Ferroelectric nanofibers with nanoconfined tellurium nanobeams for mechanical and thermal energy harvesting and wearable healthcare

Utsa Sarkar a, Hari Krishna Mishra a, Ajay Kumar a, Shanker Ram *b and Dipankar Mandal *a
aQuantum Materials and Devices Unit, Institute of Nano Science and Technology, Knowledge City, Sector-81, Mohali 140306, India. E-mail: dmandal@inst.ac.in
bMaterials Science Centre, Indian Institute of Technology, Kharagpur, 721302, India. E-mail: sram@matsc.iitkgp.ac.in

Received 4th June 2025 , Accepted 19th September 2025

First published on 22nd September 2025


Abstract

Large-area sensing devices open up great potential to trigger continuous self-powering, multifunctional sensors with responses to multiple stimuli. Herein, we report a self-powered flexible hybrid piezo- and pyro-electric nanogenerator (NG) based on Te-reinforced poly(vinylidene fluoride) (PVDF) electrospun nanofibers. It is tested to generate electricity from waste mechanical and thermal energies at room temperature. An electroactive phase (∼94%) is built in along the interface (an electrified jet) of dipoles poled perpendicular to the PVDF backbone ‘–C–C–’ chains. As a proof-of-concept, the fabricated NG generates remarkable electricity, reaching a power density of 4.2 μW cm−2 under periodic mechanical stimulation, showcasing its potential for the efficient conversion of ambient mechanical energy into electricity. The intrinsic photothermal heat-localization effect of poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) as an organic electrode enhances the pyroelectric response, yielding a pyroelectric coefficient of 40 μC m−2 K−1 (which is 4 times higher than that of the undoped fibers), making it capable of detecting low thermal oscillations. Therefore, it has the potential to monitor physiological conditions and body temperature, making it suitable for remote infectious disease surveillance.



New concepts

We introduce a self-powered, flexible, pressure- and temperature-responsive hybrid sensor based on tellurium (Te)-reinforced poly(vinylidene fluoride) (PVDF) electrospun nanofibers. This design strategically harnesses Te, a monoelemental chalcogenide with exceptional piezo- and ferro-electricity, to precisely modulate exotic charges within a 1D confinement. This approach not only boosts energy harvesting capabilities but also capitalizes on the synergistic effects of mechanical stretching and nanoconfined stress, significantly enhancing the alignment of molecular dipoles and optimizing ferroelectric properties, which are of prime importance in the era of sensor-based technologies. Our work demonstrates aligning the self-confined hexagonal phase of monoelemental Te along the fiber axis by utilizing high rotational speeds during electrospinning, thereby inducing a compressive lattice strain at the Te-PVDF interface. The quantum confinement of the Te nanobeams subtly alters their vibrational modes, fine-tuning the electronic states and significantly enhancing the charge-carrier coupling with the PVDF dipoles within the hybrid fibers. This process culminates in a robust heterojunction that amplifies both the piezoelectric and pyroelectric responses. Furthermore, we leverage organic PEDOT:PSS/Xyl electrodes to intensify photothermal heat localization compared to conventional metal electrodes, which markedly enhanced the pyroelectric response. This design achieves superior flexibility and durability while enabling highly sensitive pressure and temperature detection, establishing a new benchmark for self-powered wearable healthcare and remote infectious disease monitoring.

Introduction

Self-powered pressure sensors and electronic skin (e-skin) based devices have garnered growing attention recently in healthcare monitoring, biometrics and the ‘Internet of Things’ (IoT).1–13 The idea of ‘internet connectivity’ is expanding beyond the traditional computer platforms of physical equipment and routine medical diagnosis tools.6,10 Efforts are being directed to the development of smart ‘energy conversion’ devices based on self-sustaining piezo-, pyro- and ferro-electrics integrated with a hybrid nanogenerator to efficiently and simultaneously convert mechanical and thermal energies into a powerful electrical signal at room temperature. Several compatible dopants have been used to promote the functionalities of wearable devices.1,14 Fragility limits the utility of inorganic doping in such devices. Thus, despite their inferior piezo- and pyro-electric features, organic polymers are being used in smart sensors and NGs owing to their good compatibility, good thermal stability, flexibility, inertness in ambient air, low cost, and overall good eco-friendliness and biocompatibility.6,15

PVDF, a linear chain polymer and its light-weight copolymers are highly flexible, thermoplastic and electroactive with good piezo-, pyro- and ferro-electric properties that make them useful for smart electronic devices.10–12 Crystalline PVDF exists in three major configurations: a non-polar α-phase, a polar β-phase, and a semi-polar γ-phase based on the polarities of ˙CH2 and ˙CF2 on its ‘–C–C–’ backbone chains.10,16 The β-PVDF phase is the most electroactive ferroelectric (FE) phase for incorporation in piezo- and pyro-electric NGs, with the largest spontaneous polarization (Ps) and piezo-/pyro-electric coupling coefficients.1,17 Mechanical stretching,17,18 high electric field poling,17,19 spin coating,20 and electrospinning1,17 are used to stretch and induce the electroactive β-PVDF phase. Among them, electrospinning is highly efficient to stretch the 1D β-PVDF configuration (D: dimension) with the dipoles aligned on the PVDF macromolecular chains, as the stretching force coupled with the electric field induces in situ polarization of duly enhanced piezoelectric properties. β-PVDF chain stretching under uniaxial stress self-aligns the dipoles, poled transverse to the stretching, resulting in enhanced piezoelectric potential and thereby efficiency. Nonetheless, due to the modest energy difference, the β phase relaxes back to α-PVDF.17 Heterojunctions, built-in with a strong electric field via fillers—such as graphene, graphene oxides (GO),12,17 metals,21 carbon nanotubes (CNTs),12,22 metal sulfides/oxides,23–25 CsPbI3 type halides,26 Ag2CO3,14 and 2D transition metal dichalcogenides (TMDCs)24—have shown to keep the dipoles mutually aligned perpendicular to the β-PVDF chains, which also restrains their relaxation back to the non-polar α-PVDF phase. In particular, collective electrostatic interactions between localized charges (fillers) and poled dipoles (polymer) at the heterojunction promote the stabilization of the β-PVDF phase hybrid nanostructure.

