Interface engineered V2O5-based flexible memristors towards high-performance brain-inspired neuromorphic computing

Kumar Kaushlendra and Davinder Kaur*
Functional Nanomaterials Research Laboratory (FNRL), Department of Physics and Centre for Nanotechnology, Indian Institute of Technology Roorkee, Uttarakhand, India. E-mail: davinder.kaur@ph.iitr.ac.in

Received 5th August 2025 , Accepted 8th December 2025

First published on 12th December 2025


Abstract

Bio-inspired neuromorphic computing offers a revolutionary approach by replicating brain-like functionalities in next-generation electronics. This study presents two flexible resistive memory devices fabricated using DC magnetron sputtering, D1(Nb/V2O5/Ni) and D2(Nb/NbOx/V2O5/Ni). Device D1 exhibits abrupt SET and gradual RESET switching, while D2 demonstrates fully gradual resistive switching (GRS), highly desirable for analog synaptic behavior. Mechanistically, D1 is primarily governed by oxygen vacancies, whereas D2 benefits from the synergistic interplay between oxygen vacancies and interfacial NbOx/NiO layers, confirmed by XPS depth profiling. These interfacial layers significantly enhance D2's GRS performance and synaptic fidelity. Both devices exhibit temperature-dependent control of oxygen vacancies, which dynamically increases the memory window, lowering the ON/OFF ratio. Multilevel resistive states are generated in both devices by controlling the compliance current, with D2 outperforming D1 by exhibiting a higher memory window (∼552) and exceptional endurance beyond 7000 cycles. Moreover, both devices effectively replicate biological synaptic functions such as LTP and LTD. However, D2 also mimics complex neural dynamics, including spike time-dependent and rate-dependent plasticity. Simulation of D2's artificial neural network demonstrates ∼86.75% excellent accuracy level, attributed to its linear, symmetric analog weight modulation and multiple conductance states. These results highlight the potential of V2O5-based devices for high-performance neuromorphic computing.



New concepts

This study demonstrates a breakthrough in flexible neuromorphic device engineering through the design of interface-engineered V2O5-based memristors employing a dual-interface structure with NbOx and NiOx layers. This configuration enables precise control of oxygen vacancy migration, resulting in fully gradual resistive switching behavior essential for analog synaptic operation unlike conventional abrupt-switching V2O5 or single-layer devices. The integration of both engineered and self-formed interfacial oxides modulates electric field distribution and stabilizes filamentary conduction, achieving superior switching uniformity, multilevel conductance states, and mechanical flexibility on nickel substrates. This approach provides a fundamental differentiation from existing research by simultaneously realizing flexibility, analog tunability, and endurance in a single device architecture fabricated via scalable DC magnetron sputtering. Beyond device performance, the conduction analysis reveals a transition from Ohmic to trap-controlled space-charge-limited conduction, offering new physical insight into defect-mediated analog switching in oxide-based systems. The successful emulation of synaptic behaviors such as LTP, LTD, STDP, and SRDP, together with ANN-based image recognition, establishes this concept as a significant advance in nanoscience for developing energy-efficient, flexible, and brain-inspired neuromorphic hardware.

1. Introduction

Conventional Von Neumann systems have limitations in the emerging field of artificial intelligence, as the physical separation between memory and the computing unit results in low processing speed and high power consumption.1–7 These restrictions make it more difficult for them to satisfy the rising need for enormous processing power.8–11 Researchers have developed an artificial electronic synapse that integrates memory, computation, and signal detection into a single unit, mimicking the parallel processing capability of the human brain. Neural impulses are transmitted between pre- and post-synaptic terminals via electrical and chemical signals in a biological synapse. Neurotransmitters (Ca2+, Na+) facilitate information transmission by controlling the synaptic weight of biological memory synapses (BMS).8,12,13 Artificial memory devices have a comparable bionic structure, which includes top and bottom electrodes resembling pre- and post-synapses and vacancies, ions, and electrons like neurotransmitters in BMS.14–17

Filamentary-based resistive random-access memory (RRAM), with fast switching (∼1.2 ns), low power consumption (∼1.0 fJ), high-density integration via crossbar arrays, and excellent endurance and retention, exhibits synaptic behavior. Its metal–insulator–metal (MIM) structure enables easy fabrication.11,12,18–21 Bipolar resistive switching (BRS) can be categorized into abrupt resistive switching (ARS) and gradual resistive switching (GRS) modes. ARS exhibits a high ON/OFF ratio and strong retention but suffers from poor endurance and limited synaptic plasticity. In contrast, GRS offers multiple conductance states with smooth and controllable modulation, making it more suitable for neuromorphic computing applications. This analog-like behavior enhances the linearity of long-term potentiation (LTP) and depression (LTD), enabling precise weight updates required for spike-timing-dependent plasticity (STDP). A. K. Jena et al. reported that gradual switching reduces circuit complexity, while M. Ismail et al. demonstrated that improved LTP/LTD linearity significantly enhances computational accuracy in neuromorphic systems.9,22 Moreover, GRS devices generally operate at lower voltages and currents, leading to reduced power consumption and improved reliability. Their superior endurance, uniformity, and scalability make them promising candidates for large-scale neuromorphic hardware. The gradual modulation of resistance states is typically governed by the progressive formation and migration of oxygen vacancies, contributing to stable analog performance. Fabricating ARS devices is more straightforward than fabricating GRS devices.23 To convert ARS to GRS, researchers have optimized key parameters using various techniques. Likewise, a thermally enhanced layer is used to regulate vacancy distribution, thereby modulating conductivity, while interfacial layers restrict charge carrier mobility, both of which facilitate the transition from ARS to GRS.20,24–27

