Iván P. Francoa,
Monica Morales-Masis
b,
Iván Mora-Seró
*a and
Rosario Vidal
*a
aInstitute of Advanced Materials (INAM), Universitat Jaume I, Av. Sos Baynat, s/n, Castelló de la Plana 12071, Spain. E-mail: vidal@uji.es
bMESA+ Institute for Nanotechnology, University of Twente, Enschede 7500 AE, The Netherlands
First published on 7th July 2025
Lead zirconate titanate (PZT) is one of the most widely used piezoelectric materials due to its excellent performance. However, its lead content raises serious environmental and health concerns, prompting the search for more sustainable alternatives. In this work, we explore whether a lead-free composite based on the halide perovskite FASnI3 embedded in a polyvinylidene fluoride (PVDF) matrix could serve as a viable substitute for PZT in piezoelectric energy harvesting applications. To assess this potential, we conduct a comparative life cycle assessment (LCA) of both materials in thin-film device configurations, following a cradle-to-grave approach. The analysis includes the environmental impacts of raw material extraction, manufacturing, potential energy recovery during use, end-of-life treatments, and accidental release scenarios. The results show that PZT-based devices have consistently higher environmental impacts across all life cycle stages, mainly due to the high energy requirements for their synthesis and thin-film deposition, as well as the use of lead. In contrast, the FASnI3–PVDF composite benefits from low-temperature processing and the absence of lead, resulting in significantly lower impacts during manufacturing and the use phase. This study offers a first comparative insight into the environmental trade-offs of substituting PZT with halide perovskite-based composites, contributing to the identification of more sustainable piezoelectric solutions.
This technology has been extensively studied as a promising alternative to traditional batteries in specific applications that are exposed to consistent mechanical load. One of the most popular ideas is to implement these devices in roadways. They can be installed in all types of roads, operate independently of weather conditions, be designed in a wide range of sizes and shapes, and have no adverse effects on the lifespan of the pavement.8,9 Piezoelectric energy harvesters (PEHs) offer a sustainable solution to power electronic devices, especially in scenarios where battery replacement is impractical, costly, or environmentally detrimental. This capability is particularly relevant in applications such as structural health monitoring, remote sensors, and wearable electronics,10 where reliable and maintenance-free energy sources are crucial.
As evidenced by their widespread application in numerous electronic devices, piezoelectric materials play an important role in a variety of technologies. The global piezoelectric devices market, valued at USD 29.23 billion in 2021, is projected to grow at a compound annual growth rate (CAGR) of 5.7% over the forecast period. Key drivers of this market expansion include the demand for miniaturized electronic devices such as smartphones, wearables, and IoT products. These applications require compact and efficient components for sensing, actuation, and energy harvesting.11
Lead-zirconate-titanate (PZT) ceramics have long been at the forefront of piezoelectric technology, renowned for their exceptional ferroelectric and piezoelectric properties. These materials exhibit high spontaneous polarization, remarkable dielectric permittivity, and outstanding piezoelectric coefficients, reaching up to 1000 pm V−1 in bulk form and exceeding 300 pm V−1 in thin films.12–14 Introduced by Hans Jaffe in 1954,14 its versatility allows for the synthesis of a wide range of materials with diverse properties by forming solid solutions over a broad range of Zr:
Ti ratios. This adaptability makes PZT a crucial material in numerous applications, offering tailored solutions to meet specific technological needs. PZT is typically synthesized through a solid-state reaction involving the mixing of lead oxide (PbO), zirconium dioxide (ZrO2), and titanium dioxide (TiO2). These precursor powders are thoroughly mixed and then calcined at high temperatures, usually ranging from 800 °C to 900 °C, to promote the formation of the perovskite phase. Following calcination, the resulting powder is milled, pressed into the desired shapes, and subsequently sintered at even higher temperatures over 1000 °C.15 This sintering process enhances the material's densification and crystallinity, which are crucial for achieving the desired piezoelectric properties of PZT. However, it also significantly increases the energy required for its production. PZT is also widely fabricated in thin film form, for which solid targets composed of sintered PbO, ZrO2 and TiO2 as described above, are used as source materials for pulsed laser deposition (PLD) or sputtering deposition of PZT thin films. To achieve high quality films, substrate temperatures above 600 °C are also required.16–18 The main challenge with current PZT materials is their lead content. In the European Union, directives such as the Restriction of Hazardous Substances (RoHS),19 which limits the use of hazardous substances like lead, and the Waste Electrical and Electronic Equipment (WEEE) Directive,20 with its stringent recycling requirements, are driving the search for lead-free alternatives in piezoelectric devices. Although RoHS generally restricts the use of lead, certain exemptions still apply for specific functional materials. Notably, exemption 7(c)-IV, which allowed the use of lead in PZT-based dielectric ceramic materials for capacitors that are part of integrated circuits or discrete semiconductors, expired on 21 July 2024. In addition, exemption 7(c)-I, which permits the use of lead in glass or ceramic materials other than dielectric ceramics, such as in piezoelectric devices, is set to expire on 31 December 2027. This latter exemption has been extended due to the current lack of viable high-performance alternatives, acknowledging that the replacement of PZT remains technologically and economically unfeasible for many applications. These exemptions are subject to periodic review, creating growing regulatory pressure to accelerate the development of sustainable, lead-free piezoelectric materials.21,22