Molecular engineering of 2D architectures beyond GOs and TMDCs is a simple strategy for tuning exotic charges confined in a 2D-limited space.27,28 For example, single layers of group III to group VI elements have led to breakthrough electronic materials with game-changing properties. In particular, the chalcogen Te possesses excellent piezo-, ferro- and pyro-electricity,29–31 good thermoelectric32,33 and non-linear optical behaviors,34 and gate-tunable spin–orbit interactions useful for neuromorphic computing,28 photodetectors,35–37 and energy harvesting devices.27,28 One challenge is the fabrication of large-area sensors for healthcare, biometrics, and other applications. In this scenario, Te nanocrystals (NCs) have multiple effects on the nucleation of the electroactive phase and energy harvesting functionalities in hybrid PVDF composites, such as: (i) stretching the PVDF polymer chains, (ii) poling the dipoles at the chains, (iii) increasing exotic surface charges, and (iv) coupling the charges (also spins) and dipoles regulated at cohesive nearest-neighbour interactions in a self-powered device. The two phases create a built-in robust structure (elastomer-like) in which the stretching force acts along the electrified jets. Thus, both piezo- and pyro-electric phases can be integrated into a single NG device. The open issues are to develop the science and technologies of self-powered inorganic–organic hybrid devices.

Herein, we report electrospun nanofibers of Te-reinforced PVDF polymer, which binds the Te NCs to be self-aligned along the fibers. A self-powered NG (the first report of its type) is fabricated from the fibers with built-in ferro-, piezo- and pyro-electric responsiveness for tracking physiological diseases and human body temperature in the context of remote infectious disease monitoring applications. The electroactive β/γ-PVDF phase is enhanced as the ˙CH2 and ˙CF2 dipoles are poled at the chains aligned along the fibers. The cohesive Te-PVDF interactions pave the dipole orders at the interfaces. A device with compatible organic electrodes and a pyroelectric coefficient enhanced by four orders of magnitude (p of ∼40 μC m−2 K−1) precisely detects high-temperature fluctuations. The device is tested for a wide range of pressure–temperature sensing with good flexibility, durability, cyclic stability, and electrical output, proving its potential in detecting physiological responses and variations in temperature, expanding its applications in real-time health care monitoring.

Results and discussion

Fig. 1a presents the X-ray diffraction (XRD) pattern of the single-phase small crystallites of α-Te (average size Dc = 30 nm) with a hexagonal crystal structure,38,39 which were used to dope the PVDF polymer to form the hybrid nanocomposite of electrospun fibers. Small α-Te crystallites are critical to uniformly incorporate into PVDF fibers and regulate the Te-PVDF interactions over sufficiently large interfaces. The interplanar dhkl spacing in the XRD peaks indicated average lattice parameters of a = 0.4443 nm and c = 0.5913 nm, with a lattice volume Vc = ½(3)1/2a2c ≅ 0.1011 nm3 and X-ray crystal density ρc = 6.286 g cm−3. The small α-Te crystallites are well surface-compressed (strained) upto 0.90% larger density compared to the bulk density of 6.230 g cm−3. The α-Te lattice has been shown to contract in 1D nanowires.40 Further, the present α-Te exhibits anisotropic growth, demonstrated by its enhanced (100) peak intensity (relative to the most intense (101) peak) of I100/I101 = 28%, compared to a value of I100/I101 = 16% in the bulk state. Primarily, lateral growth dominates along the a-axis. The I100 value is further enhanced to 87% in the Te-reinforced PVDF fibers, which shows that the α-Te is aligned along the fibers (Fig. S1). A compressive strain γc ≅ (2d200d100)/d100 = (−) 2.29% was estimated (shown in the inset) from the interplanar spacings of d100 = 0.3448 nm and d200 = 0.1888 nm observed for the respective peaks. The field-emission scanning electron microscopy (FE-SEM) images in Fig. 1b display long nanoprisms (or nanobeams) (w ≅ 200–500 nm widths) grown along the (001) facets of the model given at the right. The nanobeams contain large surfaces that form effectively large interfaces when doped in the PVDF fibers.
image file: d5mh01052g-f1.tif
Fig. 1 (a) XRD pattern of α-Te nanobeams, and their (b) FE-SEM images (models at right) and (c) Raman bands with schematic energy levels shown in the inset. (d) The three A1 bands split up in Te atoms, oscillate along and perpendicular to the nanobeams. (e) A TEM image of a PVTe1.0 fiber with embedded α-Te and (f) d100 lattice images of α-Te bound by the d110* polymer chains via an interface, as shown in (g) a model structure.

In the Raman spectrum in Fig. 1c (full spectra are provided in Fig. S2), the Raman bands of three well-known groups of E1 (Te–Te bending), A1 (Te–Te breathing) and E2 (asymmetric Te–Te stretching along the c-axis) modes of Te atoms oscillating in a chain are observed at 103, 125 and 147 cm−1. Duly enhanced phonon frequencies, if compared to thin Te@SiO2 films (94, 122, and 142 cm−1, respectively),35 anticipate the α-Te is quantum confined, as also evidenced from the XRD pattern. As illustrated in Fig. 1d, the three A1 bands v1, v2 and v3 are located at 116, 125 and 131 cm−1 and correspond to a transverse, a longitudinal, and a mixed oscillation, respectively, of the Te atoms in the nanobeams. Additionally, the E2 band exhibits three parts (marked in a schematic energy level diagram in the inset in Fig. 1c). No asymmetry is visible in the E1 mode. Coupled phonons with the valence 5s2p4 Te electrons (in the Te-atoms oscillating at the triplet ground electronic state 3S1) results in the multiple states.