Bilayer memristor devices are promising candidates for mimicking biological synapses due to their low power consumption and ability to emulate key synaptic behaviors such as long-term potentiation (LTP), long-term depression (LTD), paired-pulse facilitation (PPF), spike-timing-dependent plasticity (STDP), and spike-rate-dependent plasticity (SRDP).28,29 Several studies have demonstrated the effectiveness of interfacial engineering in enhancing synaptic performance. For example, Ryu et al. reported bilayer ReRAM devices using HfO2 as interfacial layers, achieving excellent analog switching and synaptic characteristics.30 Wan et al. introduced SiO2 and graphene oxide layers to realize analog switching and neuromorphic behavior, while Ismail et al. demonstrated gradual switching in TiO2/HfO2 bilayers using a tailored pulse scheme.28,31 In our work, we successfully achieved multilevel analog switching at low operating voltages and emulated essential synaptic functions in a V2O5-based ReRAM device by introducing a NbOx interfacial layer, further validating the effectiveness of bilayer structures in neuromorphic computing applications. Niobium-oxide (NbOx) is an attractive option, as the material exhibits rich and complex electrically memristive behaviors that can vary based on the film composition, structure (e.g., crystallinity), and mode of electrical operation. In addition, it creates an oxygen vacancy gradient in the oxide layer, which enhances the switching performance. Furthermore, various transition metal oxides such as TiO2, ZnO, and NiO have demonstrated resistive switching (RS) characteristics. Although these memristors exhibit stable switching behavior and good scalability, they often suffer from limited analog tunability. In contrast, vanadium pentoxide (V2O5), a layered oxide with unique optical, electrical, and structural properties, offers significant potential for achieving gradual, analog-type switching behavior, making it highly promising for neuromorphic computing applications.32–39 Zhenni Wan et al. reported resistive switching in V2O5 with the effect of annealing temperature and abrupt SET/RESET behavior.40 However, researchers have not fully explored the tunability of oxygen vacancies and interfacial layers for achieving gradual SET/RESET, which holds significant potential for neuromorphic applications. Gradual resistive switching (GRS) behavior facilitates smooth and predictable variations in conductance, essential for implementing analog weight updates in neuromorphic systems. Nevertheless, achieving dependable GRS in single-layer systems sometimes proves difficult due to constraints in stability and variability. Bilayer memristor architectures have been developed to address this limitation. These architectures employ a regulated interfacial layer to govern the migration of oxygen vacancies and the development of filaments. These bilayer configurations maintain the slow switching characteristics of GRS while enhancing endurance, homogeneity, and analog tunability, all of which are crucial for precise synaptic emulation. Thus, the present study examines V2O5-based bilayer memristor devices, including a NbOx interfacial layer to achieve robust and reproducible GRS behavior suitable for neuromorphic computing.4,9,33 The diffusion coefficient and mobility of oxygen vacancies in active materials and vacancy-driven filament diameter modulation play a crucial role in synaptic measurements and facilitate analog switching behavior, ensuring stable endurance and long retention.41–45

Today's world is driven by an unending need for wearable, flexible, implanted sensors, rollable screens, wearable health diagnostic systems, and portable electronics. Thus, researchers focus on developing low-cost, multifunctional, and mechanically stable flexible electronics. Stainless steel, metal foils, and polyimide are examples of flexible substrates that are being thoroughly studied for the development of RRAM devices.36,46 Flexible metal foils garner significant interest due to their high-temperature applications. Flexible Ni (nickel) foil has shown great promise due to its additional features, including low cost, magnetostrictive property, broad manufacturing area capability, and high fracture strength. However, attributed to the temperature imbalance that arises from interfacial chemical effects, the direct deposition of an active V2O5 layer on Ni foil produces an interfacial layer.8,13 Nb was selected as the top electrode due to its high conductivity and reactivity. This layer enhances the switching performance of the device, making it a promising candidate for RRAM memory applications.

This work reports the development of two flexible RRAM devices, D1 and D2, fabricated on nickel substrates via DC magnetron sputtering, designed for neuromorphic computing inspired by biological synaptic behavior. Device D1 exhibits abrupt SET and gradual RESET transitions, enabling multilevel current–voltage (IV) characteristics under compliance control. In contrast, device D2 demonstrates fully gradual resistive switching (GRS) behavior in both SET and RESET states, attributed to introducing an engineered NbOx interfacial layer. This layer improves switching uniformity and enhances analog modulation, an essential feature for high-fidelity synaptic emulation. It emulates key synaptic behaviours such as LTP, LTD, STDP, and SRDP. These attributes collectively position the V2O5-based ReRAM device D2, exhibiting superior endurance and mechanical flexibility, as a promising candidate for next-generation wearable neuromorphic electronic devices.

2. Experimental section

2.1. Materials and fabrication

The magnetron sputtering technique was used to fabricate the MIM stack of Nb/V2O5 and Nb/NbOx/V2O5 over a flexible Ni substrate using a commercial Vanadium, Niobium, sputtering target (5 mm thick and 2-inch diameter).46–49 The Ni substrate was ultrasonically cleaned thoroughly before thin film deposition. Initially, the V2O5 layer was deposited directly onto the flexible Ni substrate under optimized sputtering conditions, followed by the deposition of an ultrathin NbOx interfacial layer to improve RS performance. The deposition parameters were carefully tuned and summarized in Table 1. After that, the top electrode of Nb was deposited using a shadow mask of diameter 100 µm, and the resulting devices with oxygen vacancies are named D1 (Nb/V2O5/Ni) and D2 (Nb/NbOx/V2O5/Ni). In addition, the V2O5/Ni heterostructure was annealed at 450 °C in a pure argon atmosphere for 35 minutes to create oxygen vacancies. Nb top-electrode was deposited thereafter, and the resulting devices had oxygen vacancies in devices D1 and D2.
Table 1 Optimized sputtering parameters of the as-fabricated V2O5-based RRAM devices
Sputtering parameters V2O5 (Switching layer) Nb NbOx (Top - electrode)
Substrate temperature 450 °C RT 550 °C
Target-substrate distance 5 cm 5 cm 5 cm
Pressure (Torr) 2.06 × 10−6 3.1 × 10−6 2.2 × 10−6
Deposition-pressure 7 mTorr 6 mTorr 6 mTorr
Sputtering gas Ar + O2 Ar Ar + O2
Sputtering power 130 W 50 W 130 W
Deposition time 3.5 min 40 s 30 s


2.2. Formation mechanism of nano-rod like – V2O5 structures

During sputtering, Ar+ ions transport V and O atoms to the substrate, forming initial nucleation sites for V2O5 thin films. In the early stages, the films grow layer-by-layer until a critical thickness is reached. Beyond this point, the V2O5 structure transitions into a nanoworm morphology, driven by the intrinsic nature of its 2D layered crystal structure, which is prone to curling and edge dislocations. Localized heating during deposition further contributes by introducing slight curvatures and dislocations that serve as energetically favorable sites for vertical growth. As deposition continues, these curved regions act as nucleation sites, promoting anisotropic growth and resulting in the formation of nano rod-like structures. The final entangled morphology arises from variations in local growth conditions, filament curvature, and strain, leading to the development of the observed nanorod features.

2.3. Material characterizations

The crystalline orientation and phase composition of the V2O5/Ni structure were investigated using X-ray diffraction (XRD) with a Cu Kα radiation source (λ = 1.54 Å) at a grazing incidence angle of 2°. Surface morphology and cross-sectional film thickness were characterized using field emission scanning electron microscopy (FESEM), providing high-resolution imaging of the heterostructure architecture. Elemental composition and atomic distribution were determined via energy-dispersive X-ray spectroscopy (EDX), integrated with FESEM, to confirm the presence and stoichiometry of constituent elements. X-ray photoelectron spectroscopy (XPS, PerkinElmer Model 1257) was employed to investigate the chemical states and bonding environments of elements in the active layers. The electrical and synaptic properties of the fabricated D1(Nb/V2O5/Ni) and D2(Nb/NbOx/V2O5/Ni) structure were measured using a semiconductor characterization system (SCS 4200, Keithley) and Keithley source meter 2400, 2450. The magnetic hysteresis loops were recorded using a physical property measuring system (PPMS, DYNACOOL).