Halide perovskites have recently attracted growing interest as potential piezoelectric materials, offering advantages such as low-temperature processability and possibility to present non-centrosymmetric phases that could give rise to piezoelectric properties. The low temperature and simple processing is a notable contrast to conventional oxide perovskites like PZT, which require synthesis and sintering temperatures above 1000 °C.23–25 Their compatibility with energy-efficient fabrication techniques, such as microwave-assisted synthesis, further highlights their potential by enabling rapid and uniform heating that reduces both processing time and energy consumption.26 These characteristics have positioned halide perovskites as promising candidates for piezoelectric applications in fields ranging from energy harvesting to sensors and actuators.27 Notably, some halide perovskites have demonstrated exceptional piezoelectric performance in single crystal form. In 2017 a hybrid organic-inorganic perovskite, trimethylchloromethylammonium (TMCM)-MnCl3, was reported exhibiting a piezoelectric coefficient (d33) of 185 pC N−1.28 Similarly, in 2020, vacancy-ordered double perovskite incorporating tin, TMCM2SnCl6 achieved a d33 of 137 pC N−1.29 To extend these achievements beyond single crystal formats, recent research has shifted toward the development of halide perovskite-based composites. An effective trade-off between mechanical flexibility and functional performance has been demonstrated by piezoelectric composites that embed ceramic materials within polymer matrices.30–32 These systems often outperform single-phase materials, offering enhanced electromechanical coupling, lower electrical impedance, and improved mode stability while remaining cost-effective and suitable for flexible applications. By combining the functional properties of halide perovskites with the mechanical flexibility and chemical resilience of polymers, these composites aim to achieve enhanced performance, improved stability, and easier processability. In particular, the integration of halide perovskites with polyvinylidene fluoride (PVDF), a piezoelectric polymer known for its flexibility and high piezoelectric response, has shown great promise.33
The FASnI3–PVDF composite, for instance, exemplifies this approach. While FASnI3 provides the semiconducting and piezoelectric functionality, PVDF contributes mechanical robustness and environmental compatibility. This synergy has led to a composite material capable of delivering strong piezoelectric performance (d33 = 73 pC N−1)34 at relatively low processing temperatures (∼70 °C for the active layer), making it a compelling lead-free alternative to PZT for thin-film, flexible piezoelectric devices. Thin-film piezoelectrics have a wide range of potential applications, including lightweight, flexible substrates, and integration into wearable or miniaturized IoT devices.35–37
Conducting a life cycle assessment (LCA) is crucial for evaluating the environmental impacts of piezoelectric materials throughout their entire lifecycle, from raw material extraction to end-of-life disposal. These assessments are necessary to justify the replacement of one material with another that is environmentally superior. The pioneering work by Ibn-Mohammed between 2016 and 2018 was among the first to explore the environmental profiles of lead-free alternatives like potassium sodium niobate (KNN) and sodium bismuth titanate (NBT), finding that these materials offer no clear environmental advantage over PZT.38,39 He also contributed to a 2019 paper that delves into the advantages and the need for applying LCA analysis to ceramic materials.40 Hazeri's studies in 2017 and 2022 focused on energy harvesting in pneumatic tires, comparing the environmental impact of piezoelectric tires with conventional ones.41,42 Further advancements were published in 2024, continuing the exploration of the environmental impacts of KNN piezoelectrics. This study considered material density and concluded that KNN has lower environmental impacts, but also highlighted the need for more efficient extraction and purification methods for Nb2O5 to reduce energy consumption and waste emissions.43 The number of publications on the piezoelectric properties of halide perovskites has increased significantly in recent years,44 and there were also 155 active patents for KNN-based piezoelectric ceramics between 2017 and 2022.43 However, despite this growing interest, only these few studies have focused on conducting LCAs to validate the environmental benefits of these emerging materials.
This research aims to explore environmentally sustainable alternatives to conventional lead-based piezoelectrics by applying a life cycle assessment (LCA) methodology. As a case study, we evaluate a lead-free composite system consisting of the halide perovskite FASnI3 embedded in a PVDF matrix. While the analysis is limited to this specific configuration, it serves as a starting point for understanding the broader sustainability implications of emerging lead-free piezoelectric materials. The assessment follows a cradle-to-grave approach, including manufacturing impacts, energy harvesting performance during use, and end-of-life considerations. By focusing on a realistic device architecture, the study highlights key environmental trade-offs and potential benefits associated with moving away from traditional PZT ceramics, while acknowledging the stability challenges of FASnI3 and the environmental burden of PVDF. These findings contribute to the ongoing search for safer and more sustainable piezoelectric technologies for energy harvesting.
The system boundary for the LCA of piezoelectric materials spans from cradle-to-grave, encompassing cradle-to-gate, use phase, and end-of-life, see Fig. 1. In the cradle-to-gate study, both production processes and raw materials preparation for the piezoelectric materials were evaluated. An analysis of the industrial-scale process was performed, assuming finished products of piezoelectric thin films. Various compositions of PZT were selected, assuming the same manufacturing system for all of them, with changes mainly in the quantities of materials precursors used in this process. In the use phase, the devices are applied in energy harvesting scenarios, generating energy in response to mechanical stress. One notable advantage of piezoelectric materials is their ability to replace batteries and operate autonomously throughout their lifespan. This fact eliminates the need for periodic battery replacements and is particularly beneficial in environments where replacing batteries may pose significant challenges due to complexity. Additionally, Ibn-Mohammed has recently emphasized the importance of performing accident analysis alongside a conventional LCA to assess the environmental impacts of unforeseen events.47 In this study, such an accident scenario was considered, assuming the device breaks and disperses all the piezoelectric material onto the ground. This perspective complements the cradle-to-grave analysis by accounting for potential risks during the use phase.47
The study on end-of-life scenarios employed the model for waste-specific and regionalized models for residual material landfills and slag compartments published by Gabor Doka in the 2023 version.48 It has been assumed that the materials are deposited in a residual material landfill. This system boundary assessment enables a complete cradle-to-grave analysis, offering a clear understanding of the lifecycle impacts. It facilitates a meaningful comparison between traditional lead-based piezoelectrics and emerging lead-free perovskites.