The typical transmission electron microscopy (TEM) image in Fig. 1e shows how a reinforced fiber PVTe1.0 embeds α-Te of small rhomboids, which are grown edge-on towards the (001) facets lying along the PVDF fibers. This means the PVDF dictates the α-Te to order, in support, on and along a stable interface of the Te atoms bonded to the ˙CH2 and ˙CF2 dipoles along the fibers. A higher-magnified view reveals that the d100 = 0.3450 nm Te lattices (Fig. 1f) are inclined at ϑ = 32°, bonding to the PVDF β-phase at the d110* = 0.3825 nm chains via an interface (disordered, w = 2.0 nm) in an integrated structure. Fig. 1g illustrates a model in which the two phases are mutually ordered via an interface in a fiber. The aligned d110* polymer chains are reflected in the duly enhanced intensity of the d110* XRD peak (Fig. S1b). The FESEM images in Fig. 2a and b present bead-free fibers (insets show histograms of the fibers with uniformly distributed diameters) electrospun at Rs = 500 rpm (w = 246 ± 13 nm) and 3000 rpm (w = 235.4 ± 17 nm), respectively. Relatively narrow distributions are observed for the fibers processed under these two conditions, which confirms their homogeneity across the macroscopic area of the sample. A large fraction φ → 90%, of oriented fibers is achieved at the optimized Rs of 3000 rpm (≤15% at Rs = 500 rpm) (Fig. 2c and d). The fast Fourier transform (FFT) images (insets) present diverged intensities at all angles (a Gaussian-shaped curve) in the randomly oriented fibers, while majority of the fibers are oriented at 92° when electrospun at the faster Rs. The pixels, scattered in an arc in the randomly oriented fibers, converged in an ellipsoid in the preferentially aligned fibers. Elemental mapping (Fig. 2e) confirmed that C, F and Te were uniformly distributed in the α-Te reinforced PVTe1.0 fibers. In the PVTe nanofibers (Fig. 2f, g and h), the (001) planes of the α-Te NCs face up along thick fibers, (w ≥ 150 nm), while the lateral α-Te faces are set along thinner fibers, as projected in the models.


image file: d5mh01052g-f2.tif
Fig. 2 FESEM images of the PVTe1.0 fibers (distributions of their diameters are shown in histograms in the insets) electrospun at (a) 500 rpm and (b) 3000 rpm and (c and d) their different fractional (φ) orientations, respectively, with the FFT images in the insets. (e) A typical TEM image of a Te-nanobeam (rhomboid) embedded in a PVDF fiber, with the C, F and Te mapping shown at right. A closer view of (f) randomly oriented and (g and h) aligned fibers, showing that the Te NCs are embedded either along the (001) facets, or the lateral facets.

The FTIR bands (Fig. 3a, full spectra are provided in Figures S3 and S4) reveal successive increase in the β-PVDF phase fraction with α-Te doping (plotted in Fig. 3b). A quantitative assessment of the concentration dependence of the β-PVDF phase fraction on the Te doping (Fig. S5, associated discussion S2) shows a maximum Fβ of 82% (electroactive fraction FEA ∼94% at Rs = 3000 rpm). The characteristic β-PVDF phase bands at 840 and 1277 cm−1 are enhanced in PVTe1.0 at the expense of the α-PVDF phase bands at 763, 796 and 977 cm−1, which were used to estimate FEA and Fβ as follows:41

 
image file: d5mh01052g-t1.tif(1)
 
image file: d5mh01052g-t2.tif(2)
where AEA840 and Aα763 are the absorbance at the 840 cm−1 and 763 cm−1 bands with the absorption coefficients KEA840 and Kα763 (KEA840/Kα763 ∼ 1.26), respectively, and A1277 and A1234 are the absorbance intensities at 1277 and 1234 cm−1, which correspond to the β- and γ-PVDF phases, respectively. Thus, the maximum Fβ and FEA (Table S1) are tuned in uniaxial stretching along the applied field E and stress, with the dipoles poled perpendicular to the polymer chains.42,43 A cohesive interaction of surface Te charges with β-PVDF dipoles results in softer ˙CH2 stretching bands (Fig. S6) due to damped oscillation with the enhanced effective ˙CH2 mass,14
 
image file: d5mh01052g-t3.tif(3)
where 2rdc is the damping constant, f0 and fint are the frequencies (˙CH2 stretching) before and after the doping, respectively, and c is the velocity of light. A 2rdc value of ∼5.8 × 1013 s−1 was thus estimated for the Te-doped fibers (Fig. 3b).


image file: d5mh01052g-f3.tif
Fig. 3 (a) FTIR spectra (* signifies bands common to the different phases) of electrospun NPV and PVTe fibers (Rs = 3000 rpm), and (b) variation in fractions (%) Fα/Fβ of the α-/β-phases and damping factor 2rdc with the Te content. (c) Schematic representations of the polymer chain axis (–C–C– bond) and transitional dipole moments (μ) along the a, b and c-axes with respect to the direction of the polarizer. (d) FTIR spectra of the aligned nanofibers measured using parallel and perpendicular polarized IR beams. (e) C 1s, (f) F 1s and (g) Te 3d XPS bands of the fibers.

The alignment of the dipoles in the fibers are observed as polarized FTIR vibrational bands. As shown in the schematics in Fig. 3c, if the fibers are stretched towards the c-axis along the rotation of the collector (Fig. S7), the polymer chains would be oriented along the rotation that makes an angle at the c-axis. If the c-axis lies parallel to the polarized IR beam, the component of the transition dipole moment (μc) along the c-axis induces intensified bands (Fig. 3d, full spectra provided in Fig. S8) of ω(CH2) + νas(CC) at 1404 cm−1, and νs(CC) + ω(CF2) at 1070 cm−1 in a parallel mode (ω = wagging and νas (s) = asymmetric (symmetric) stretching). In a perpendicular polarized IR beam, the component μb along the b-axis induces intensified bands of νs(CF2) + νs(CC) at 1277 cm−1 and νs(CF2) + νs(CC) at 840 cm−1, with decreased intensities of the 1404 cm−1 and 1070 cm−1 bands (as listed in Table S2).44 This indicates the preferential molecular chain orientation and improved dipole alignment at faster Rs compared to slower Rs (∼83% and ∼54%, respectively, in PVTe1.0) (Figures S9, S10, associated discussion S3). The PVDF β-phase has a 2.3% shorter d110 = 0.4285 nm (PVT1.0) in the XRD pattern (Fig. S1) compared to the virgin fibers (d110 = 0.4385 nm). The XPS C 1s bands of the ˙CH2 and ˙CF2 moieties (Fig. 3e) show higher binding energy (BE) by ΔEc = 0.35 eV and 0.20 eV, respectively, over the virgin states, with a CH2 band intensity enhancement ΔhCH2 of ∼ 5%. The reduced gap of 4.45 eV (4.60 eV before doping) between the two C 1s bands indicates Te-mediated Te-PVDF charge–dipole interactions. In Fig. 3f and g, the F1s band is also blue shifted by 0.1 eV, and the interface Te-3d3/2,5/2 bands have values 2.15 eV higher than those of the primary bands. Accordingly, the Fermi surface is pushed closer to the conduction band in a finely tuned band structure. Bare Te (Fig. S11) has 0.8 eV lower Te-3d3/2,5/2 values in the absence of interactions with ˙CH2 and ˙CF2 dipoles at the polymer chains.