3. Results and discussion

Fig. 1a schematically illustrates the DC magnetron sputtering process for fabricating V2O5-based ReRAM devices (D1 and D2) on a flexible Nickel (Ni) substrate. The X-ray diffraction (XRD) pattern of Nb2O5/V2O5/Ni in Fig. 1b reveals diffraction peaks at 2θ = 15.44°, 20.04°, 21.7°, 24.92°, 26.18°, 31.06°, 32.5°, 34.42°, 38.14°, 41.26°, 45.46°, 47.38°, 49.6°, 53.66°, 55.68°, 61.16°, and 62.16° angles. These peaks are attributed to (200), (001), (101), (201), (110), (301), (011), (310), (401), (002), (411), (600), (112), (220), (021), (321), and (710) diffraction planes, confirming the successful synthesis of V2O5 with orthorhombic crystal structure (JCPDS Card No: 89-2482).34,50,51 Additional peaks at 2θ = 22.8°, 28.8°, and 29.3° correspond to the (001), (180), and (200) planes of orthorhombic Nb2O5 (JCPDS No. 30-0873). Furthermore, the diffraction peaks observed at 44.18° and 51.3°, corresponding to the (111) and (200) crystallographic planes, respectively, are in good agreement with the standard diffraction pattern of Ni substrate (JCPDS Card No: 01-087-0712).52 The alignment of diffraction peaks with established data underscores the structural precision and reliability of the fabrication process.
image file: d5nh00562k-f1.tif
Fig. 1 (a) Schematic diagram for DC magnetron sputtering and as-fabricated ReRAM devices, namely Ni/V2O5/Nb (D1), and Ni/V2O5/NbOx/Nb (D2), along with their digital photographs, surface FE-SEM images. (b) XRD pattern of Nb2O5/V2O5/Ni. (c) XPS full survey spectrum for V2O5/Ni. (d and e) XPS high-resolution spectra of O 1s and V 2p regions. (f) XPS depth profile. (g) Raman spectra.

X-ray photoelectron spectroscopy (XPS) was employed to analyze the chemical composition and electronic bonding states of the V2O5 thin film. The survey spectra confirmed the presence of vanadium (V), oxygen (O), and carbon (C), as shown in Fig. 1c.53 The C 1s peak at 284.7 eV, attributed to adventitious carbon contamination, was utilized as an internal reference for binding energy calibration. High-resolution XPS spectra of the V 2p region revealed two distinct peaks at binding energies of 524.1 eV and 517.3 eV, corresponding to the V 2p1/2 and V 2p3/2 spin–orbit components, respectively (Fig. 1d). Quantitative elemental analysis was performed by integrating the areas under the deconvoluted peaks. Further insight into the oxidation states of vanadium was obtained through deconvolution of the V 2p3/2 peak, which revealed components at 516.4 eV and 517.4 eV, corresponding to V4+ and V5+ oxidation states, respectively. The estimated V4+/V5+ ratio of ∼10.2% confirms the existence of oxygen vacancies within the V2O5 active layer, which plays a critical role in modulating the resistive switching behavior of the device.54 The O 1s core-level spectrum in Fig. 1e exhibited a primary peak at 530.2 eV, indicative of stoichiometric V2O5, and an additional peak at 531.6 eV, which is associated with V–O bonding involving non-lattice (dangling) oxygen atoms, signifying oxygen vacancy-related states. Fig. 1f presents the X-ray photoelectron spectroscopy (XPS) depth profile of the Nb/NbOx/V2O5 multilayer structure deposited on a Ni substrate. The depth profile of elemental concentration confirms the existence of a NbOx layer at the Nb–V2O5 interface and formation of an additional NiOx interfacial layer at the V2O5–Ni interface. The comprehensive depth-resolved elemental analysis reveals a well-defined multilayer structure comprising Nb/NbOx/V2O5/NiO sequentially stacked on the flexible Ni substrate. Raman spectra in the 100–1200 cm−1 wave number range were used to analyze the fabricated vanadium pentoxide (V2O5), as shown in Fig. 1g. The crystalline and layered structure of V2O5 is confirmed by wave numbers 145, 197, 284, 304, 407, 490, 690, and 994 cm−1.55 The Raman peaks at 994 and 490 cm−1 represent the doubly coordinated oxygen bond between vanadium and oxygen (V[double bond, length as m-dash]O) and the stretching mode of V–O3–V, respectively. The vibration at 145 is due to the VO5–VO5 mode. The 284 and 407 cm−1 peaks confirm the bending vibration mode of V[double bond, length as m-dash]O.41,42

High-resolution transmission electron microscopy (HRTEM) was systematically performed to elucidate the nanostructural features of the V2O5 active layer integrated within the RRAM architecture. HRTEM images (Fig. 2a–c) reveal a nanosheet-like morphology that facilitates multiple ionic conduction paths, promoting robust filament formation. A zoomed-in region of Fig. 2b and c at 20 and 10 nm scale was selected and analyzed using fast Fourier transformation (FFT) to assess the lattice orientation. The resulting FFT pattern (right side of inset Fig. 2e and f) displays well-defined diffraction spots corresponding to the (110) and (010) planes of orthorhombic V2O5. Inverse FFT (IFFT) images of the selected region are further generated (left side of inset Fig. 2e and f), highlighting the coherence and periodicity of the atomic planes. HRTEM images in Fig. 2e and f confirm the layered crystalline structure, with interplanar spacings of d110 = 0.341 nm and d200 = 0.575 nm identified via FFT and inverse FFT analysis. Additionally, the selected area electron diffraction (SAED) pattern (Fig. 2d) shows polycrystalline diffraction rings corresponding to the (200), (110), (301), (310), (411), (112), (021), (412), (710), and (712) lattice planes, further confirming the phase purity and crystallinity of V2O5. Energy-dispersive X-ray spectroscopy (EDX) elemental mapping (Fig. 2g–i) reveals a homogeneous spatial distribution of vanadium (V) and oxygen (O) elements across the examined region.


image file: d5nh00562k-f2.tif
Fig. 2 (a–c) Surface HRTEM image showing porous V2O5 nanosheets recorded at different magnifications. (d) SAED pattern. (e and f) HRTEM image showing a zoomed-in region of V2O5 nanosheets with resolved planes (Inset shows the FFT and IFFT image). (g–i) EDX elemental mapping of V2O5.

Field-emission scanning electron microscopy (FE-SEM) images of the V2O5 layer at different magnifications (25 kx, 50 kx, and 100 kx), as shown in Fig. 3a and b, reveal a uniform nanorod-like morphology of the active switching layer. This well-aligned architecture suggests a high degree of surface regularity, which is advantageous for consistent filamentary switching behavior in resistive memory applications. Fig. 3c presents the FESEM cross-sectional image revealing the structural configuration of the device. The energy dispersive X-ray spectroscopy (EDX) spectrum in Fig. 3c confirms the presence of V and O with atomic percentages of 30.82% and 69.18%, respectively, values in close agreement with the ideal stoichiometry of V2O5. The uniform elemental distribution of V and O across the active layer is further verified through elemental electron imaging, as shown in Fig. 3d–g. The combined XRD, Raman, XPS, FE-SEM, and HRTEM results confirm the successful synthesis of V2O5.


image file: d5nh00562k-f3.tif
Fig. 3 (a and b) FE-SEM images of V2O5 layer illustrating the nanorod-like surface morphology. (c) Shows the FESEM cross-section image of NbOx/V2O5/Ni. (d) EDX analysis and elemental composition of V2O5 thin film. (e–g) EDX elemental mapping of vanadium (V) and oxygen (O) elements.