The scarce literature revealed a diversity in the choice of functional units, ranging from specific components, such as a wheel in piezoelectric studies related to vehicular applications,41,42 to mass-based units like kilograms38,39,49 or per volume in m3.43 However, given the focus of this study on materials with varying densities and their application in energy harvesting, adopting area, measured in square meters (m2), as the functional unit emerged as a more pragmatic approach. This decision is particularly rational for piezoelectric materials used in thin films, as it allows for the consideration of the contact area where force is applied, enabling a more precise assessment of their performance in energy harvesting applications. Energy harvesting is another key parameter in evaluating piezoelectric performance; however, it is highly dependent on factors such as the operating frequency of the device and the applied pressure. To account for this energy in the analysis, this study evaluates energy harvesting under standardized conditions, specifically at a frequency of 20 Hz and applied pressures of 10 MPa and 100 MPa.
Fig. 2a illustrates the intricate process involved in the fabrication of traditional PZT thin films. The procedure starts with precision batch weighing of PbO, TiO2, and ZrO2. These components are then subjected to ball milling in the presence of isopropanol, forming a homogenous mixture. The resulting slurry is dried and subjected to a calcination process at elevated temperatures to achieve the desired crystalline structure. A subsequent round of ball milling, followed by another drying phase, prepares the material for sintering at an elevated temperature of 1000 °C. The sintered material is then deposited onto a substrate heated to 600 °C using magnetron sputtering to produce PZT thin films of the desired thickness, in this case 800 nm for easy comparison with the reported FASnI3–PVDF thickness. A final annealing step is employed to enhance the piezoelectric and structural properties of the films, optimizing them for their intended applications.38
The synthesis process of FASnI3–PVDF perovskite thin films, see Fig. 2b, starts with the dissolution of tin iodide (SnI2) and formamidinium iodide (FAI) in dimethylformamide (DMF). This mixture is then subjected to magnetic stirring at ambient conditions to ensure the formation of a homogenous FASnI3 precursor solution. Simultaneously, PVDF is dissolved in DMF and agitated magnetically at a controlled temperature of 60 °C to obtain a stable solution. These two solutions are carefully blended in a 0.5:
0.5 volume ratio and subsequently annealed at 70 °C to promote the formation of the perovskite structure. The annealed mixture is spin-coated onto a pre-prepared poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) film for a precise duration of one minute, laying the foundation for the 800 nm thick perovskite thin film. To optimize the film's performance, a final annealing process to 70 °C is implemented. This step helps to refine the microstructure and bolster the intrinsic piezoelectric and electronic characteristics of the film, ensuring that its properties are not only enhanced but also consistent.34
In Table 1, the different types of piezoelectric materials analyzed are documented alongside their d33 value. Efforts were made to identify materials with similar piezoelectric coefficients. For the perovskite FASnI3–PVDF, a d33 value of 73 pC N−1 was determined,34 while for the sputtered PZT ceramic, the d33 value was found to be 75 pC N−1.50 This equivalence in d33, a measure of the material's ability to convert mechanical stress into electrical charge, serves as a critical benchmark for assessing performance under similar operational conditions. It allows for a direct and fair comparison of their efficiency and effectiveness in energy conversion applications. Additionally, to analyze the usage stage, more commercial PZTs were considered to explore how their environmental impacts varies due to their different piezoelectric properties. Specifically, soft PZT materials such as PZT-5H, PZT-5A, and PZT-5J, see Table 1, were compared. They are commonly used as reference materials due to their large transduction coefficient which helps to generate high energy and power densities.51 These materials outperform hard PZT counterparts under low-frequency off-resonance excitation conditions.52
For the PZT laboratory case, the inventory of material and energy inputs was modeled based on data provided by Ibn-Mohammed's work (Table S1†).38 Adjustments were made to increase the amount of PbO (Table S2†) used compared to the initial data, compensating for lead oxide losses due to evaporation at high temperatures during the synthesis of PZT materials.55 Additionally, isopropanol was introduced as a solvent in the solid-state reaction.56 At industrial scale, a 95% solvent recirculation rate was assumed, and electricity consumption was calculated using theoretical equations derived from an industrial-scale thermodynamic study conducted by Piccinno.57 In both laboratory and industrial contexts, sputtering has been identified as one of the most widely used and effective techniques for depositing PZT in thin films.58 For laboratory-scale sputtering, a material loss of 61% was considered, while for industrial-scale sputtering, losses were assumed to be only 5%, based on the utilization of the entire target material and the implementation of a protective shield to recover residual material.59
The lead-free hybrid perovskite system consists of FASnI3 embedded in a soft PVDF polymer matrix with a volumetric ratio of 0.5:
0.5. The quantities for halide perovskite composite synthesis were derived from the study conducted by Pandey (Tables S3 and S4†).34 Electricity consumption for the spin-coating deposition process was measured at our laboratory using a CIRCUTOR MYeBOX 1500 Network Analyzer, capturing data every second from a Laurell WS-650Lz-23NPPB spin coater.