At frequencies below 104 Hz, PVTe1.0 (Fig. 4a) retains a steady dielectric permittivity εr of ∼16 (11 before doping) at a well-controlled power loss (tan[thin space (1/6-em)]δ) of ≤0.02, making it usable for applications over wide bands of frequencies. A reasonably enhanced ε persists at the dipoles aligned at the Te-PVDF interfaces due to the Maxwell–Wagner–Sillars (MWS) effect.45 A stable interface regulates the charge order of the enhanced εr values that govern the piezo- and ferro-electric features.26 A peak is observed near tanδ = 2 × 106 Hz due to the rather metallized states induced on the charge-order. Fig. 4b and c display the switching response of PVTe1.0 (deposited on ITO glass) to an AC voltage applied between the probe and the substrate with a ±10 V DC bias. At positive bias, the phase-hysteresis shows that the dipoles are switched within the fibers up to a (−) 90° phase reversal to attain saturation. Under reversed fields, they reorient oppositely up to a 90° phase value in the local FE response. The enhanced magnitude of piezoelectric coefficient image file: d5mh01052g-t4.tif (13±2 pm V−1 before doping, Fig. S12), at a change of Δz in the lattice strain (amplitude response) with a change of ΔV in the applied field,8,46 indicates that they could be useful for developing self-powered devices. This value is superior (see comparison in Table S3) to those of most advanced piezoelectric materials.


image file: d5mh01052g-f4.tif
Fig. 4 (a) Permittivity (εr) and power loss (tan[thin space (1/6-em)]δ) for the NPV and PVTe1.0 nanofibers. (b) Ferroelectric hysteresis loop (PVTe1.0) showing a 180° phase reversal at switching of the dipoles, and (c) a piezoelectric butterfly hysteresis loop with d33 = 21 ± 1 pm V−1. (d) Schematic of the fabricated NG device used to generate a (e) VOC at varied pressures, (f) Isc profile at 128 kPa, (g) its electrical output voltage (VL) and power density (Pd) at varied loads (equivalent circuit in the inset), (h) its mechano-sensitivity under different pressures, and (i) its long-term mechanical stability over 4500 cycles, which was maintained even after 6 months.

The PVTe1.0 NG device depicted in the schematic in Fig. 4d was tested to validate its viability for energy harvesting applications. When subjected to an external mechanical stress, a piezo-potential develops across the device. Thus, a peak-to-peak open-circuit voltage Voc of ∼8.5 V (≤2.4 V before doping, Fig. S13a) was obtained (Fig. 4e) using an imparting machine (with a motor used as the source of impact) under a periodic applied pressure (at a frequency of 10 Hz) of up to 128 kPa, with a short-circuit current Isc of ∼1.2 μA (Fig. 4f) (∼0.53 μA before doping, Fig. S13b). Furthermore, the device demonstrates a rapid response time of ∼12 ms, as determined from the output voltage profiles under different applied pressure levels (Fig. S14). A common switching-polarity test validated the piezoelectric nature of the sign of Voc, which reverses under forward and reverse conditions under an equivalent imparting pressure of 68 kPa (Fig. S15). The instantaneous voltage VL gradually increased to 5.1 V (Fig. 4g), giving rise to a maximum power density Pd of ∼4.2 μW cm−2 at the critical load resistance RL = 0.8 MΩ across the device with an effective area of A,

 
image file: d5mh01052g-t5.tif(4)

These values represent marked improvements over previously reported values (Table S4). The mechano-sensitivity S = ∂V/∂σa shows two linear pressure response regimes (Fig. 4h): the former with mechano-sensitivity of 101 mV kPa−1 below 30 kPa, enabling the detection of pressures as low as 5 kPa; lowered to 21 mV kPa−1 in the latter regime due to the theoretical limit of the effective strain in piezoelectric materials in the high-pressure region. The mechanical durability and stability prolong safely for up to 4500 cycles, with no significant change in the amplitude of the voltage signal under a constant imparting force throughout the extended testing period, even after 6 months (Fig. 4i). This validates the long-term operational stability of the device even in harsh conditions.

The sensing performance of the device was tested towards different mechanical deformation, human motion and vibration stimuli.5,47 To demonstrate its versatility and tolerance, it was placed under mechanical deformation by bending at 30°, 60° and 90° angles. A quick bending motion (in an arc) generates Voc, which increases with increasing bending towards a maximum of 90° (Fig. 5a). The sensing behaviour of the device was systematically evaluated under realistic physiological conditions. In due flexibility and sensitivity towards bending, it converted several physiological motions (illustrated in Fig. 5b) into a Voc signal upon deformation. The pattern differs (Fig. 5c) for the bending of different synovial joints, i.e., the finger, neck, elbow and wrist. The dynamic parameters can be utilized in diagnosing various infectious diseases.7 Our device showed repeatable sensing of the thoracic pressure developed in the muscle tissues around the vocal cords during the action of coughing at the larynx (Fig. S16). The results demonstrate its stability and reproducibility towards subtle changes in instantaneous pressure, thereby allowing the detection of tactile stimuli and vocal cord vibrations for a wide range of real-time applications.


image file: d5mh01052g-f5.tif
Fig. 5 (a) Electrical outputs of the flexible NG used for mechanical sensing when bent at different angles. (b) A schematic diagram illustrating pertinent physiological gesture detections during ‘bending’ of the finger, neck, elbow and wrist, and the corresponding (c) induced electrical output signals.