To investigate the role of interfacial engineering in resistive switching characteristics, two MIM-structured RRAM devices (D1 and D2) were designed and fabricated using vanadium pentoxide (V2O5) as the active functional layer.12 Both devices employ a nickel (Ni) substrate as the bottom electrode (BE) and niobium (Nb) as the top electrode (TE); however, device D2 is uniquely engineered with an ultrathin NbOx interlayer between the V2O5 and Nb layers (Fig. 4a). The resulting configurations are denoted as D1(Nb/V2O5/Ni) and D2(Nb/NbOx/V2O5/Ni). The electrical characterization (Fig. 4b and c) demonstrates a significant disparity in switching behavior between the two devices. Device D1 exhibits a highly abrupt SET transition followed by a relatively gradual RESET process, demonstrating asymmetric BRS behavior governed by stochastic filament dynamics and electric field. This abrupt switching is often associated with uncontrolled filament formation and rupture, leading to variability in device performance. However, the D2 demonstrates a continuous, gradual conductance modulation during both SET and RESET characteristics. This gradual switching behavior suggests enhanced control over the migration of oxygen ions and vacancy-mediated filament. The improved switching uniformity and reversibility in D2 are attributed to the functional role of the inserted NbOx interfacial layer, which acts as an energy barrier. The NbOx layer redistributes the local electric field at the V2O5 interface, smooths potential gradients, and suppresses abrupt filament growth, facilitating controlled resistive switching.


image file: d5nh00562k-f4.tif
Fig. 4 (a) Schematic presentation of electrical measurement. (b and c) IV characteristics of Ni/V2O5/Nb (D1) and Ni/V2O5/NbOx/Nb (D2) (inset shows an analytical model for SET and RESET of D1 and D2).

Structurally, the NbOx interlayer stabilizes the growth of the conductive filament and promotes improved charge injection symmetry by introducing a customized energy environment through interfacial engineering. This design successfully lowers switching variability, increases device durability, and improves analog tuning capabilities, all of which are critical features for multilevel memory systems and neuromorphic computing applications. Moreover, in both D1 and D2 configurations, the Ni bottom electrode chemically interacts with the V2O5 active layer during the film deposition, leading to interfacial modifications. XPS depth profiling reveals the emergence of a thin interfacial nickel oxide (NiOx) layer at the V2O5/Ni interface. This spontaneous reaction is driven by interfacial oxygen exchange, resulting in a modified electrode interface that can influence barrier height and interfacial defect chemistry. The NiOx layer may act as an additional resistive element, further modulating the bottom interface's electrochemical properties and impacting the overall switching kinetics. Dual-interface engineering via intentional NbOx insertion at the top and spontaneous NiOx formation at the bottom offers a synergistic pathway for fine-tuning resistive switching behavior in V2O5-based RRAM devices. These findings underscore the critical importance of atomic-scale interface control in governing switching reliability, endurance, and performance consistency in emerging memory technologies.

When a positive voltage is applied to the Nb top electrode (TE) while the Ni bottom electrode (BE) is grounded in devices (D1), Oxygen ions (O2−) migrate toward the TE. This ion migration creates oxygen vacancies in the V2O5 and NiOx layers, both acting as active switching layers.53 The oxygen vacancies tend to align and drift downward, forming a conductive filament. This filament formation results in an abrupt transition from a high-resistance state (HRS) to a low-resistance state (LRS), representing the SET process as shown in Fig. 4b and supported by an analytical model that explains the SET process (inset of Fig. 4b). Conversely, applying a negative voltage to the TE causes oxygen vacancies in D1 to migrate away from the filament (Fig. 4b). As a result, the device gradually transferred from LRS to HRS, corresponding to the RESET process, as illustrated in the inset of Fig. 4b. In contrast, Device D2 exhibits gradual SET and RESET switching when positive and negative voltage is applied to TE. This behavior is described as the introduction of an additional interfacial NbOx layer. The observed difference arises from variations in Gibbs’ free energy, activation energy, and diffusion coefficients of NbOx, NiO, and V2O5, which govern the mobility of the oxygen vacancies, leading to non-uniform filament formation.28,56 Upon the application of the positive bias, oxygen vacancies accumulate asymmetrically near the conductive filament, resulting in an increase in filament diameter, as depicted in the inset of Fig. 4b. Instead of undergoing abrupt rupture, the filament experiences partial and gradual dissolution, initiated by the outward migration of oxygen vacancies. This controlled degradation reduces filament conductivity progressively, thereby facilitating a smooth transition from a low-resistance state (LRS) to a high-resistance state (HRS), as illustrated in the inset of Fig. 4c. Furthermore, device D2 exhibits lower conductivity and a higher ON/OFF ratio due to an additional interfacial layer NbOx, which influences charge transport and resistive switching behavior.

Furthermore, Fig. 5a and b illustrates the manifestation of the multilevel switching (MLS) behavior as a direct function of varying compliance currents during the RS process. This behavior stems from the dynamic modulation of the filamentary conduction pathway within the switching dielectric layer, governed by the magnitude of the compliance current. A higher compliance current induces a more pronounced soft breakdown in the dielectric medium, promoting enhanced ion migration and accumulation, leading to a more robust and wider conductive filament. This increase in filament diameter reduces the effective resistance in the low-resistance state (LRS) due to the enlarged conductive cross-section. Moreover, the rise in filament density reduces charge carrier hopping distances, facilitating more efficient charge transport across the filament. Consequently, the compliance current modulates the LRS and high-resistance state (HRS) levels, reflecting the filament's morphological evolution and its impact on the overall resistance states. The endurance and retention characteristics of devices D1 and D2 under varying compliance currents are presented in the inset of Fig. 5a and b, respectively. These plots highlight the multilevel endurance behavior, demonstrating the stability and reproducibility of discrete resistance states over repeated switching cycles. Notably, the device D1 exhibits endurance of 7000 cycles and retention of 6000 s, whereas D2 demonstrates endurance of 5000 cycles and retains data for up to 4500 s. Fig. 5g illustrates the re-plot of flexible RRAM devices D2 having IV characteristics on a double logarithmic (log[thin space (1/6-em)]I–log[thin space (1/6-em)]V) scale to provide a deeper understanding of the underlying conduction mechanism. The analysis reveals that the conduction process is primarily governed by the space charge-limited conduction (SCLC) mechanism, which implies that electron trapping and detrapping in oxygen-related traps/defects are intimately connected to the conduction process. In LRS, the slope of the log[thin space (1/6-em)]I–log[thin space (1/6-em)]V curve is approximately 1.02 across the entire voltage range, indicating the Ohmic behavior (IV), a characteristic of thermally generated free carrier conduction.12 In the high-resistance state (HRS), the device exhibits three distinct conduction regions. In the low-voltage region (0 < V < 0.15 V), the slope of 0.902 indicates Ohmic conduction, suggesting that the current is limited by thermally generated free carriers. As the voltage increases, the conduction mechanism transitions to Schottky emission in the intermediate region (0.15 < V < 0.35 V), which is confirmed by the linear behavior observed in the ln(I) versusV graph plotted in Fig. 5h (The detailed analysis is provided in the SI).23,57 At higher voltages, the slope increases to approximately 4.55, indicating a transition from the trap-filled limit (TFL) to the Child's law region of space-charge-limited conduction (SCLC), where the current follows the relation IVn (n > 2). These progressively increasing slope values reflect a gradual trap-filling process, in which charge transport becomes increasingly constrained by the availability of free carriers and trap states. Therefore, the dominant conduction mechanism in device D2, particularly in the HRS regime, is attributed to trap-controlled SCLC, governed by carrier injection and transport through defect states within the switching layer. Furthermore, the combined effects of temperature and compliance current (CC) on the IV characteristics of devices D1 and D2 were investigated, as shown in Fig. 5c and d. Fig. 5e and f illustrate the temperature-dependent behavior alone, where an increase in conductivity with rising temperature was observed. This phenomenon can be attributed to the enhanced generation and mobility of oxygen vacancies, which facilitate charge transport within the switching layer.