At an industrial scale, a 95% solvent recirculation rate was also assumed for perovskite deposition. While spin coating is commonly employed in laboratory settings, it is inefficient for large-scale production. The literature highlights several deposition methods suitable for industrial applications. For this study, slot-die coating was selected as the industrial deposition method due to its scalability, cost-effectiveness, and reproducibility. Energy consumption associated with slot-die coating was estimated using a reliable industrial-scale ref. 60.
Li-ion battery cell featuring a nickel manganese cobalt 811 (NMC811) cathode and a silicon-coated graphite-based anode has been utilized for energy study, sourced from Ecoinvent. This cell provides a specific energy capacity of 0.209 kWh kg−1. This comparison is significant as piezoelectrics could avoid the use of batteries for certain applications and consequently the impacts of their fabrication, see below.
Some data are not available in the Ecoinvent 3.9 database and have been modeled independently. Details on the modeling approach for these compounds are provided in the ESI.†
![]() | ||
Fig. 3 Schematic diagram of a typical piezoelectric device for energy harvesting, illustrating the encapsulation layer, substrate, and the central piezoelectric film between two electrodes. |
This study focuses on the comparison of the piezoelectric layer, assuming that the other layers remain constant across different materials. By holding these layers constant, we eliminate their influence on the results, allowing for a clearer environmental assessment of the piezoelectric materials being compared.
By treating the piezoelectric element as a capacitor, we can effectively model and compute the potential energy harvesting capabilities, thereby establishing a fundamental basis for assessing the efficiency and efficacy of these materials in converting mechanical energy into electrical energy.66 For the open circuit, voltage and current are determined by the following equations:67
![]() | (1) |
![]() | (2) |
Energy is computed by considering capacitance and time, taking into account that the energy stored in a capacitor follows the formula E = 1/2CV2. This energy formula can be expressed as follows:67
![]() | (3) |
![]() | (4) |
As a result, the energy equation can be written in the following way:67
![]() | (5) |
Coefficients such as d33 and εT33 have been found to generally increase with temperature for both soft and hard PZT samples. Additionally, thermal hysteresis was observed in all the studied material coefficients over the temperature cycle.68 Additionally, dielectric losses (tanδ) at low frequencies were found to be negligible, simplifying the comparison between different materials.69
Larger dimensions of the piezoelectric device yield greater energy output. Moreover, increasing the d33 coefficient and the applied force significantly enhances the energy harvested. Therefore, attaining high d33 values and low εT33 permeabilities will be crucial for achieving greater energy output. To ensure a fair comparison of the materials, all simulations in this study were performed under identical mechanical and geometric conditions. The same applied pressure, device area, and thickness were used in each case. With these parameters held constant, any variation in energy output can be attributed solely to the intrinsic electromechanical properties of the materials. Under such conditions, the ratio is recognized as a figure of merit (FoM) for energy harvesting applications.67,70 The corresponding FoM values for the materials considered are summarized in Table 1. A constant ambient temperature will be assumed to ensure consistency and comparability across the case studies. This decision stems from the observed variations in certain piezoelectric parameters with temperature.71
We use eqn (5) to represent the NET environmental impacts during the use of piezoelectric devices, considering both the energy produced over a given period and the battery impacts avoided as a result. This metric is crucial for evaluating the energy efficiency and environmental feasibility of these devices, as it provides insights into when the energy invested in their production will be offset.
NET = piezoelectric impacts − avoided battery impacts | (6) |
The avoided battery impacts (eqn (7)) were calculated based on the environmental impacts of producing an NMC811 battery, with an energy density of 0.209 kWh kg−1, as provided in the Ecoinvent database. Using this energy density and the energy generated by the piezoelectric device over a given period, we determined the equivalent amount of battery production that was avoided.
![]() | (7) |
Impact category | Unit | Abbreviation |
---|---|---|
Acidification | mol Heq+ | AP |
Climate change | kg CO2 eq | GW |
Ecotoxicity, freshwater | CTUe | EcoT |
Particulate matter | Disease inc. | PM |
Eutrophication, marine | kg Neq | EP-m |
Eutrophication, freshwater | kg Peq | EP-f |
Eutrophication, terrestrial | mol Neq | EP-t |
Human toxicity, non-cancer | CTUh | HT-nc |
Human toxicity, cancer | CTUh | HT-c |
Ionising radiation | kBq U-235eq | IR |
Land use | Pt | LU |
Ozone depletion | kg CFC11 eq | OD |
Photochemical ozone formation | kg NMVOCeq | POP |
Resource use, minerals and metals | kg Sbeq | RU-m |
Water use | m3 depriv. | WD |
Cumulative energy demand | MJ | CED |
The LCA calculations were performed twice: first using a custom spreadsheet tool in Microsoft Excel and then directly in SimaPro 9.5. Monte Carlo simulations with 1000 iterations were conducted in SimaPro to quantify uncertainties, providing results within a 95% confidence interval for environmental impacts.