To investigate its pyroelectric performance, a device architecture was engineered using PEDOT:PSS/Xyl composite electrodes. This material system was selected for its synergistic properties of high electrical conductivity, intrinsic photothermal conversion abilities, and mechanical flexibility. The primary objective of this study was to elucidate the role of the electrode composition in modulating the pyroelectric response dynamics under photothermal excitation. The pyroelectric response of the device was tested via IR heating at image file: d5mh01052g-t6.tif (Fig. 6a and b) at a 0.1 Hz switching frequency, which induces an Isc of ∼0.4 nA at a faster response time, τ of 0.12 s (Fig. 6c and d), reported so far.48 The decreasing charge polarization with rising temperature (T) induces a positive current response.49 The reversed Isc during subsequent cooling anticipates enhanced charge interactions,

 
image file: d5mh01052g-t7.tif(5)
where h is the height and w is the width of the nanowires.50 The enhanced p value of 40 μC m−2 K−1 (10 μC m−2 K−1 in the bare fibers, see Fig. S17), with a significant Voc of 145 mV (Fig. S18) demonstrates the merits the of the device as a thermal energy harvester and breathing sensor.48 The observed enhancement arises from both the device engineering and the intrinsic thermal properties of the Te-doped PVDF fibers, which effectively regulate the pyroelectric response. To further elucidate the thermal transport characteristics, the specific heat capacities (Cp) of both the undoped and doped fibers (Fig. S19) were measured to calculate the volume-specific heat CE = ρcCp (where the density ρc is ∼1.74 g cm−3).51 Assuming similar thermal diffusivity values (α) for both fibers, the thermal conductivity κ = αCE for PVTe1.0 was estimated to be significantly lower (∼0.55 W m−1 K−1) than that of NPV (0.91 W m−1 K−1).52


image file: d5mh01052g-f6.tif
Fig. 6 Pyroelectric performance of PVTe1.0 using PEDOT:PSS/Xyl gel as the electrode. (a) Input temperature profile, (b) corresponding heat rate (at 0.1 Hz switching frequency), (c) induced Isc during the heating and cooling cycle, and (d) a comparison of the pyroelectric current response-time of different pertinent materials, showing the fast response of our device. (e) Schematic diagrams of the working mechanism of the pyroelectric nanogenerator showing (i) the initial polarization state and change in the polarization state upon (ii) heating and (iii) cooling, causing alternate current flow in external circuit. (f) Comparison of the pyroelectric coefficients of different organic and inorganic materials (ref. 50 and 54–68).

The κ value is markedly reduced upon the doping of α-Te in the PVDF fibers, as it acts as a phonon-scattering centre, thus impeding heat transport.53 The restricted heat flow promotes photothermal localization, resulting in the build-up of elevated local temperatures in the PVTe1.0 nanofibers. Consequently, thermal gradients are retained more effectively, minimizing heat transfer to the surroundings. The localized thermal strain between the PEDOT:PSS/Xyl electrode and PVTe1.0 interface is therefore more pronounced compared to that in the NPV system, leading to a notable boost in the pyroelectric output. Moreover, the PEDOT:PSS/Xyl film electrode demonstrates a notably higher absolute temperature (T) and an increased temperature difference (ΔT) of 12 K, along with a higher dT/dt magnitude compared to the metal electrode (Fig. S20) (for associated discussion, see S4). This observation confirms the superior photon energy absorption efficiency of our system with organic electrodes.

Pyroelectric materials demonstrate a temperature-dependent spontaneous polarization (P) arising from their non-centrosymmetric crystal structure, which generates surface charges in the absence of an external electric field. At thermal equilibrium (dT/dt = 0), the static polarization induces bound surface charges that are neutralized by adsorbed environmental ions or electrons, resulting in no net current flow through externally connected electrodes (Fig. 6e(i)). This equilibrium persists until thermal perturbations disrupt the system. When subjected to heating (dT/dt > 0), the increased lattice thermal energy disrupts the alignment of the intrinsic electric dipoles within the material. This disordering reduces the spontaneous polarization (ΔPT < 0), thereby decreasing the surface charge density. The resulting excess compensation charges are expelled from the surfaces of the material, generating a measurable pyroelectric current in the external circuit (Fig. 6e(ii)). Conversely, during cooling (dT/dt < 0), the lattice contracts and the dipoles align, leading to an increase in the P value. This reordering elevates the surface charge density as free charges migrate from the environment to re-establish the electrostatic equilibrium (Fig. 6e(iii)). The resultant reverse current, flowing under short-circuit conditions, confirms the reversibility of the pyroelectric effect, highlighting its dependence on the direction and magnitude of the thermal fluctuations. It is worth mentioning that our device exhibits a significantly higher p value than those that have been reported for pyroelectric devices based on organic and inorganic materials (Fig. 6f).54–68

Conclusions

Te-reinforced PVDF nanofibers have been shown to exhibit enhanced pyro-electric features along with a piezo-electric response suitable for the fabrication of wearable pressure and temperature sensors. Optimized doping with 1 wt% Te (nanobeams) leads to poled, greatly promoted electroactive β/γ-phases at the interfaces, with a highly enhanced d33 value of 21 ± 1 pm V−1 at a 10 V bias. The device generates an 8.5 V signal (4.2 μW cm−2 power density), sufficient to run self-powered electronics. As a proof-of-concept, it was tested in the detection of a variety of human physiological signals, including the motions of joints during bending and the changes in thoracic pressure during coughing, showing its potential in biomedical diagnosis and voice recognition systems. Its pyroelectric features surpass those of bare PVDF fibers to detect minor thermal fluctuations, giving it great potential for real-world application in multimodal smart sensors for acoustic and medical devices.