image file: d5nh00562k-f5.tif
Fig. 5 (a and b) Multilevel IV behavior in devices D1 and D2 under compliance current (Inset shows the endurance and retention characteristics of D1 and D2). (c and d) combined effect of Temperature-dependent and CC I–V response in D1 and D2 (inset shows device-to-device measurement of D1 and D2). (e and f) Temperature dependent IV response with constant CC (g) log[I] versus log[V] curve for D2. (h) (b) Schottky conduction of the HRS (high-voltage region of negative bias data) by fitting. (i) Memory window versus temperature response of D1 and D2.

The device-to-device variability measurements for devices D1 and D2 are presented in the inset of Fig. 5c and d, respectively. The device D2 demonstrated superior stability, attributed to the valence change mechanism (VCM) and an additional interfacial NbOx layer. This interfacial layer plays a crucial role in enhancing the RS mechanism by stabilizing the diameter of conductive filaments, thereby improving the overall reliability and performance of the device. In devices D1 and D2, the memory window initially increases with temperature, reaching a maximum at 335 K. This enhancement is attributed to the thermally activated generation of oxygen vacancies, which facilitates the formation of stable conductive filaments, thereby improving the ON/OFF ratio. Within the 300–335 K range, the gradual increase in oxygen vacancy concentration leads to a wider memory window. However, beyond 335 K, particularly between 335–350 K, the memory window decreases due to excessive vacancy accumulation. This surplus disrupts the filament formation dynamics, leading to instability in resistive switching and a decline in performance (Fig. 5i). Thus, while oxygen vacancies initially enhance switching behavior, their overaccumulation at higher temperatures negatively impacts device reliability.

Fig. 6a illustrates the key components of a biological synapse in the human brain, including dendrites, axons, pre- and post-synaptic neurons, the synaptic cleft, and neurotransmitters. In the artificial device, the top and bottom electrodes correspond to pre- and post-synaptic neurons, while electrons, oxygen ions, and vacancies function analogously to neurotransmitters. The device's electrical conductivity variations, representing weak and strong conductive states, correlate with the modulation of synaptic weight (strengthening and weakening) in biological memory systems (BMS).9,13 Fig. 6b, the sequential application of five positive voltage pulses to the top electrode of device D2 leads to a gradual increase in current. In comparison, five negative voltage pulses decrease the previously incremented current. This current and voltage versus time response is related to synaptic functions such as potentiation and depression of synaptic weight.


image file: d5nh00562k-f6.tif
Fig. 6 (a) Visual representation of synaptic activity between the pre-synaptic and post-synaptic neurons and ionic memristor interface on flexible Ni substrate. (b) IV characteristics of multiple consecutive voltage sweeps of device D2. (c and d) LTP and LTD cycles of D2 and D1. (e) Comparison of LTP and LTD for D1 and D2. (f and g) LTP and LTD at different amplitudes and pulse widths in D1 and D2.

To investigate electronic synaptic behavior, 40 identical positive (1.1 V, 1 ms) and negative (−1.1 V, 1 ms) voltage pulses were applied to devices exhibiting gradual (D1 and D2) switching. The LTP and LTD cycles of devices D1 and D2 are shown in Fig. 6c and d. D2 achieves higher LTP/LTD cycle counts than D1, highlighting the enhanced filament control and repeatability enabled by the interfacial engineering. Fig. 6e shows conductance modulation in D1 and D2, where gradual and linear LTP/LTD occurs due to defects, vacancies, ion migration, and the interfacial layer, which is beneficial for neuromorphic applications. The conductivity of D2 is lower than that of D1 due to the interfacial layer NbOx.58 Fig. 6f and g shows the LTP and LTD measurements for device D2 under different pulse width trains and amplitudes, leading to more efficient synaptic weight control and indicating synaptic amplitude-dependent plasticity.

After successfully emulating a biological synapse, we investigated a fundamental synaptic learning mechanism, paired-pulse facilitation (PPF). This process involves the modulation of synaptic conductance by the cumulative effect of consecutive input pulses and stimuli, thereby enabling the memristive device to mimic short-term neuromorphic functionalities such as synaptic adaptation and temporal filtering. For the PPF measurements, two consecutive voltage pulses of 3V amplitude and 10 ms duration were applied, with the inter-pulse interval (Δt) varied from 0.1 to 15 ms (Fig. 7a). A significant increase in output current was observed in response to the second stimulus relative to the first, indicating facilitation. This enhancement occurs because the synaptic response from the first pulse does not fully decay before the arrival of the second pulse. When the second stimulus arrives within this short recovery window, it results in an elevated conductance response analogous to biological synaptic behavior. The observed increase in current reflects a temporary modulation of the device's conductance induced by the applied pulse sequence. This behavior is attributed to the migration of oxygen vacancies (Vo) towards the conductive filament region, even after removing the potentiating pulse.9,58 When the interval between two successive pulses is shorter than the relaxation time of the oxygen vacancies, the residual ionic redistribution from the first pulse enhances the effect of the second, leading to a higher conductance state. Conversely, when the inter-pulse interval exceeds the relaxation time, the vacancies return to their equilibrium positions, and the facilitation effect saturates or diminishes. The PPF can be calculated as

 
image file: d5nh00562k-t1.tif(1)
where Af and Ai are the post-synaptic current readouts after the first and second pulses, respectively.


image file: d5nh00562k-f7.tif
Fig. 7 (a) Schematic illustration of the PPF rule of bio-synapse between pre- and post-synaptic neurons and diameter modulation of the artificial device. (b) PPF ratio measurement for device D2 using a pulse of 3 V and 10 ms. (c and d) STDP and SRDP for device D2.