The manufacturing process of FASnI3–PVDF composite demonstrates a lower environmental footprint, primarily due to its simpler fabrication, which avoids high-temperature treatments and minimizes energy usage. In contrast, PZT production is dominated by the impacts of material inputs, such as lead (Pb), contributing heavily to acidification and ecotoxicity, as well as the energy-intensive processes of electricity and heat consumption, which drive its contributions to climate change and CED.
These findings align with previous studies, which emphasize that energy consumption during PZT manufacturing is a major driver of environmental impacts, particularly in the climate change and CED categories39 Additionally, these results are consistent with the latest environmental assessment on piezoelectric devices, which indicates that the greatest impacts in the manufacturing process arise from raw materials, particularly PbO, and electricity consumption.43 This data indicates, on one hand, that the use of a Pb-free materials has a huge effect in the reduction of environmental impacts of piezoelectric materials input. While alternative methods such as hydrothermal synthesis,73 co-precipitation74 or pulsed wire discharge75 offer potential for lowering sintering temperatures, the heat treatments required for PZT ceramics remain more energy-intensive than those for the FASnI3–PVDF composite.
Fig. 4b illustrates the lab-scale distribution of manufacturing impacts for the FASnI3–PVDF composite across all contributing sources: PVDF emerges as the largest contributor to climate change, while DMF is the dominant factor in cumulative energy demand (CED). At industrial scale, however, the implementation of solvent recovery systems can significantly reduce the energy impact associated with DMF, leaving PVDF as the main environmental hotspot—surpassing even the contributions from the perovskite precursors FAI and SnI2, as well as other process-related emissions and energy use.
When normalized to a one-kilogram functional unit, the cradle-to-gate impacts of the FASnI3–PVDF composite are 24.2 kg CO2-eq per kg for climate change and 443.6 MJ kg−1 for cumulative energy demand (CED). These values are substantially lower than those reported at laboratory scale for other lead-free ceramic alternatives, such as KNN (248.6 kg CO2-eq per kg and 4123.7 MJ kg−1)38 and NBT (30.4 kg CO2-eq per kg and 447.0 MJ kg−1).39 Moreover, the implementation of solvent recovery at industrial scale is expected to further reduce these impacts, reinforcing the environmental advantage of FASnI3–PVDF composite over both PZT and other Pb-free piezoelectric materials.
The environmental impacts increase further during the deposition stage, as shown in Fig. 4c and d. For PZT, thin film deposition causes a 97% increase in its climate change impact and a 145% rise in its CED. In contrast, the deposition process for FASnI3–PVDF composite also contributes to its impacts, but these remain significantly lower overall. This stage is dominated by spin-coating, a widely used laboratory technique with high material losses (up to 99%). As a result, the climate change impact of FASnI3–PVDF composite reaches a mean value of 0.25 kg CO2 eq per m2, while its CED is 6.51 MJ m−2—reductions of 85.69% and 84.56%, respectively, compared to PZT. The laboratory data underscore that the use of lead-free materials, such as FASnI3–PVDF composite, can substantially reduce the environmental burdens of piezoelectric materials.
The analysis reveals significant differences between the FASnI3–PVDF perovskite composite and PZT ceramics across various environmental categories. As shown in Fig. 5a, PZT consistently exhibits higher impacts, aligning with laboratory-scale observations. For the FASnI3–PVDF composite, industrial-scale slot-die deposition demonstrates a notable advantage over the spin-coating process used in laboratories, dramatically reducing environmental impacts due to significantly lower material losses. Consequently, the industrial-scale environmental impacts of FASnI3–PVDF are markedly lower than those of PZT, further supported by laboratory-scale comparisons (Fig. 4b and c), which highlight the superior environmental performance of the perovskite composite.
Fig. 5b focuses on the cumulative energy demand (CED) and climate change categories, two critical metrics in this study. FASnI3–PVDF exhibits substantially lower values than PZT, primarily due to reduced electricity consumption during industrial manufacturing, leading to lower emissions and potential cost savings. The inclusion of 95% confidence intervals ensures data reliability while accounting for uncertainty in the analysis.
Additionally, Fig. 5c compares various PZT compositions, showing relatively similar environmental impacts across formulations, with a slight increase observed in titanium-rich compositions due to higher density and associated resource demands. In PZT manufacturing, electricity consumption during deposition and the use of raw materials, particularly lead, are the primary contributors to environmental impacts. For acidification and ecotoxicity categories, raw materials account for 30% and 50% of impacts, respectively, with lead playing a dominant role.
From a performance perspective, the largest piezoelectric constant of the PZT ceramics here studied occurs with a Zr/Ti ratio of 52:
48, corresponding to a morphotropic phase boundary (MPB) that enhances electromechanical properties. This makes PZT highly effective in applications such as sensors, actuators, and transducers.55 However, different PZT phases—tetragonal, rhombohedral, and occasionally cubic—exhibit unique electrical and structural properties, which cater to specific applications, including high-temperature sensors, MEMS devices, and ferroelectric memory.76,77
Despite its technical advantages, PZT's inherent reliance on lead and higher needs of electrical power contributes significantly to its environmental burden, as shown in Fig. 5c. Addressing these challenges requires transitioning toward lead-free materials with less fabrication power demands, such as FASnI3–PVDF halide perovskite composite. The industrial-scale optimization of the FASnI3–PVDF composite manufacturing process underscores its potential as a safer and more sustainable alternative, offering significant reductions in environmental impacts. By minimizing energy usage and material losses, this composite enables simpler, greener production while maintaining strong performance as a piezoelectric material.