Experimental details

Synthesis

The α-Te NCs were prepared from Na2TeO3 using a reduction reaction in hydrazine (N2H4·H2O) at 180 °C for 36 h (for associated discussion, see S1). After the reaction (2Na2TeO3 + N2H4·H2O → 2Te + 4NaOH + H2O + 2NO↑), the recovered Te sample was dispersed in a solution of PVDF in N,N-dimethylformamide (DMF) and acetone (6[thin space (1/6-em)]:[thin space (1/6-em)]4 v/v ratio) via stirring at 60 °C to obtain a homogeneous precursor of nanocolloids. Hydrazine is a mild reducing agent that yields tiny α-Te NCs dispersed in aqueous NaOH as a slurry.35 The polar liquid DMF mediates the dispersion of the α-Te NCs on the PVDF chains in the nanocolloids. The sample was injected through a syringe under a voltage of 14 kV between a metallic needle tip and an aluminum-foil-wrapped drum collector (separation: 12 cm), at which it was electrospun as nanofibers (Fig. S7). Using this method, two samples of Te-reinforced PVDF containing 0.5 wt% (PVTe0.5) and 1.0 wt% Te (PVTe1.0) were obtained (at drum rotation speeds of Rs = 500 rpm and 3000 rpm) as hybrid structures with functional properties.

Characterizations

The electroactive β/γ-PVDF phases of the oriented dipoles formed in the nanofibers were assessed using Fourier transform infrared (FTIR) spectra recorded using a Thermo Scientific, iS20 Nicolet instrument with a KRS-5 polarizer. Their crystallographic structures were analyzed from the XRD patterns recorded using an X-ray diffractometer with CuKα X-ray radiation (λ = 0.15410 nm). Field emission scanning microscopy (FE-SEM) and high-resolution transmission electron microscopy (HRTEM) were used to study the surface topologies of the bare α-Te-NCs, PVDF fibres and Te-PVDF nanocomposite fibers. The XPS analysis was performed using a Thermo Scientific K-Alpha spectrophotometer equipped with a monochromatic Al Kα X-ray source with an energy of = 1486.67 eV. The system employs a dual-beam charge compensation system that uses both a low-energy electron flood gun and an Ar ion flood gun to prevent local electrostatic charge build-up on the sample surface. The binding energies of the XPS bands were corrected with respect to the adventitious carbon C 1s band at 284.8 eV. The dielectric properties were measured using an impedance analyzer (E4990A, Keysight). The PFM (piezoelectric force microscopy) signal was studied using a Nanosurf CoreAFM system, and the electrical response of the NG device was collected using a digital storage oscilloscope (Keysight, DSOX1102G) and electrometer unit (Keithley, 6514). Other details are the same as reported previously.49

Fabrication

The NG for examining piezoelectric performance consisted of a nanofiber mat (2 × 5 cm2) sandwiched between the top and bottom electrodes of a synthetic Ni–Cu textile, which was then machined to remove air gaps so the electrodes would form contacts with the mat connected in a circuit. Finally, as reported earlier,49 the device was covered in 400-μm-thick polydimethylsiloxane using the elastomer Sylgard 184 to protect it from mechanical damage. For the preparation of the PEDOT:PSS/Xyl electrodes, a suspension of PEDOT:PSS was combined with a few drops of 4N HCl, followed by the addition of xylitol pellets in a weight ratio of 6[thin space (1/6-em)]:[thin space (1/6-em)]1 (w/w). The resulting mixture was stirred for 1 h and then placed in a sealed water bath at 100 °C for 2 h. The PEDOT:PSS/Xyl film electrodes were then fabricated by spreading the synthesized gel onto both sides of a nanofiber mat of the desired size and heating the assembly at 80 °C in air.

Author contributions

US synthesized the material, executed the experiments for material and device characterization and wrote the MS. HKM planned and carried out the device measurements, provided critical feedback and revised the MS. AK carried out PFM measurement and contributed to data analysis. SR contributed to data analysis of data obtained from TEM and XRD and revised the MS. DM planned and supervised the entire work. The manuscript was written with the contributions of all the authors.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Supplementary information: The supporting information contains associated discussions and detailed synthetic procedures, XPS spectra, a table of the calculated electroactive phases, FTIR data of the various –CH2 stretching, a table of the assignments of the vibrational bands, comparison of the degrees of the oriented dipoles, XRD patterns, the PFM response of NPV, piezoelectric and pyroelectric output responses of NPV, PVTe1.0 and the polarity reversal test. See DOI: https://doi.org/10.1039/d5mh01052g.

Acknowledgements

This work is supported by the ANRF (SERB/CRG/2020/004306), the Government of India, for financial support. US and AK are thankful to the University Grant Commission (UGC) for providing the fellowships.