Fig. 7b illustrates that the PPF value progressively decreases as the interval between consecutive pulses increases. At shorter intervals, the diameter of the conductive filament formed by the first pulse partially persists, resulting in an enhanced response to the second pulse and a larger change in current. In contrast, the filament partially dissolves at longer intervals before the second pulse arrives, leading to a minor increase in filament diameter and a lower current response, thereby reducing the PPF ratio value. The extracted fitting parameters are c1 = 26, c2 = 103.5, τ1 = 1.519 ms, and τ2 = 468.69 ms, here, τ1 corresponds to the fast relaxation process, typically associated with short-term synaptic response dynamics, while τ2 represents the slower relaxation process, linked to longer-lasting synaptic modulation. The obtained values are consistent with the characteristic time scales reported in memristive synapses, further validating the reliability of our measurements. The increasing and decreasing characteristics of PPF with time intervals closely mimic biological synaptic behavior.

To study the SRDP and STDP, which are fundamental associative learning rules governed by Hebbian plasticity, it is crucial to analyse the role of frequency-dependent and temporally structured input stimuli in modulating synaptic weight. In the case of SRDP, synaptic weight adaptation can be achieved by varying the average firing rate of input pulses. Conversely, the relative timing between pre- and post-synaptic spikes determines STDP-induced synaptic weight modulation. In this experiment, SRDP has been simulated by varying the time interval between voltage write pulses (+1.5 V) and erase pulses (−2.0 V), both applied with a pulse width (PW) of 1 ms. The read pulses were delayed by 1 ms relative to the write/erase pulse train. Fig. 7c displays the changes in synaptic weight (Δω) as a function of the time interval (Δt), calculated using the formula.

 
image file: d5nh00562k-t2.tif(2)
where Kf = current value at final Vread, Ki = Current value at initial Vread.

The weight changes for both LTP and LTD were well-fitted with the equation:

 
image file: d5nh00562k-t3.tif(3)
where, τ is the time constant for LTP and LTD in SRDP.

For STDP, similar experiments have been measured utilizing an adjustment in pulse mode as shown in Fig. 7d, which is explained by the relationship between time interval (Δt) and synaptic weight (Δω). The gap in time between them is described as

 
Δt = tpretpost (4)
where tpre and tpost were the pulse times that came to the post-synaptic and pre-synaptic electrodes. A pair of pulses was applied to the pre-synaptic electrode (1.1 V, 1 ms) and post-synaptic electrode (−1.1 V, 1 ms), and initial and final Vread (0.5 V, PW = 1 ms). The synaptic weight connection increases, indicating the relationship between the two synapse connections is potentiated. However, synaptic weight decreases, indicating that the connection between two synapses is weakened.

In addition, the Modified National Institute of Standards and Technology (MNIST) handwritten digits dataset is used for the simulation. The simulation is performed using MATLAB and a three-layer network (one hidden layer) with 400 input neurons, corresponding to the input black and white images with 20 × 20 pixels encoded from the MNIST dataset, and 100 hidden neurons. The pattern recognition results (0–9) are presented through the 10 output neurons (Fig. 8a). The inner product between the input neuron signals and the corresponding synaptic weights obtained from experimental measurements is computed by the first synapse array and transmitted to the hidden neuron layer. A similar computation is performed between the hidden and output layers via the second synapse array, completing one training epoch. During each epoch, the artificial neural network (ANN) is trained using 60[thin space (1/6-em)]000 randomly selected images from the MNIST training dataset, and its recognition accuracy is evaluated on a separate test subset (detailed analysis in SI).9 Fig. 8b shows the device structure of the device D2. Our synaptic device achieves a validation accuracy of 95.5%, closely approaching the ideal software-based implementation, which attains a validation accuracy of 93.5%, in the pattern recognition task (Fig. 8c and d).


image file: d5nh00562k-f8.tif
Fig. 8 (a) Illustration of a hardware-based neural network configuration employing memristive synapses for real-time recognition of handwritten digits, mimicking brain-inspired computation for efficient pattern classification. (b) Schematic of the D2(Nb/NbOx/V2O5/Ni). (c) Accuracy versus epoch plot showing the learning progression during image recognition. (d) Comparison of the accuracy of the ideal and D2 device.

The schematic illustration of the flexible V2O5-based ReRAM devices architecture highlights its mechanical adaptability (Fig. 9a). The device's mechanical flexibility has been tested in three bending modes: tensile, compressive, and no bending, which shows no change in IV characteristics, as well as LTP and LTD (Fig. 9b and c). As presented in Fig. 9d, the device exhibited one bending mode that maintains stable performance around 1300 cycles between LRS and HRS, thus confirming the futuristic memory application of flexible RRAM. To further investigate angular stability, the device was bent at multiple angles (30° → 45° → 60° → 90° → 120° → 150°), with no observable degradation in either the LRS or HRS, as presented in Fig. 9e. The measurement setup used for this angular flexibility assessment is schematically depicted in Fig. 9f. Furthermore, cyclic bending tests were performed to probe under mechanical stress. The performance metrics summarized in Table 2 confirm that the Nb/NbOx/V2O5/Ni-based device achieves reliable resistive switching along with outstanding mechanical resilience and data stability, making it a strong candidate for flexible neuromorphic applications.


image file: d5nh00562k-f9.tif
Fig. 9 (a) Schematics of the present flexible RRAM device. (b) Gradual characteristics of heterostructure (D2). (c) LTP and LTD of RRAM devices (D2). (d) one bending cycle. (e) Endurance characteristics at various bending angles (D2). (f) Digital photographs showing the device at different bending angles.
Table 2 Comparison of the present active switching layer with the previously reported data for RRAM applications
Active layer OFF/ON Ratio Voltage (V) Flexibility Endurance Synaptic measurement SET/RESET Ref.
VO2 60 −6 to 6 No 100 No Both abrupt 59
VOx >100 −0.4 to 0.4 No 10 No Both abrupt 60
SrVOx >100 −1 to 2 No No Abrupt/Gradual 61
V2O5 350 −1 to 1 Yes 500 No Both abrupt 55
V2O5/WO3 −8 to 8 No Yes Both Gradual 62
NbOx/V2O5 5 × 102 −1.5 to 1.5 Yes 1300 Yes Both Gradual This work


4. Conclusions

In conclusion, this research demonstrates the successful realization of bipolar resistive switching and synaptic functionalities in V2O5-based flexible Re-RAM devices, namely, D1(Nb/V2O5/Ni) and D2(Nb/NbOx/V2O5/Ni) structures, fabricated via DC magnetron sputtering. Device D1 exhibits a sharp SET and a gradual RESET behavior. In contrast, device D2 demonstrates fully gradual SET and RESET transitions, enabling multilevel RS with an enhanced memory window. The superior performance of D2 is attributed to incorporating an NbOx interfacial layer, which modulates oxygen vacancy dynamics and stabilizes filamentary conduction. Moreover, D2 effectively mimics essential synaptic functionality, including LTP, LTD, SRDP, and STDP, making it highly promising for neuromorphic computing. The conduction mechanism follows a combination of Ohmic and SCLC processes, consistent with the behavior of oxygen vacancy-mediated switching. Mechanical flexibility is validated through stable performance over 1300 bending cycles, confirming excellent structural integrity and endurance. These results highlight the potential of V2O5-based flexible memory devices as scalable, wearable, and energy-efficient electronics for high-performance neuromorphic computing applications.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

Data will be made available on request.