In the context of piezoelectric energy harvesting, it is more meaningful to analyze device performance in terms of cycles rather than operational time, given the wide range of frequencies at which these devices can operate. Higher operating frequencies enable more energy generation in less time but also shorten the device's lifespan due to increased mechanical stress. To provide a practical benchmark, a reference frequency of 20 Hz was selected, as it yielded the best performance for the FASnI3–PVDF composite studied. This aligns with typical applications in literature, where many piezoelectric harvesters are designed to operate within low-frequency ranges (0–100 Hz), commonly targeting vibrations from human motion (2–5 Hz),78 road traffic (3–10 Hz at 30–100 km h−1),79 and household appliances such as refrigerators or air conditioners (25–100 Hz).80 It is important to emphasize the importance of resonance frequency, where energy output is maximized. Aligning this frequency with the target range can be achieved through design strategies such as modifying geometric dimensions, adding mass, or utilizing flexible materials. Frequency tuning mechanisms and multi-modal designs further enhance efficiency, particularly in environments with variable or low-frequency vibrations.
The pressure exerted on piezoelectric materials plays a crucial role in determining both the amount of energy generated and the durability of the device over multiple cycles. Most studies on piezoelectric energy harvesters focus on pressure ranges between 0.1 and 10 MPa, commonly associated with applications such as human motions (e.g., hand movements81 or walking), tire pressures from cars and trucks,82,83 or energy harvesting from machine vibrations. However, PZT-based devices are capable of withstanding pressures exceeding 100 MPa over extended cycles,84,85 demonstrating their suitability for high-pressure environments and long-term applications.
The evolution of net energy at constant pressures of 10 MPa and 100 MPa is analyzed for the FASnI3–PVDF composite and various piezoelectric ceramics, as shown in Fig. 6. A discontinuous bar indicating the time required to complete a specific number of cycles at 20 Hz has also been added to provide a more tangible understanding of these values. It is important to note that these results assume that a piezoelectric material can endure the required number of cycles. However, given the uncertainty regarding the lifetime of the perovskite composite, it is crucial to explore the implications if the material cannot withstand such a high number of cycles, as discussed in the next section.
![]() | ||
Fig. 6 NET cumulative energy demand of PZTs and FASnI3–PVDF perovskite, (a) at 10 MPa and (b) 100 MPa. The operating time is indicated next to the dashed line for a frequency of 20 Hz. Note that NET CED is calculated from eqn (6) and considering just the piezoelectric active layer. |
At 10 MPa, the halide perovskite composite requires less than 106 cycles to return the energy consumed during its fabrication, while the piezoelectric ceramics need cycles of 2–4 orders of magnitude more, see Fig. 6a. Among the PZT compositions, the PZT with a d33 coefficient similar to that of the perovskite offers the poorest performance in terms of recovering the energy required for its fabrication. The commercial PZT ceramics exhibit a similar behavior, with PZT-5H showing the highest efficiency due to its high d33 coefficient, being able to return the energy consumed in its fabrication after 107 cycles, followed closely by PZT-5J and finally PZT-5A. At 100 MPa, the results are analogous, with the added benefit of recovering the energy investment much sooner due to the higher operating pressure, see Fig. 6b. The perovskite would need less than 104 cycles, while PZT-5H would require more 105 cycles, and the PZT with a low d33 coefficient would require 108 cycles. Although the PZT-5H ceramics exhibit the highest room temperature piezoelectric properties and can reduce manufacturing impacts more quickly, this composition has a Tc value of 170 °C, which is very low compared to other materials. This can be a limiting factor for applications in systems where the device could be exposed to temperatures exceeding its Curie point.12,86
Fig. 6 shows that the FASnI3–PVDF composite material can recover the energy invested in its production significantly faster than piezoelectric ceramics under these conditions. This superior performance is attributed to its lower CED during fabrication, starting from a much smaller initial energy input. Additionally, the lower permeability of the perovskite enables it to generate more energy in open-circuit conditions.
We also analyze in Fig. S1† the time required for both materials to recover the energy consumed during their fabrication, considering only the energy harvested through energy harvesting and without accounting for the avoided impacts from replacing batteries. The recovery time is significantly longer under these conditions, being approximately two orders of magnitude higher compared to scenarios where the avoided impacts from battery replacement are included. The PZT-5H was selected to evaluate the remaining net impacts across various impact categories due to its superior operational performance. Comparisons were made with the FASnI3–PVDF composite and a PZT material exhibiting a d33 value similar to that of the perovskite-based composite. The calculations for these categories were carried out at a pressure of 10 MPa, as shown in Fig. 7. The categories of ecotoxicity and acidification require fewer cycles to offset the impact produced during manufacturing, for both materials. In these categories, the halide perovskite needs approximately 104 cycles, while the PZT-5H requires more than 106 cycles. In terms of climate change and ecotoxicity, the duration needed to offset the environmental impact is longer. FASnI3–PVDF composite material requires 105 cycles to neutralize its impact, whereas the PZT-5H ceramic demands a substantially greater number of 107 cycles. While both materials are capable of mitigating their initial production impacts over time, the halide perovskite composite demonstrates significantly greater effectiveness in achieving substantial environmental benefits within a practical operational timeframe. This further underscores the superior potential of the halide perovskite composite in providing long-term sustainability compared to the PZT-5H ceramic. When comparing the PZT with a d33 value similar to that of the perovskite composite, its environmental performance is markedly worse across all categories. This material demands significantly more cycles to offset its initial production impacts.