References

  1. C. Chang, V. H. Tran, J. Wang, Y. K. Fuh and L. Lin, Nano Lett, 2010, 10, 726–731 CrossRef CAS PubMed.
  2. Y. Yang, W. Guo, K. C. Pradel, G. Zhu, Y. Zhou, Y. Zhang, Y. Hu, L. Lin and Z. L. Wang, Nano Lett, 2012, 12, 2833–2838 CrossRef CAS.
  3. M. Amjadi, K. U. Kyung, I. Park and M. Sitti, Adv. Funct. Mater., 2016, 26, 1678–1698 CrossRef CAS.
  4. D. Y. Park, D. J. Joe, D. H. Kim, H. Park, J. H. Han, C. K. Jeong, H. Park, J. G. Park, B. Joung and K. J. Lee, Adv. Mater., 2017, 29, 1702308 CrossRef PubMed.
  5. M. Lou, I. Abdalla, M. Zhu, X. Wei, J. Yu, Z. Li and B. Ding, ACS Appl. Mater. Interfaces, 2020, 12, 19965–19973 CrossRef CAS.
  6. R. A. Surmenev, R. V. Chernozem, I. O. Pariy and M. A. Surmeneva, Nano Energy, 2021, 79, 105442 CrossRef CAS.
  7. X. Huang, Q. Qin, X. Wang, H. Xiang, J. Zheng, Y. Lu, C. Lv, K. Wu, L. Yan, N. Wang, C. Xia and Z. L. Wang, ACS Nano, 2021, 15, 19783–19792 CrossRef CAS PubMed.
  8. H. K. Mishra, V. Gupta, K. Roy, A. Babu, A. Kumar and D. Mandal, Nano Energy, 2022, 95, 107052 CrossRef CAS.
  9. C. Ye, S. Yang, J. Ren, S. Dong, L. Cao, Y. Pei and S. Ling, ACS Nano, 2022, 16, 4415–4425 CrossRef CAS.
  10. C. M. Costa, V. F. Cardoso, P. Martins, D. M. Correia, R. Gonçalves, P. Costa, V. Correia, C. Ribeiro, M. M. Fernandes, P. M. Martins and S. Lanceros-Méndez, Chem. Rev., 2023, 123, 11392–11487 CrossRef CAS.
  11. P. Chen, C. Cheng, X. Yang, T. T. Sha, X. Zou, F. Zhang, W. Jiang, Y. Xu, X. Cao, Y. M. You and Z. Luo, ACS Nano, 2023, 17, 25625–25637 CrossRef CAS PubMed.
  12. A. Huang, Y. Zhu, S. Peng, B. Tan and X. Peng, ACS Nano, 2024, 18, 691–702 CrossRef CAS PubMed.
  13. A. Kumar, R. Haldar, E. E. Siddharthan, R. Thapa, M. Shanmugam and D. Mandal, Nano Lett., 2024, 24, 7861–7867 CrossRef CAS.
  14. H. K. Mishra, D. Sengupta, A. Babu, B. M. Pirzada, R. Sarkar, B. S. Naidu, T. K. Kundu and D. Mandal, Sustainable Energy Fuels, 2022, 6, 1625–1640 RSC.
  15. S. Mishra, L. Unnikrishnan, S. K. Nayak and S. Mohanty, Macromol. Mater. Eng., 2019, 304, 1–25 CrossRef.
  16. A. D. Phule, S. Ram, S. K. Shinde, J. H. Choi and A. K. Tyagi, J. Phys. Chem. Solids, 2018, 115, 254–264 CrossRef CAS.
  17. X. Liu, J. Ma, X. Wu, L. Lin and X. Wang, ACS Nano, 2017, 11, 1901–1910 CrossRef CAS.
  18. L. Li, M. Zhang, M. Rong and W. Ruan, RSC Adv, 2014, 4, 3938–3943 RSC.
  19. A. Morali, A. Mandal, M. Skorobogatiy and S. Bodkhe, RSC Adv, 2023, 13, 31234–31242 RSC.
  20. M. Fortunato, D. Cavallini, G. De Bellis, F. Marra, A. Tamburrano, F. Sarto and M. S. Sarto, Polymers, 2019, 11, 1096–1110 CrossRef PubMed.
  21. A. Gebrekrstos, T. S. Muzata and S. S. Ray, ACS Appl. Nano Mater, 2022, 5, 7632–7651 CrossRef CAS.
  22. L. Yang, Q. Zhao, K. Chen, Y. Ma, Y. Wu, H. Ji and J. Qiu, ACS Appl. Mater. Interfaces, 2020, 12, 11045–11054 CrossRef CAS.
  23. T. Yang, H. Pan, G. Tian, B. Zhang, D. Xiong, Y. Gao, C. Yan, X. Chu, N. Chen, S. Zhong, L. Zhang, W. Deng and W. Yang, Nano Energy, 2020, 72, 104706 CrossRef.
  24. D. Bhattacharya, S. Bayan, R. K. Mitra and S. K. Ray, Nanoscale, 2021, 13, 15819–15829 RSC.
  25. X. García-Casas, F. J. Aparicio, J. Budagosky, A. Ghaffarinejad, N. Orozco-Corrales, K. (Ken) Ostrikov, J. R. Sánchez-Valencia, Á. Barranco and A. Borrás, Nano Energy, 2023, 114, 108686 CrossRef.
  26. Y. Xue, T. Yang, Y. Zheng, K. Wang, E. Wang, H. Wang, L. Zhu, Z. Du, H. Wang, K. C. Chou and X. Hou, Adv. Sci., 2023, 10, 2300650 CrossRef CAS.
  27. N. R. Glavin, R. Rao, V. Varshney, E. Bianco, A. Apte, A. Roy, E. Ringe and P. M. Ajayan, Adv. Mater., 2020, 32, 1–22 Search PubMed.
  28. G. Rao, H. Fang, T. Zhou, C. Zhao, N. Shang, J. Huang, Y. Liu, X. Du, P. Li, X. Jian, L. Ma, J. Wang, K. Liu, J. Wu, X. Wang and J. Xiong, Adv. Mater., 2022, 34, 1–10 CrossRef.
  29. Q. Lian, W. Liu, D. Ma, Z. Liang, Z. Tang, J. Cao, C. He and D. Xia, ACS Nano, 2023, 17, 8755–8766 CrossRef CAS PubMed.
  30. J. Zhang, J. Zhang, Y. Qi, S. Gong, H. Xu, Z. Liu, R. Zhang, M. A. Sadi, D. Sychev, R. Zhao, H. Yang, Z. Wu, D. Cui, L. Wang, C. Ma, X. Wu, J. Gao, Y. P. Chen, X. Wang and Y. Jiang, Nat. Commun., 2024, 15, 7648 CrossRef CAS PubMed.
  31. H. K. Mishra, A. Jain, D. Saini, B. Mondal, C. Bera, S. Ram and D. Mandal, Phys. Rev. B, 2025, 111, 155436 CrossRef CAS.
  32. A. Chauhan, A. K. Das, D. K. Satapathy and M. Battabyal, ACS Appl. Energy Mater., 2025, 8, 6044–6055 CrossRef CAS.
  33. D. Liu, D. He and H. Huang, Appl. Phys. Lett, 2025, 126, 043505 CrossRef CAS.