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: provides additional details on the Artificial Neural Network (ANN) simulation, the corresponding confusion matrix, and discussions related to neuromorphic computing. It also includes an analysis of device-to-device stability and the fitting of the IV characteristics based on the Richardson–Schottky equation. See DOI: https://doi.org/10.1039/d5nh00562k.

Acknowledgements

We appreciate the financial assistance given by the Defense Research and Development Organization (DRDO) under EP & IPR 2025 with reference no. ERIP/ER/202311006/M/01/1852. We also appreciate the financial assistance given by the Science and Engineering Research Board (SERB), Department of Science and Technology (DST), India, with reference No. CRG/2020/005265. Davinder Kaur sincerely acknowledges the Shastri Indo-Canadian Institute, India, for awarding the SMP fellowship. Kumar Kaushlendra sincerely acknowledges MHRD for the Senior Research Fellowship (SRF) award.

References

  1. M. Awais, N. Othman, M. D. Shafiq, F. Zhao and K. Y. Cheong, Small, 2025, 2504260 CrossRef CAS PubMed.
  2. R. Sun, Z. Hou, Q. Chen, B. Zhu, C. Zhu, P. Huang, Z. Hu, L. Zhen, F. Zhou, C. Xu and J. Qin, Adv. Mater., 2025, 37, 2409017 CrossRef CAS PubMed.
  3. H. Na and S. Kim, Adv. Funct. Mater., 2025, 2500956 CrossRef CAS.
  4. H. Sun, H. Tian, Y. Hu, Y. Cui, X. Chen, M. Xu, X. Wang and T. Zhou, Adv. Sci., 2024, 11, 2406242 CrossRef PubMed.
  5. G. Kim, D. Yoo, H. So, S. Park, S. Kim, M. J. Choi and S. Kim, Mater. Horiz., 2024, 12, 915–925 Search PubMed.
  6. H. Paik, J. Lim, H. Seo, T. K. Kim, J. Shin, H. Song, D. G. Kim, W. Lee, D. S. Kwon and C. S. Hwang, Mater. Horiz., 2025, 12, 7305–7317 RSC.
  7. S.-M. Chen, L.-C. Shih, J.-C. Gao, S.-X. You, K.-T. Chen, P.-L. Lin, K.-S. Hsu, C.-C. Chen, W.-L. Chen and J.-S. Chen, Nanoscale Horiz., 2025, 10, 2068–2079 RSC.
  8. S. Shankhdhar, D. Arora, H. Rahman, R. Kumar, B. Ranjan, K. Kaushlendra and D. Kaur, Nano Energy, 2025, 133, 110441 CrossRef CAS.
  9. A. K. Jena, M. C. Sahu, K. U. Mohanan, S. K. Mallik, S. Sahoo, G. K. Pradhan and S. Sahoo, ACS Appl. Mater. Interfaces, 2023, 15, 3574–3585 CrossRef CAS PubMed.
  10. R. Khan, N. U. Rehman, R. Thangappan, A. Saritha and S. Sangaraju, Nanoscale, 2025, 17, 11152–11190 RSC.
  11. S. M. Chen, L. C. Shih, J. C. Gao, S. X. You, K. T. Chen, P. L. Lin, K. S. Hsu, C. C. Chen, W. L. Chen and J. S. Chen, Nanoscale Horiz., 2025, 10, 2068–2079 RSC.
  12. A. Kumar, S. Pawar, S. Sharma and D. Kaur, Appl. Phys. Lett., 2018, 112, 262106 CrossRef.
  13. K. Kaushlendra and D. Kaur, ACS Appl. Electron. Mater., 2023, 5, 3362–3372 CrossRef CAS.
  14. G. Hu, Z. Yu, H. Qu, Y. Yuan, D. Li, M. Zhu, J. Guo, C. Xia, X. Wang, B. Wang, G. Ma, H. Wang and W. Dong, Appl. Phys. Lett., 2024, 124, 142109 CrossRef CAS.
  15. X. Chen, X. Wang, Y. Pang, G. Bao, J. Jiang, P. Yang, Y. Chen, T. Rao and W. Liao, Small Methods, 2023, 7, 2201156 CrossRef CAS.
  16. Y. F. Jiang, H. Y. Peng, Y. Cai, Y. T. Xu, M. Y. Fu, M. Feng, B. W. Wang, Y. Q. Wang, Z. Guan, B. Bin Chen, N. Zhong, C. G. Duan and P. H. Xiang, Mater. Horiz., 2025, 12, 6928–6937 RSC.
  17. Y. Wang, H. Liu, W. Wu, G. Yin, M. Zhang and J. Su, J. Mater. Chem. C, 2025, 13235–13245 RSC.
  18. X. Liu, J. Chen, Y. Wang, S. Han and Y. Zhou, Adv. Funct. Mater., 2021, 31, 2004733 CrossRef CAS.
  19. P. P. Patil, S. S. Kundale, S. V. Patil, S. S. Sutar, J. Bae, S. J. Kadam, K. V. More, P. B. Patil, R. K. Kamat, S. Lee and T. D. Dongale, Small, 2023, 19, 2303862 CrossRef CAS.
  20. Z. Cao, B. Sun, G. Zhou, S. Mao, S. Zhu, J. Zhang, C. Ke, Y. Zhao and J. Shao, Nanoscale Horiz., 2023, 8, 716–745 RSC.
  21. T. Guo, K. Pan, Y. Jiao, B. Sun, C. Du, J. P. Mills, Z. Chen, X. Zhao, L. Wei, Y. N. Zhou and Y. A. Wu, Nanoscale Horiz., 2022, 7, 299–310 RSC.
  22. M. Ismail, U. Chand, C. Mahata, J. Nebhen and S. Kim, J. Mater. Sci. Technol., 2022, 96, 94–102 CrossRef CAS.
  23. R. Deb, M. G. Nair, U. Das and S. R. Mohapatra, J. Mater. Chem. C, 2023, 11, 7629–7640 RSC.
  24. R. Islam, H. Li, P.-Y. Chen, W. Wan, H.-Y. Chen, B. Gao, H. Wu, S. Yu, K. Saraswat and H.-S. Philip Wong, J. Phys. D: Appl. Phys., 2019, 52, 113001 CrossRef.
  25. K. Chatterjee, J. Stanley, V. P. Dravid, R. dos Reis and S. E. Skrabalak, Appl. Phys. Lett., 2023, 123, 154101 CrossRef CAS.
  26. Y. Zhu, M. Chen, H. Lu, P. Mi, D. Luo, Y. Wang, Y. Liu, R. Xiong and H. Wang, Appl. Phys. Lett., 2024, 124, 063504 CrossRef CAS.