![]() | ||
Fig. 7 Net environmental impacts of FASnI3–PVDF, PZT with similar d33 and PZT5H under a pressure of 10 MPa across multiple cycles: (a) climate change, (b) freshwater ecotoxicity, (c) acidification, and (d) resource use. Note that NET impacts are calculated from eqn (6) and considering just the piezoelectric active layer. |
Although piezoelectric materials as energy harvesters cannot generate as much energy as other renewable sources like solar power, they are indispensable in specific applications. For instance, replacing batteries in large-scale sensor networks can be costly and challenging, especially in hazardous or remote environments.87 In this context, the FASnI3–PVDF composite not only exhibits lower initial environmental impacts but also enhances its environmental footprint by eliminating the need for batteries in specific applications in a shorter period of time. However, the longevity of perovskites still falls short of that of PZT ceramics. Literature reports that similar perovskite composites, such as MASnI3-PVDF, can endure up to 104 cycles under 0.5 MPa,88 whereas PZT ceramics demonstrate significantly greater durability, with lifecycles ranging from 107 to 1010 cycles in low-frequency applications.89–92 This durability advantage of PZT ceramics highlights the importance of improving the operational lifespan of perovskite-based materials to fully realize their potential as sustainable alternatives in piezoelectric applications.
In the scenario presented, the total environmental impact of the piezoelectric device is the combination of the impacts generated during its production (from cradle-to-gate) and those caused by its accidental release into the soil. The calculations show that 99.79% of the impact is in the human toxicity, non-cancer category for PZT result from its production, see Fig. 8a. While accidental release accounts for only 0.21% of the total impacts, it is noteworthy that lead is the dominant contributor, responsible for 99.99% of the non-cancer toxicity. This highlights the importance of considering the entire lifecycle and potential accidental releases when assessing the environmental impacts of piezoelectric materials.
Previous studies on piezoelectric materials have primarily concentrated on the environmental impacts associated with their manufacturing processes, often overlooking the impacts of their use and end-of-life stages. However, a cradle-to-grave analysis requires a comprehensive assessment that includes end-of-life impacts, as this phase addresses disposal, recycling, and the potential release of hazardous substances. Fig. 8b and c presents a comparison of end-of-life impacts for both PZT ceramics and FASnI3–PVDF composite materials in the categories of climate change and CED. The results indicate that the perovskite-based composite generates roughly 34% of the impacts associated with the ceramic piezoelectrics. This difference primarily arises due to the end-of-life methods applied, which have a greater influence on the impact categories than the material composition itself. Furthermore, due to the lower density of halide perovskites, the perovskite composite demonstrates reduced environmental impacts per square meter, which is the functional unit adopted for this study. In terms of waste management, the recovery of PZT remains a major challenge. Although several recycling strategies have been proposed, they are still in early stages and face technical barriers such as energy-intensive processing and material contamination.93 For example, a new recycling approach has been proposed where crushed PZT ceramics are reassembled using low-temperature sintering techniques, significantly reducing the energy demand compared to traditional processes. However, these methods are still proof-of-concept and unproven at industrial scale.94 Moreover, the intrinsic lead content of PZT poses serious environmental and regulatory challenges even if recycling were viable.
Recycling research is also emerging for halide perovskites. For example, a recent aqueous-based method demonstrated the separation of halogens from perovskite layers via distillation, enabling the recovery of valuable elements with minimal material degradation. Although these strategies are still just beginning, they illustrate the growing focus on enabling material circularity for lead-free perovskite technologies.95 Similarly, PVDF recycling is already practiced in the context of silicon-based solar cells, where it is used as a backsheet. Several methods have been developed to recover PVDF from end-of-life photovoltaic modules, demonstrating its potential for separation and reuse.96 However, many of these methods rely on pyrolysis, which, although effective in breaking down polymeric materials, requires high temperatures.97,98 This leads to significant energy consumption and the release of harmful fluorine compounds, raising concerns about the environmental impact and safety of this recycling method.
In the cradle-to-grave analysis, a sensitivity study was conducted considering different lifetimes for the perovskite-based devices, see Fig. 8d and e, assuming operation at 10 MPa. The scenarios evaluated include a currently reported lead-free perovskite lifetime of 104 cycles,88 as well as extended lifetimes of 105 and 106 cycles. At the end of the device's operational life, replacing the FASnI3–PVDF piezoelectric with additional units is required to meet the desired number of cycles. In these scenarios, eqn (6) and (7) are also applied to calculate the avoided impacts associated with not using conventional batteries.
For 104 cycles, all scenarios involving the FASnI3–PVDF composite exhibit a lower environmental impact compared to PZT. However, for 105 cycles, the composite with a lifetime of 104 cycles matches the impact of PZT due to the need for 10 additional devices to achieve the required operational duration. Similarly, for 106 cycles, the composite with a lifetime of 105 cycles reaches the same environmental impact as PZT.
As a result, for lifetimes of 104 and 105 cycles, the halide perovskite system struggles to mitigate the environmental impacts associated with manufacturing multiple devices to meet the required cycles. In these scenarios, both PZT5H and the PZT material with a d33 similar to that of the FASnI3–PVDF composite demonstrate better environmental performance. However, for a lifetime of 106 cycles, the halide perovskite composite achieves a sufficiently long operational period to offset its production impacts, making it a competitive alternative. In contrast, PZT5H would require over 107 cycles, and the PZT with similar d33 would need an even higher number of cycles to achieve a comparable level of environmental impact mitigation.