  34. J. Zha, D. Dong, H. Huang, Y. Xia, J. Tong, H. Liu, H. P. Chan, J. C. Ho, C. Zhao, Y. Chai and C. Tan, Adv. Mater., 2024, 36, 2408969 CrossRef CAS PubMed.
  35. Y. Du, G. Qiu, Y. Wang, M. Si, X. Xu, W. Wu and P. D. Ye, Nano Lett, 2017, 17, 3965–3973 CrossRef CAS PubMed.
  36. M. Amani, C. Tan, G. Zhang, C. Zhao, J. Bullock, X. Song, H. Kim, V. R. Shrestha, Y. Gao, K. B. Crozier, M. Scott and A. Javey, ACS Nano, 2018, 12, 7253–7263 CrossRef CAS.
  37. C. Niu, G. Qiu, Y. Wang, Z. Zhang, M. Si, W. Wu and P. D. Ye, Phys. Rev. B, 2020, 101(205414), 6 Search PubMed.
  38. Joint Committee on Powder Diffraction Standards, X-ray Powder Diffraction File No. 00-036-1452, α-Te (International Centre for Diffraction Data, Swarthmore, PA, 2019).
  39. H. K. Mishra, A. Jain, D. Saini, B. Mondal, C. Bera, S. Ram and D. Mandal, Phys. Rev. B, 2025, 111(7), 155436 CrossRef CAS.
  40. H. S. Qian, S. H. Yu, J. Y. Gong, L. B. Luo and L. F. Fei, Langmuir, 2006, 22, 3830–3835 CrossRef CAS PubMed.
  41. J. Gomes, J. S. Nunes, V. Sencadas and S. Lanceros-Mendez, Smart Mater. Struct., 2010, 19, 065010 CrossRef.
  42. M. M. Abolhasania, S. Azimia and H. Fashandi, RSC Adv, 2015, 5, 61277 RSC.
  43. R. K. Singh, S. W. Lye and J. Miao, Polymer, 2021, 214, 123366 CrossRef CAS.
  44. D. Mandal, S. Yoon and K. J. Kim, Macromol. Rapid Commun., 2011, 32, 831–837 CrossRef CAS.
  45. A. Kumar and D. Mandal, J. Appl. Polym. Sci., 2023, 140, e53652 CrossRef CAS.
  46. A. Jalalian, A. M. Grishin, X. L. Wang, Z. X. Cheng and S. X. Dou, Appl. Phys. Lett, 2014, 104, 103112 CrossRef.
  47. D. W. Lee, D. G. Jeong, J. H. Kim, H. S. Kim, G. Murillo, G. H. Lee, H. C. Song and J. H. Jung, Nano Energy, 2020, 76, 105066 CrossRef CAS.
  48. K. Roy, S. K. Ghosh, A. Sultana, S. Garain, M. Xie, C. R. Bowen, K. Henkel, D. Schmeiβer and D. Mandal, ACS Appl. Nano Mater, 2019, 2, 2013–2025 CrossRef CAS.
  49. D. Saini, D. Sengupta, B. Mondal, H. K. Mishra, R. Ghosh, P. N. Vishwakarma, S. Ram and D. Mandal, ACS Nano, 2024, 18, 11964–11977 CrossRef CAS.
  50. C. R. Bowen, J. Taylor, E. Leboulbar, D. Zabek, A. Chauhan and R. Vaish, Energy Environ. Sci., 2014, 7, 3836–3856 RSC.
  51. Y. Zhang, M. Xie, J. Roscow, Y. Bao, K. Zhou, D. Zhang and C. R. Bowen, J. Mater. Chem. A, 2017, 5, 6569–6580 RSC.
  52. Y. Zhang, X. Zhang, L. Yang, Q. Zhang, M. L. Fitzgerald, A. Ueda, Y. Chen, R. Mu, D. Li and L. M. Bellan, Soft Matter, 2018, 14, 9534–9541 RSC.
  53. X. Cui and Y. Cai, ACS Appl. Nano Mater, 2023, 6, 17457–17463 CrossRef CAS.
  54. D. Berlincourt, H. Jaffe and L. R. Shiozawa, Phys. Rev., 1963, 129, 1009–1017 CrossRef CAS.
  55. M. H. Litt, C. H. Hsu and P. Basu, J. Appl. Phys, 1977, 48, 2208–2212 CrossRef CAS.
  56. V. S. Likhovidov, V. V. Golovanov and A. V. Vannikov, Polym. Sci. U. S. S. R., 1978, 20, 80–87 CrossRef.
  57. S. B. Lang, A. S. DeReggi, M. G. Broadhurst and G. T. Davis, Ferroelectrics, 1981, 33, 119–125 CrossRef CAS.
  58. M. H. Lee, R. Guo and A. S. Bhalla, J. Electroceram., 1998, 2, 229–242 CrossRef CAS.
  59. V. Bharti and R. Nath, J. Phys. D: Appl. Phys., 2001, 34, 667–672 CrossRef CAS.
  60. Y. Yang, J. H. Jung, B. K. Yun, F. Zhang, K. C. Pradel, W. Guo and Z. L. Wang, Adv. Mater., 2012, 24, 5357–5362 CrossRef CAS.
  61. Y. Yang, W. Guo, K. C. Pradel, G. Zhu, Y. Zhou, Y. Zhang, Y. Hu, L. Lin and Z. L. Wang, Nano Lett, 2012, 12, 2833–2838 CrossRef CAS PubMed.
  62. Q. Leng, L. Chen, H. Guo, J. Liu, G. Liu, C. Hu and Y. Xi, J. Mater. Chem. A, 2014, 1–3 Search PubMed.
  63. D. Isakov, E. de M. Gomes, B. Almeida, A. L. Kholkin, P. Zelenovskiy, M. Neradovskiy and V. Y. Shur, Appl. Phys. Lett., 2014, 104(7), 032907 CrossRef.
  64. T. Park, J. Na, B. Kim, Y. Kim, H. Shin and E. Kim, ACS Nano, 2015, 9, 11830–11839 CrossRef CAS PubMed.
  65. S. K. Ghosh, M. Xie, C. R. Bowen, P. R. Davies, D. J. Morgan and D. Mandal, Sci. Rep., 2017, 7(7), 16703 CrossRef PubMed.
  66. P. Sadhukhan, S. Q. Wu, J. I. Long, T. Nakanishi, S. Kanegawa, K. Gao, K. Yamamoto, H. Okajima, A. Sakamoto, M. L. Baker, T. Kroll, D. Sokaras, A. Okazawa, N. Kojima, Y. Shiota, K. Yoshizawa and O. Sato, Nat Commun, 2021, 12(7), 4836 CrossRef CAS.
  67. Deepak, Z. Mallick, U. Sarkar, D. Mandal and R. K. Roy, Chem. Mater., 2023, 35, 3316–3328 CrossRef CAS.
  68. R. Haldar, A. Kumar, D. Mandal and M. Shanmugam, Mater. Horiz., 2024, 11(7), 454 RSC.

Footnote

Authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.