  27. Z. Wang, M. Yin, T. Zhang, Y. Cai, Y. Wang, Y. Yang and R. Huang, Nanoscale, 2016, 8, 14015–14022 RSC.
  28. M. Ismail, U. Chand, C. Mahata, J. Nebhen and S. Kim, J. Mater. Sci. Technol., 2022, 96, 94–102 CrossRef CAS.
  29. Z. Wang, S. Joshi, S. E. Savel’ev, H. Jiang, R. Midya, P. Lin, M. Hu, N. Ge, J. P. Strachan, Z. Li, Q. Wu, M. Barnell, G. L. Li, H. L. Xin, R. S. Williams, Q. Xia and J. J. Yang, Nat. Mater., 2017, 16, 101–108 CrossRef CAS.
  30. J. H. Ryu and S. Kim, Chaos, Solitons Fractals, 2020, 140, 110236 CrossRef.
  31. T. Wan, B. Qu, H. Du, X. Lin, Q. Lin, D. W. Wang, C. Cazorla, S. Li, S. Liu and D. Chu, J. Colloid Interface Sci., 2018, 512, 767–774 CrossRef CAS.
  32. K. Gao, B. Sun, B. Yang, Z. Cao and Y. Cui, Appl. Mater. Today, 2025, 43, 102628 CrossRef.
  33. Y. Qiao, F. Wang, W. Guo, Z. He, L. Yao, J. Li, N. Sun, Y. Wang and F. Wang, J. Alloys Compd., 2023, 967, 171712 CrossRef CAS.
  34. B. Ranjan and D. Kaur, ACS Appl. Mater. Interfaces, 2024, 16, 14890–14901 CrossRef CAS.
  35. S. Yin, C. Song, Y. Sun, L. Qiao, B. Wang, Y. Sun, K. Liu, F. Pan and X. Zhang, ACS Appl. Mater. Interfaces, 2019, 11, 43344–43350 CrossRef CAS.
  36. S. Sharma, K. Kumar, K. Kaushlendra and D. Kaur, ACS Appl. Electron. Mater., 2024, 6, 875–886 CrossRef CAS.
  37. V. R. Nallagatla, T. Heisig, C. Baeumer, V. Feyer, M. Jugovac, G. Zamborlini, C. M. Schneider, R. Waser, M. Kim, C. U. Jung and R. Dittmann, Adv. Mater., 2019, 31, 1903391 CrossRef CAS PubMed.
  38. K. Kaushlendra and D. Kaur, ACS Appl. Electron. Mater., 2024, 6, 4548–4559 CrossRef CAS.
  39. Q. Tian, X. Shan, J. Bian, Y. Cheng and J. Zheng, J. Mater. Chem. C, 2025, 12046–12065 RSC.
  40. Z. Wan, H. Mohammad, Y. Zhao, C. Yu, R. B. Darling and M. P. Anantram, J. Appl. Phys., 2018, 124, 105301 CrossRef.
  41. L. Gao, Q. Liao, X. Zhang, X. Liu, L. Gu, B. Liu, J. Du, Y. Ou, J. Xiao, Z. Kang, Z. Zhang and Y. Zhang, Adv. Mater., 2020, 32, 1906646 CrossRef CAS.
  42. X. Zhang, Q. Liao, S. Liu, Z. Kang, Z. Zhang, J. Du, F. Li, S. Zhang, J. Xiao, B. Liu, Y. Ou, X. Liu, L. Gu and Y. Zhang, Nat. Commun., 2017, 8, 1–8 CrossRef.
  43. R. Yang, Chin. Phys. B, 2020, 29, 097305 CrossRef CAS.
  44. R. Ge, X. Wu, L. Liang, S. M. Hus, Y. Gu, E. Okogbue, H. Chou, J. Shi, Y. Zhang, S. K. Banerjee, Y. Jung, J. C. Lee and D. Akinwande, Adv. Mater., 2021, 33, 2007792 CrossRef CAS.
  45. G. Zhang, Y. Xu, S. Yang, S. Ren, Y. Jiao, Y. Wang, X. Ma, H. Li, W. Hao, C. He, X. Liu and J. Zhao, Nano Energy, 2023, 106, 108074 CrossRef CAS.
  46. B. Ranjan and D. Kaur, Small, 2024, 20, 2307723 CrossRef CAS.
  47. R. Kumar, B. Ranjan and D. Kaur, J. Energy Storage, 2024, 84, 110866 CrossRef.
  48. H. Rahman, G. K. Sharma, P. Singh and D. Kaur, ACS Appl. Electron. Mater., 2024, 6, 2400–2412 CrossRef CAS.
  49. R. Kumar, B. Ranjan, K. Kumar, S. Shankhdhar and D. Kaur, ACS Appl. Nano Mater., 2024, 7, 7663–7673 CrossRef CAS.
  50. K. Kaushlendra, P. Kumar, D. Arora, B. Ranjan and D. Kaur, Ceram. Int., 2023, 49, 9543–9550 CrossRef CAS.
  51. H. Raha, N. Dey and P. Kumar, J. Energy Storage, 2025, 128, 117177 CrossRef.
  52. B. Ranjan and D. Kaur, J. Energy Storage, 2023, 71, 108122 CrossRef.
  53. B. Sharmila, P. Divyashree and P. Dwivedi, Sens. Actuators, A, 2024, 367, 114979 CrossRef.
  54. S. G. Choi, H. J. Seok, S. Rhee, D. Hahm, W. K. Bae and H. K. Kim, J. Alloys Compd., 2021, 878, 160303 CrossRef CAS.
  55. B. P. Yalagala, P. Sahatiya, C. S. R. Kolli, S. Khandelwal, V. Mattela and S. Badhulika, ACS Appl. Nano Mater., 2019, 2, 937–947 CrossRef CAS.
  56. K.-C. Chuang, C.-Y. Chu, H.-X. Zhang, J.-D. Luo, W.-S. Li, Y.-S. Li and H.-C. Cheng, IEEE J. Electron Devices Soc., 2019, 7, 589–595 CAS.
  57. S. Zhu, B. Sun, G. Zhou, T. Guo, C. Ke, Y. Chen, F. Yang, Y. Zhang, J. Shao and Y. Zhao, ACS Appl. Mater. Interfaces, 2023, 15, 5420–5431 CrossRef CAS.
  58. S. K. Mallik, R. Padhan, M. C. Sahu, S. Roy, G. K. Pradhan, P. K. Sahoo, S. P. Dash and S. Sahoo, ACS Appl. Mater. Interfaces, 2023, 15, 36527–36538 CrossRef CAS PubMed.
  59. J. Li, X. Zhou, L. Xu, J. Wang, B. Wu and C. Wang, Mater. Res. Express, 2022, 9, 035003 CrossRef CAS.
  60. J. Ryu, K. Park, D. P. Sahu and T. S. Yoon, ACS Appl. Mater. Interfaces, 2024, 16, 26450–26459 CrossRef CAS.
  61. T.-J. Lee, S.-K. Kim and T.-Y. Seong, Sci. Rep., 2020, 10, 5761 CrossRef CAS PubMed.
  62. B. Sharmila and P. Dwivedi, ACS Appl. Electron. Mater., 2025, 7, 3191–3200 CrossRef.

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