,Zr)O3 ceramics, ” J. Mater. Sci., vol. 55, no. 23, pp. 9972–9992, Aug. 2020, https://doi.org/10.1007/s10853-020-04771-8. 41. H. Palneedi, M. Peddigari, G. T. Hwang, D. Y. Jeong, and J. Ryu, “High-Performance Dielectric Ceramic Films for Energy Storage Capacitors: Progress and Outlook,” Adv. Funct. Mater., vol. 28, no. 42, p. 1803665, Oct. 2018, https://doi.org/10.1002/adfm.201803665. S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 245 42. Y . Zhou, Q. Lin, W. Liu, and D. Wang, “Composition- al dependence of electrocaloric effect in lead-free (1 - X)Ba(Zr0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 ceram- ics, ” RSC Adv., vol. 6, no. 17, pp. 14084–14089, 2016, https://doi.org/10.1039/c5ra26692k. 43. C. Zhao, J. Yang, Y. Huang, X. Hao, and J. Wu, “Broad-temperature-span and large electrocalor- ic effect in lead-free ceramics utilizing successive and metastable phase transitions, ” J. Mater. Chem. A, vol. 7, no. 44, pp. 25526–25536, 2019, https://doi.org/10.1039/c9ta10164k. 44. S. Patel, P. Sharma, and R. Vaish, “Enhanced elec- trocaloric effect in Ba 0.85 Ca 0.15 Zr 0.1 Ti 0.9–x Sn x O 3 ferroelectric ceramics,” Phase Transitions, vol. 89, no. 11, 2016, https://doi.org/10.1080/01411594.2016.1144752. 45. S. Merselmiz et al., “Enhanced electrical proper- ties and large electrocaloric effect in lead-free Ba0.8Ca0.2ZrxTi1−xO3 (x = 0 and 0.02) ceramics,” J. Mater. Sci. Mater. Electron., vol. 31, no. 19, pp. 17018–17028, Oct. 2020, https://doi.org/10.1007/s10854-020-04259-w. 46. L. B. Kong, H. Huang, and S. Li, “Fundamentals of Ferroelectric Materials,” in Ferroelectric Materi- als for Energy Applications, Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, 2018, pp. 1–31, https://doi.org/10.1002/9783527807505.ch1. 47. B. Zhang et al., “Enhanced electrocaloric effect in the Sm and Hf co-doped BaTiO3 ceramics, ” Ceram. Int., vol. 47, no. 1, pp. 1101–1108, Jan. 2021, https://doi.org/10.1016/j.ceramint.2020.08.226. 48. P. Wu et al., “Direct and indirect measurement of electrocaloric effect in lead-free (100-x) Ba(Hf0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 ceramics near multi-phase boundary,” J. Alloys Compd., vol. 725, pp. 275–282, Nov. 2017, https://doi.org/10.1016/j.jallcom.2017.07.103. 49. X. Wang et al., “Giant electrocaloric effect in lead- free Ba 0.94 Ca 0.06 Ti 1− x Sn x O 3 ceramics with tunable Curie temperature,” Appl. Phys. Lett., vol. 107, no. 25, p. 252905, Dec. 2015, https://doi.org/10.1063/1.4938134. 50. S. Liu et al., “Tunable electrocaloric and energy storage behavior in the Ce, Mn hybrid doped Ba- TiO3 ceramics,” J. Eur. Ceram. Soc., vol. 38, no. 14, pp. 4664–4669, Nov. 2018, https://doi.org/10.1016/j.jeurceramsoc.2018.06.020. 51. X. Q. Liu, T. T. Chen, M. Sen Fu, Y. J. Wu, and X. M. Chen, “Electrocaloric effects in spark plasma sin- tered Ba0.7Sr 0.3TiO3-based ceramics: Effects of domain sizes and phase constitution,” Ceram. Int., vol. 40, no. 7 PART B, pp. 11269–11276, Aug. 2014, https://doi.org/10.1016/j.ceramint.2014.03.175. 52. Z. Li, C. Molin, A. Michaelis, and S. E. Gebhardt, “Modified (Ba,Sr)(Sn,Ti)O3 via hydrothermal syn- thesis for electrocaloric application, ” Open Ceram., vol. 16, p. 100502, Dec. 2023, https://doi.org/10.1016/j.oceram.2023.100502. 53. Z. Liu et al., “ Large electrocaloric and pyroelectric energy harvesting effect over a broad tempera- ture range via modulating the relaxor behavior in non-relaxor ferroelectrics ,” J. Mater. Chem. A, vol. 9, no. 38, pp. 22015–22024, 2021, https://doi.org/10.1039/d1ta03894j. 54. Y . Zhao, X. Q. Liu, S. Y . Wu, and X. M. Chen, “Electro- caloric effect and pyroelectric energy harvesting in diffuse ferroelectric Ba(Ti1-xCex)O3 ceramics,” J. Electroceramics, vol. 43, no. 1–4, pp. 106–116, Dec. 2019, https://doi.org/10.1007/s10832-019-00183-6. Arrived: 22. 05. 2024 Accepted: 12. 09. 2024 S. Merselmiz et al.; Informacije Midem, Vol. 54, No. 4(2024), 237 – 245 Copyright © 2024 by the Authors. This is an open access article dis- tributed under the Creative Com- mons Attribution (CC BY) License (https://creativecom- mons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.