Moreover, the fabrication of these piezoelectric ceramics requires high temperatures, which introduces further risks and significantly increases energy consumption. In contrast, Pb-free halide perovskite materials can be produced at ambient temperatures, significantly reducing associated hazards and energy demands. There is a study that has examined the human toxicity and ecotoxicity of perovskite solar cells (PSCs), and the results indicate that PbI and PbO exhibit very similar toxicity levels in these aspects. Therefore, producing Pb-based halide perovskites might not be a favorable option, and it would be better to avoid the use of lead altogether.100 Additionally, when evaluating alternatives to replace Pb in PCSs, Sn emerges as a less toxic option due to its degradation into the environmentally benign compound SnO2 upon exposure to air.101,102
The main challenge in fully adopting the substitution of lead with tin-based materials lies in establishing characterization factors for human health impact categories related to tin, which are currently unavailable. Critical effects such as neurotoxicity and carcinogenicity remain under debate, creating a significant gap in assessing the environmental and health impacts of tin-based perovskites. Without these toxicity factors, it is not possible to make a fair comparison between lead-based and tin-based materials in terms of human toxicity.
Meanwhile, halide perovskite-based materials show promise due to their lower environmental impact compared to traditional PZT ceramics, it is also important to consider the toxicity of the solvents used in their production. The most commonly used solvent for perovskite layer deposition, N,N-dimethylformamide (DMF), is listed as a substance of very high concern (SVHC) by the European Chemical Agency under the REACH regulation.103 For FASnI3 perovskite, DMF contributes approximately 20% of the environmental impacts associated with perovskite fabrication, emphasizing the importance of exploring alternative solvents to mitigate these impacts. Studies suggest that dimethyl sulfoxide (DMSO), a less toxic solvent, could serve as a suitable alternative.104 Formulations incorporating this solvent, such as a DMSO:
DMF (9
:
1, v/v) mixture,105 have been developed to significantly reduce DMF usage, thereby further minimizing the environmental footprint of halide perovskite production. This highlights the importance of replacing SVHCs with less dangerous substances or technologies wherever feasible. Consequently, there is a critical need for further research and development in identifying and utilizing safer alternatives in the production of halide perovskite materials to minimize their overall environmental and health impacts. In particular, PVDF is chemically related to per- and polyfluoroalkyl substances (PFAS), a group of persistent and bioaccumulative compounds increasingly scrutinized due to their toxicity and resistance to degradation. While PVDF is not currently classified as hazardous under the ECHA's harmonized classification system. No hazards have been officially recognized but some health risks are routinely reported including skin and eye irritation and respiratory effects.106 As part of its broader PFAS restriction strategy, ECHA is expected to assess possible regulatory measures specifically targeting PVDF by the end of 2025107 This anticipated regulatory development underscores the importance of monitoring policy changes and exploring non-fluorinated alternatives for future applications.
Additionally, the end-of-life treatment of both PZT and halide perovskite materials includes considerations for disposal in safety landfills, introducing variability based on local regulations, landfill technologies, and material composition. This aspect introduces additional uncertainty into the LCA and highlights the need for improved recycling and disposal strategies to mitigate potential environmental harm. Other components that encompass the final product, such as electrodes or substrates, have not been considered. There are numerous possibilities for these layers, and it would be valuable to study their impact on environmental effects in the future.
These limitations underscore the necessity for ongoing research and development to refine life cycle assessments of piezoelectric materials, particularly in the context of emerging lead-free alternatives like tin-based halide perovskites. Addressing these challenges will be crucial for accurately evaluating their environmental performance and potential for sustainable applications.
A detailed examination of the manufacturing processes reveals that the raw materials used in PZT production generate significantly higher environmental impacts than those required for lead-free halide perovskites, driven in part by the inclusion of lead, especially in the categories related to toxicity. Additionally, PZT ceramic production involves high-temperature processes, resulting in greater energy consumption and elevated manufacturing risks. In contrast, the synthesis of halide perovskites occurs at much lower temperatures (approximately 100 °C), making it a safer and less energy-intensive process. Notably, the fabrication of FASnI3–PVDF composites generates less than 20% of the environmental impacts associated with conventional PZT ceramics, or even less for certain impacts.
For thin-film devices, halide perovskites composites also support more energy-efficient deposition techniques which not only reduce energy costs for manufacturers but also significantly lower the overall environmental footprint. These advancements align with the overarching goal of developing sustainable, lead-free piezoelectric materials.
The application of these devices in energy harvesting further highlights their benefits. By using these materials, the continuous replacement of batteries in complex environments where consistent energy supply is necessary, such as sensor systems, can be avoided. Halide perovskites or their composites, in particular, offer significant advantages due to their low energy requirements for fabrication and reduced permeability. The FASnI3–PVDF composite achieves a competitive environmental advantage over conventional batteries when it operates for a lifetime of at least 106 cycles for 10 MPa (tire pressure), as this duration offsets the environmental impacts of its manufacturing, least 104 cycles for 100 MPa.
Our findings demonstrate that FASnI3–PVDF composites have significant potential for sustainable piezoelectric energy harvesting. This highlights lead-free halide perovskites as promising solutions for energy harvesting applications and represents a step toward more environmentally responsible technologies.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5se00717h |
This journal is © The Royal Society of Chemistry 2025 |